You are on page 1of 361

Fundamentals of Dynamics

and Control of Space


Systems

Krishna Dev Kumar


ii

Dr. Krishna. D. Kumar


Professor and Canada Research Chair in Space Systems
Department of Aerospace Engineering
Ryerson University
Toronto, Ontario
Canada M5B 2K3
Email: kdkumar@ryerson.ca
http://www.ryerson.ca/ kdkumar

Copyright c 2012 by Krishna Dev Kumar. All rights are reserved whether
the whole or part of the material is concerned, especially the right of transla-
tion, reprinting, reuse of illustrations, recitation, broadcasting, reproduction
on microfilm or any other ways, and storage in data banks.

Cover page: Full view of the International Space Station as photographed


from the Space Shuttle Discovery during the STS-114 Return to Flight mis-
sion, following the undocking of the two spacecraft. (Courtesy of NASA)
Fundamentals of Dynamics
and Control of Space
Systems

Krishna Dev Kumar


Professor and Canada Research Chair in Space Systems
Department of Aerospace Engineering
Ryerson University
Toronto, Canada
iv
Preface

This book is an outcome of the lectures delivered by the author for engi-
neering students at Kyushu University, Japan, Korea Advanced Institute of
Science and Technology, South Korea, and Ryerson University, Canada. The
text started when the author was at Kyushu University, Japan. The author
has been doing research on the subject of “Dynamics and Control of Space
Systems” for the last fifteen years. During the course of research and teach-
ing this subject, the author realized that there is an earnest need of a text
which may provide basic understanding of inherent complexity in modelling
space systems including environmental forces and torques. This text fulfils
this goal through a step by step approach that explains each concept in a
simple way with mathematical details and simple notations.
This textbook is written for engineers and students pursuing study or
research on the dynamics and control of space systems. The author has used
all contents of this textbook in a core course on Space Systems (one semester,
four hours a week). This text assumes that the reader is well acquainted with
elementary calculus and linear algebra. Enough care has been taken so that
the reader may attempt different types of problems. Each concept has been
discussed with sufficient mathematical background supported by examples.
A complete set of problems has been added to the end of every chapter. A
“Solution Manual” accompanying this book provides a solution to most
of the these exercise problems. In addition, a “Laboratory Manual” with
this textbook provides the practical aspects of designing space systems with
a focus on the construction of a Can-sized satellite.
During the course of writing I have received several suggestions and com-
ments from my colleagues and students. I especially thank my students Arad-
hana Choudhuri, Mike Alger, Geoffrey McVittie, Mike Tai, Kamran Shahid,
Antonio Mauro, and Tarunkumar Patel who have gone through previous
drafts of this book, and provided me with valuable suggestions.
Finally, I express my sincere appreciation to my parents, my wife, Annu,
and children, Diksha and Saurabh for their support and patience, without
which it would have been impossible to complete this book. This textbook
vi PREFACE

is dedicated to my parents and family for their inspiration and support.


In writing this textbook, I have tried my best to provide the best rele-
vant materials on the subject. I encourage and welcome all comments and
suggestions to make this book more relevant to the contents of your course
or research study. Please address your comments to kdkumar@ryerson.ca.

September 3, 2012 Krishna Kumar


Contents

Preface v

1 Introduction to Space Systems 1


1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Space System . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3 Images of Space Systems . . . . . . . . . . . . . . . . . . . . . 21
1.3.1 Earth Orbiting Missions . . . . . . . . . . . . . . . . . 21
1.3.2 Small Satellites . . . . . . . . . . . . . . . . . . . . . . 23
1.3.3 Communication Satellites . . . . . . . . . . . . . . . . 24
1.3.4 Astronomy Satellites . . . . . . . . . . . . . . . . . . . 26
1.3.5 Interplanetary Space Systems . . . . . . . . . . . . . . 27
1.3.6 Manned Orbiting Platforms . . . . . . . . . . . . . . . 29
1.3.7 Space Robotics . . . . . . . . . . . . . . . . . . . . . . 32
1.3.8 Launch and Deployment Systems . . . . . . . . . . . . 36
1.3.9 Future Space Systems . . . . . . . . . . . . . . . . . . 37
1.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

2 Kinematics, Momentum and Energy 39


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.2 Point Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.2.1 Position, Velocity and Acceleration . . . . . . . . . . . 40
2.2.2 Momentum . . . . . . . . . . . . . . . . . . . . . . . . 60

vii
viii CONTENTS

2.2.3 Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
2.3 Rigid Body . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
2.3.1 Defining Attitude . . . . . . . . . . . . . . . . . . . . . 71
2.3.2 Position, Velocity and Acceleration . . . . . . . . . . . 77
2.3.3 Momentum . . . . . . . . . . . . . . . . . . . . . . . . 78
2.3.4 Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
2.4 Flexible Body . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
2.4.1 Defining Deformation . . . . . . . . . . . . . . . . . . 95
2.4.2 Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
2.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
2.5.1 Kinematics: Point Mass . . . . . . . . . . . . . . . . . 110
2.5.2 Kinematics: Rigid Body . . . . . . . . . . . . . . . . . 112
2.5.3 Kinematics: Flexible Body . . . . . . . . . . . . . . . 114
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

3 Forces and Torques 123


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
3.2 Gravitational Force . . . . . . . . . . . . . . . . . . . . . . . . 124
3.2.1 Simplified Gravitational Force Model . . . . . . . . . . 125
3.2.2 Generalized Gravitational Force Model . . . . . . . . . 126
3.3 Aerodynamic Forces . . . . . . . . . . . . . . . . . . . . . . . 128
3.3.1 Simplified Aerodynamic Force Model . . . . . . . . . . 128
3.3.2 Free-Molecular Aerodynamic Force Model . . . . . . . 131
3.4 Solar Radiation Pressure Force . . . . . . . . . . . . . . . . . 133
3.4.1 Solar Radiation Pressure Force Model I . . . . . . . . 133
3.4.2 Solar Radiation Pressure Force Model II . . . . . . . . 137
3.4.3 Earth shadow . . . . . . . . . . . . . . . . . . . . . . . 138
3.5 Magnetic Field Torque . . . . . . . . . . . . . . . . . . . . . . 139
3.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
3.7 Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
3.7.1 Atmospheric Data . . . . . . . . . . . . . . . . . . . . 147
CONTENTS ix

3.7.2 International Geomagnetic Reference Field . . . . . . 151


Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

4 Dynamics I 159
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
4.2 Newton’s Laws of Motion . . . . . . . . . . . . . . . . . . . . 160
4.3 Kepler’s Laws of Orbit Motion . . . . . . . . . . . . . . . . . 162
4.4 Two-Body Motion . . . . . . . . . . . . . . . . . . . . . . . . 163
4.4.1 Angular Momentum Vector . . . . . . . . . . . . . . . 166
4.4.2 Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
4.4.3 Eccentricity Vector . . . . . . . . . . . . . . . . . . . . 167
4.4.4 Solution of Relative Motion . . . . . . . . . . . . . . . 168
4.5 Conic sections . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
4.5.1 Ellipse . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
4.5.2 Parabola . . . . . . . . . . . . . . . . . . . . . . . . . 173
4.5.3 Hyperbola . . . . . . . . . . . . . . . . . . . . . . . . . 174
4.6 Orbit Motion in Relation with Time . . . . . . . . . . . . . . 178
4.7 Describing Orbit Motion in Space . . . . . . . . . . . . . . . . 183
4.8 Relating Orbital Elements with Position and Velocity Vectors 186
4.8.1 Determination of Orbital Elements from Position and
Velocity Vectors . . . . . . . . . . . . . . . . . . . . . 186
4.9 Orbital Perturbations . . . . . . . . . . . . . . . . . . . . . . 190
4.9.1 Non-Keplerian Orbit Motion . . . . . . . . . . . . . . 190
4.9.2 Angular Momentum Vector . . . . . . . . . . . . . . . 191
4.9.3 Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . 192
4.9.4 Eccentricity Vector . . . . . . . . . . . . . . . . . . . . 192
4.9.5 Equations of Motion (Orbital Elements) . . . . . . . . 193
4.9.6 Earth’s Oblateness . . . . . . . . . . . . . . . . . . . . 201
4.9.7 Planetary Gravitational Perturbations . . . . . . . . . 211
4.9.8 Aerodynamic Drag . . . . . . . . . . . . . . . . . . . . 218
4.10 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
x CONTENTS

5 Dynamics II 225
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
5.2 Euler Equations of Motion . . . . . . . . . . . . . . . . . . . . 225
5.3 Lagrange’s Equations of Motion . . . . . . . . . . . . . . . . . 226
5.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260

6 Mathematical Analysis and Simulation 263


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
6.2 Equilibrium Condition . . . . . . . . . . . . . . . . . . . . . . 264
6.3 Linearization . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
6.4 Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
6.5 Closed-form Solution . . . . . . . . . . . . . . . . . . . . . . . 270
6.6 Numerical Simulation . . . . . . . . . . . . . . . . . . . . . . 282
6.7 Maple . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 286
6.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 294
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 296

7 Control System 301


7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302
7.2 Linear Controllers . . . . . . . . . . . . . . . . . . . . . . . . 303
7.3 Nonlinear Controllers . . . . . . . . . . . . . . . . . . . . . . . 309
7.3.1 Fuzzy Logic Controllers . . . . . . . . . . . . . . . . . 309
7.3.2 Neural Networks Controllers . . . . . . . . . . . . . . 310
7.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316

8 Formation Flying 323


8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
8.2 Formation Classification . . . . . . . . . . . . . . . . . . . . . 325
8.3 Leader Satellite in Circular Orbit . . . . . . . . . . . . . . . . 326
CONTENTS xi

8.4 Leader Satellite in Elliptic Orbit . . . . . . . . . . . . . . . . 338


8.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345

Index 347
xii CONTENTS
Chapter 1

Introduction to Space
Systems

This chapter gives an overview of past, present, and future space systems,
highlighting the inherent complexity in modelling these systems for study-
ing system dynamics and control. The chapter starts with a brief history of
space systems: satellites, space stations, manned missions, and launch sys-
tems. Next, the anatomy and system architecture of a satellite are discussed.
Finally, the chapter concludes with a detailed summary of mission failures
and a discussion of the importance of dynamics and control in space systems
design.

1.1 Introduction
The Space Era began with the launch of Sputnik by the former USSR in
October 4, 1957. This successful mission was followed by the launch of sev-
eral satellites by the USSR (now Russia) and USA. Historically significant
satellites are summarized in Table 1.1. Early satellites typically weighted 100
kg with sizes around 1 m (Example Sputnik, see Table 1.1 and Fig. 1.12-
1.13). In the 80’s and 90’s, the trend moved towards larger and more complex
satellites with weights typically ranging from 1,000 kg to 10, 000 kg. Two
examples of this class of satellites are the Hubble space telescope (on-orbit
mass: 11,000 kg, size: length of 13.3 m and max. dia. of 4.3 m, Orbit: 600
km, Period: 96-97 min, Inclination: 28.5 deg, Launch: April 24, 1990, see
Fig. 1.9-1.10) and the Chandra X-ray observatory (on-orbit mass: 4800 kg,
size: wing-span of 19.5m and length of 11.8 m (with sun-shade open), Orbit:
10,000 km x 140,161 km, Inclination: 28.4 deg, Period: 64.3 hr, Launch: July
23, 1999, Fig. 1.1-1.6).
2 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

Table 1.1 Major Milestones in Space History.


Satellite Origin Year Remarks
Sputnik-1 USSR 1957 First artificial satellite
(Orbit: 500 miles, Mass: 83 kg,
size: 58 cm dia. ball)
Luna 1 USSR January 2, 1959 First lunar flyby spacecraft
(Mass: 361 kg)
Luna 2 USSR September 12, 1959 First spacecraft to impact the surface
of the moon on September 14, 1959
(Mass: 387 kg)
TIROS1 USA April 1, 1960 First weather satellite
Vostok1 USSR April 12, 1961 First manned mission to space
Crew member: Yuri Gagarin
(Mass: 2270 kg,
Size: length 2.3 m x dia. 2.3 m)
Lambda 1 USA April 18, 1962 First Astronomy satellite
(Dry mass: 1500 kg,
size: 6 m long and 1.5 m dia.)
Telstar USA 1962 First active, direct relay
communication satellite
Syncom-2 USA 1963 First geosynchronous
communication satellite
(Dry mass 145.7 kg)
Apollo 11 USA July 16, 1969 First manned lunar landing
on July 20, 1969
(Mission: July 16-24, 1969)
(Mass 43,811 kg)

1
TIROS stands for Television InfraRed Observational Satellite.
1.1. INTRODUCTION 3

The cost of a typical launch to low earth orbit(up to 185 km) is US$10,000
per kilogram[2]. The total cost for Geostationary Operational Environmental
Satellite (GOES) (Launch: May 2000) was US$290 million while the Missile
Warning Satellite (Launch: May 2000) costed US$682 million. This pro-
hibitive expense coupled with recent advances in miniaturization and manu-
facturing techniques and a commercial/educational interest in space systems
resulted in the design and launch of small satellites (Figs. 1.16 and 1.17).
This trend towards miniaturization and lowering of launch cost per satel-
lite (a CubeSat (size: 10x10x10 cm, mass: 1 kg) costs around US$50,000
to launch) has prompted active research even into developing Femto class
satellites (mass less than 100 gm).
The number of satellites has grown enormously over time (more than 3000
satellites as of year 2000), and therefore, we can classify these satellites into
groups based on mass (Table 1.2), orbit (Table 1.3), and functionality (Table
1.4).
4 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

Table 1.2 Classifications of Satellites Based on Satellite Mass.


Satellites Wet Mass Example
(Including Fuel)
Large satellite > 1000 kg Galileo
(Mass 2222 kg, Launch Oct. 18, 1989,
Jupiter orbiter)
Medium sized satellite 500-1000 kg NEAR1
(Mass 805 kg, Launch Feb. 17, 1996)
Minisatellite 100-500 kg Lunar Prospector
(Mass 295 kg, Launch Jan. 6, 1998)
Microsatellite 10-100 kg MOST
(Mass 53 kg, size 0.7x0.7x0.3m,
Orbit: 820 km, Period: 100 min,
June 30, 2003)
Nanosatellite 1-10 kg ASU-OSCAR 37
(Mass 6 kg, Launch Jan. 27, 2000,
Orbit: 799x746 km,
Inclination: 100.19 deg)
Picosatellite 0.1-1 kg DARPA Picosat 7 and 8 (MEPSI-2)
(Mass: 250g, Size 100x750x250mm,
Orbit: 511x539km,
Inclination: 97.8deg,
Year: 2000)
Femtosatellite < 100 g Not yet flown

1
Near Earth Asteroid Rendzvous. 2 Microvariability and Oscillations of Stars
1.1. INTRODUCTION 5

Table 1.3 (a) Classifications of Satellites Based on Satellite Orbit.


Satellites Orbit Example
(orbital period)
Low Earth Orbit 200-1200 km Iridium satellite constellations
(LEO) (90 min - 2hrs) (Orbit: 780 km,
Near Polar Orbit,
Year: 1998, 66 satellites)
Medium Earth Orbit 12,000-20,000 km Global Positioning System
(MEO) or Intermediate Circular (2 - 12 hrs) (GPS)
Orbit (ICO) (Orbit: 20,200 km,
Near Circular Orbit,
Inclination 55 degree,
Period: 12 hrs, 820 kg,
24 satellites (now 29 satellites)
in six orbital planes
(complete in 1994)
6 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

Table 1.3 (b) Classifications of Satellites Based on Satellite Orbit.


Satellites Orbit Example
(orbital period)
Geosynchronous Orbit1 35,786 km Intelsat 7A series
(GEO) (24 hrs) (Year 1995, Mass 4180 kg)

Molniya Orbit Higly elliptic Molniya 3-47


Inclination: 63.4 deg (Orbit: 900 km x 39000 km,
Mass: 1,600 kg,
Size: Length 4.4 m x
Maximum Dia. 1.4 m,
Span: 8.20 m)

Sunsynchronous Orbit2 MEO or LEO Radarsat


(Year 1995, Orbit: 793x821 km,
Inclination: 98.6 deg,
Mass: 2,713 kg,
Size: projected length 1.5 m x
max. dia. 1.2 m,
Span: 15 m )

1
Geostationary Orbit(GSO) is a zero inclination geosynchronous orbit. 2 The
Sunsynchronous orbit is a special case of a polar orbit that precesses at
exactly 1 degree per day to remain in the same local time plane as the Earth
rotates about the Sun
1.1. INTRODUCTION 7

Table 1.4 Classifications of Satellites Based on Satellite Functions.


Satellite Orbit Example
Astronomy Highly Inclined Hubble Space Telescope(1990)
Eccentric Orbit
Atmospheric Studies GEO, MEO, LEO NASA’s Polar Satellite(1996)
Communication GEO, Telstar(1962)
Molniya orbit,
Low polar orbit
Remote Sensing Sun synchronous Radarsat(1995),
LEO LANDSAT 7(1999)
Navigation MEO GPS
Search and Rescue MEO, LEO Cospas(1982)
Space Exploration Galileo(1989)
Reconnaissance LEO, Sunsynchronous Lacrosse(1988)
Weather GEO and Polar Tiros I(1960)

In addition to launch of autonomous satellites, long duration manned


missions have been undertaken. These missions include Skylab, Mir space
station, and the International space station (ISS). The history of manned
skylab missions is summarized in Table 1.5. Following the skylab missions,
Mir space stations was constructed by the USSR in 1996. The American space
station Freedom (Alpha) remained a paper project and was later integrated
into the design for the ISS. Table 1.6 summarizes Mir and ISS space stations.
8 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

Table 1.5 History of Manned Skylab Missions.


Skylab 2 Skylab 3 Skylab4
Launch 05-25-73 07-28-73 11-16-73
Launch Vehicle Saturn 1B Saturn 1B Saturn 1B
Orbit 268.1 x 269.5 Miles 268.1 x 269.5 Miles 268.1 x 269.5 Miles
Inclination 50 deg 50 deg 50 deg
Period 93 min 93 min 93 min

Mission- 28 days 59 days 84 days


Duration 49 min 11 hrs 9 min 1 hr 16 min

Table 1.6 Mir and ISS space stations.


Mir ISS
Origin USSR USA, Russia, Japan and
Canada (CSA), and 10 European
countries (represented by ESA)
Launch February 19, 1986 November 20, 1998
Fully constructed 1996 Estimated 2010
Reentry March 23, 2001
Orbit 385 x 393 km 352.8 x 354.2 km
Period 89.1 min 91.61 min
Inclination 51.6 deg 51.64 deg
Mass 124,340 kg 183,283 kg (August 28, 2005)
Size Length: 13.1 m Length: 44.5 m
(along core) (along core)
Diameter: 4.2 m Height: 27.5 m

Manned missions started with the launch of Vostok 1 in 1961 (see Table
1.1) which orbited the Earth once (orbit: 169 x 315 km). The next significant
milestones were the Apollo missions. Apollo-8 was the first manned lunar fly-
around and safe Earth return (December 21-27, 1968) and Apollo-11 (July
1.1. INTRODUCTION 9

16-24, 1969) was the first manned lunar landing on July 20, 1969. The
landing site was Mare Tranquillitatis at latitude 0o 670 N and longitude 23o 490 .
Apollo 11’s crew members were Neil A. Armstrong, Edwin E. Aldrin, Jr., and
Michael Collins. The Apollo project was conducted untill 1975 with a total of
six successful landings. As of the year 2006, there have not been any further
manned missions beyond low earth orbit. Canada has also made a significant
contribution to manned space missions aboard the Space shuttle and the ISS
in the form of the Canadaarm (see Table 1.7 and Figs. 1.31-1.33).
All satellites and space missions require launchers to reach desired orbits.
Some of the launchers currently in use include: Ariane series (ESA), Atlas
series (USA), Delta series (USA), Titan series (USA), Space Shuttle (USA),
Cosmos series (Russia), Pegasus (USA, air launch), Proton (Russia), Soyuz
(Russia), Zenit (privately owned, Sea Launch), Polar Satellite Launch Vehicle
(India), Geosynchronous Satellite Launch Vehicle (India), H-2A (Japan), and
Long March (China). Pictures of the Titan series (Fig. 1.39) and space
shuttle (Fig. 1.41) are included in Section 1.3.8.
10 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

Table 1.7 Details of Canadarm 1 and Canadarm 2


Canadarm 1 Canadarm 2
Launch November 13, 1981 (STS-2) April 2001 (STS-100)
Degrees of Freedom 6 7
Joints (shoulder: 2, elbow: 1, (shoulder: 3, elbow: 1,
wrists: 3) wrists: 3)
Change configuration
without moving hands.
Joint Rotation 160 deg elbow rotation Full joint rotation
Joints (7) rotate 540 degrees.
Larger range of motion
than a human arm.
Mass 410.5 kg 1,800 kg
Size Length: 15 m 17.6 m
Diameter: 33 cm 35 cm
(exterior dia. of
(composite boom)
Mass Handling 29,484 kg 116,000 kg
Capacity (design case handling (design case handling
payload) payload)
Upgraded to 266,000 kg
Speed of Unloaded: 60 cm/s Unloaded: 37 cm/s
Operations Loaded: 6 cm/s Loaded:
Station Assembly: 2 cm/s
EVA Support: 15 cm/s
Orbiter: 1.2 cm/s
Cameras 2 4 color cameras
(one on the elbow and (one at each side of the elbow,
one on the wrist) the other two on the Latching
End Effectors)
1.2. SPACE SYSTEM 11

1.2 Space System


A space system is comprised of the following subsystems:

• Body: Structures and Mechanisms


• Orbit Determination and Control Subsystem: Guidance, Navigation
and Control System (GN&C)
• Attitude determination and control subsystem (ADCS)
• Communication System
• Internal Computer: Command and Data Handling Subsystem (C&DH)
or Telemetry Tracking and Control Subsystem (TT&C)
• Power System
• Thermal Control Subsystem (TCS)
• Payload

We illustrate satellite subsystems with some real mission examples. Fig-


ure 1.1 shows the Chandra X-ray Observatory which is the largest and heavi-
est satellite (mass 4800 kg) ever launched on a Space Shuttle. In this satellite,
two different sets of thrusters are used; one for propulsion and the other for
momentum dumping (Fig. 1.3). The thermal control subsystem is comprised
of a cooling radiator, insulators, heaters and thermostats. The satellite re-
ceives approximately two kilowatts of electrical power from solar arrays; this
power is used for heaters, science instruments (Fig. 1.2), computers, trans-
mitters and other components. Some components of this satellite are shown
in Figs. (1.4)-(1.6).
12 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

Figure 1.1: Overview of systems on Chandra X-ray Observatory.

Figure 1.2: Science instruments onboard Chandra X-ray Observatory.


1.2. SPACE SYSTEM 13

Figure 1.3: Thrusters onboard Chandra X-ray Observatory. (Courtesy of


NASA)

Figure 1.4: Low gain antenna onboard Chandra X-ray Observatory. (Cour-
tesy of NASA)
14 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

Figure 1.5: Sunshade door of Chandra X-ray Observatory. (Courtesy of


NASA)

Figure 1.6: Camera onboard Chandra X-ray Observatory. (Courtesy of


NASA)
1.2. SPACE SYSTEM 15

Landsat 7, a remote sensing satellite, is illustrated in Fig. 1.7. Details of


the satellite’s system block diagram are shown in Fig. 1.8.

Figure 1.7: Landsat 7. (Courtesy of NASA)


16 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

Figure 1.8: System block diagram for Landsat 7. (Courtesy of NASA)


1.2. SPACE SYSTEM 17

Figure 1.9 shows an overview of the Hubble space telescope (HST). Sen-
sors and actuators onboard the HST are detailed in Fig. 1.10. Figure 1.11
illustrates sensors and actuators that comprise the ADCS subsystem onboard
the ASCA spacecraft.

Figure 1.9: Details of Hubble Space Telescope. (Courtesy of NASA)

Figure 1.10: HST sensors and actuators. (Courtesy of NASA)


18 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

Figure 1.11: Attitude control system of ASCA spacecraft. (Courtesy of


Japanese Space Agency)

If any of the above subsystems do not perform as desired, a system fail-


ure or even mission loss may occur. Some mission failures are summarized
in Table 1.8. Analyzing failures, most common cause was launch failure fol-
lowed by ADCS failure. Approximately 23% of all Guidance, Navigation, and
Control Anomalies were ADCS related[5]. ADCS anomalies were the single
most common cause in those missions that were a total loss. Therefore, the
study of dynamics and control of space systems is vital to the success of
any space mission. This study involves orbital and attitude dynamics. Or-
bital dynamics deals with point mass model of the satellite while attitude
dynamics required rigid and flexible body models. In addition to these mod-
els, environmental models (that include gravity, solar radiation, magnetic
fields, aerodynamic forces, and free molecular reaction forces) are needed for
accurate modelling and control of spacecraft. This text book presents a sys-
tematic and comprehensive study of dynamics and control as they apply to
space systems.
We begin the second chapter with Kinematics of space systems including
Momentum and Energy for point mass, and rigid and flexible body models.
Chapter 3 focuses on detailed treatment of forces and torques in a space en-
vironment. Orbital dynamics is discussed in Chapter 4: Dynamics I, followed
1.2. SPACE SYSTEM 19

by attitude dynamics and coupled orbit and attitude dynamics in Chapter 5:


Dynamics II. After the development of relevant theory, Chapter 6 presents
mathematical analysis and methods for numerical simulation. Various control
algorithms applicable to space systems are discussed in Chapter 7. Finally,
the textbook ends with an introduction to formation flying.
20 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

Table 1.8 Space Missions Failures.

Satellite Date Event


SOHO June 2003 High-gain antenna stuck, unknown reason
MSAT 1 4 May 2003 Loss of two power amplifiers
Nimiq 2 20 Feb 2003 Malfunction affects available power,
some transponders off
Thaicom 3 6 Feb 2003 Temporary outages caused by short circuit
in solar array drive mechanism
Radarsat 1 30 Dec 2002 Attitude control restored using torque rods
Radarsat 1 27 Nov 2001 Backup momentum wheel fails,
loss of attitude
EchoStar VIII Oct 2002 Second onboard thruster anomaly
MSG-1 17 Oct 2002 IOT: Failure of Solid State Power
Amplifier C (SSPA-C)
TDRS I 30 Sep 2002 Spacecraft reaches final orbit although
one propellant fuel tank did not properly
pressurize after launch
Nozomi 23 Sep 2002 Contact with spacecraft re-established
after knock-out by Coronal Mass Ejection
on 21 Apr 2002
EchoStar VIII Sep 2002 Onboard thruster anomaly
CONTOUR 15 Aug 2002 Contact with spacecraft lost, reason unknown
Globalstar 12 Aug 2002 Company announces recovery of satellites
”in many cases;” only two declared failed
Echostar VI Aug 2002 Three solar array strings lost since 2001
Echostar V Jul 2002 Two solar array strings lost since 2001
1.3. IMAGES OF SPACE SYSTEMS 21

1.3 Images of Space Systems


This section consists of various photographs and diagrams of important satel-
lites and space systems.

1.3.1 Earth Orbiting Missions

Figure 1.12: Sputnik 1 shown with four whip antennas. (Courtesy of NASA)

Figure 1.13: Sputnik 1 interior. (Courtesy of NASA)


22 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

Figure 1.14: A GPS satellite. (Courtesy of U.S. Army)

Figure 1.15: GPS satellite constellation. (Courtesy of U.S. U.S. Department


of Defense)
1.3. IMAGES OF SPACE SYSTEMS 23

1.3.2 Small Satellites

Figure 1.16: Mepsi-3 Picosatellite. (Courtesy of NASDA)

Figure 1.17: Artistic impression of DARPA picosat. (Courtesy of Aerospace


corporation)
24 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

1.3.3 Communication Satellites

Figure 1.18: Syncom 2 (1963), the world’s first geosynchronous satellite.


(Courtesy of NASA)

Figure 1.19: Syncom 2 inside its launch vehicle. (Courtesy of Hughes Space
and Communications)
1.3. IMAGES OF SPACE SYSTEMS 25

Figure 1.20: Telestar 1 (1962). (Courtesy of NASA)


26 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

1.3.4 Astronomy Satellites

Figure 1.21: Hubble Space Telescope. (Courtesy of NASA)

Figure 1.22: An artist’s illustration of the Chandra X-ray Observatory.


(Courtesy of NASA)
1.3. IMAGES OF SPACE SYSTEMS 27

1.3.5 Interplanetary Space Systems

Figure 1.23: Opportunity Rover. (Courtesy of NASA)

Figure 1.24: Mars 2001 Odyssey. (Courtesy of NASA)


28 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

Figure 1.25: Cassini interplanetary trajectory. (Courtesy of NASA)


1.3. IMAGES OF SPACE SYSTEMS 29

1.3.6 Manned Orbiting Platforms

Figure 1.26: Skylab. (Courtesy of NASA)

Figure 1.27: Mir Space Station.(Courtesy of NASA)


30 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

Figure 1.28: Elements of Mir Space Station docked with Space Shut-
tle.(Courtesy of NASA)

Figure 1.29: ISS as photographed from the Space Shuttle Discovery during
the STS-114 Return to Flight mission, following the undocking of the two
spacecraft.(Courtesy of NASA)
1.3. IMAGES OF SPACE SYSTEMS 31

Figure 1.30: ISS elements - as of April 2006. (Courtesy of NASA)


32 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

1.3.7 Space Robotics

Figure 1.31: Canada arm. (Courtesy of NASA)

Figure 1.32: Special Purpose Dexterous Manipulator (SPDM) of Canadarm2.


(Courtesy of NASA)
1.3. IMAGES OF SPACE SYSTEMS 33

Figure 1.33: Photo of Hurricane Emily with Canadarm2 taken on July 17,
2005 by astronaut John Phillips from onboard the International Space Sta-
tion.(Courtesy of NASA)

Figure 1.34: STS-114 Mission Specialist Stephen K. Robinson is attached to


a foot restraint on Canadarm2. (Courtesy of NASA)
34 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

Figure 1.35: European Robotic Arm.(Courtesy of ESA)

Figure 1.36: Sprint, shuttle tail, RMS end-effector, and the Earth during
Flight STS 87 in 1997.(Courtesy of NASA)
1.3. IMAGES OF SPACE SYSTEMS 35

Figure 1.37: The Engineering Test Satellite VII. (Courtesy of NASDA)

Figure 1.38: Engineering Test Satellite VII (ETS-VII) (Launched in 1997


by National Space Development Agency of Japan (NASDA)).(Courtesy of
JAXA)
36 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

1.3.8 Launch and Deployment Systems

Figure 1.39: Titan series of launch vehicle. (Courtesy of NASA)

Figure 1.40: Space Shuttle flight sequence. (Courtesy of NASA)


1.3. IMAGES OF SPACE SYSTEMS 37

Figure 1.41: Space Shuttle. (Courtesy of NASA)

1.3.9 Future Space Systems

Figure 1.42: Space Elevator. (Courtesy of NASA)


38 CHAPTER 1. INTRODUCTION TO SPACE SYSTEMS

1.4 Summary
In this chapter, an overview of the history of space systems is presented.
Classifications of various satellites based on mass, orbit, and functionality
are discussed. The anatomy of typical space systems is illustrated. Previous
space failures are presented followed by a discussion of the importance of
studying the dynamics and control of spacecraft. More information about
space systems and the history of spaceflight are available in the References.

References

1. Peter W. Fortescue and John P. W. Stark, editors, Spacecraft Systems


Engineer- ing, John Wiley & Sons, Chichester, 1991.
2. Wiley J. Larson and James R. Wertz, editors, Space Mission Analysis
and De- sign, Microcosm, Inc., Torrance, CA, second edition, 1995.
3. Larson, W. J. and Wertz, J. R. (Editors), Space Mission Analysis and
Design,Second edition, Microcosm, Inc., Torrance, CA, 1995.
4. Vincent L. Pisacane and Robert C. Moore, editors, Fundamentals of
Space Sys- tems, Oxford University Press, Oxford, 1994.
5. Robertson, B. and Stoneking, E., Satellite GN&C Anomaly Trends,”
AAS Guidance and Control Conference, 2003.
6. Satellite Encyclopedia. http://www.tbs-satellite.com/tse/online/
7. Wikipedia Encyclopedia. http://en.wikipedia.org/wiki/
8. NASA. http://www.nasa.gov/
9. Encyclopedia Astronautica. http://www.astronautix.com/
10. European space agency. http://www.esa.org/
11. Space News. http://www.space.com
12. http://chandra.harvard.edu.

Problem Set 1

1.1 List all the space missions that have failed due to a launch failure.
1.2 List all the space missions that have failed due to attitude determination
and control system failure.
Chapter 2

Kinematics, Momentum
and Energy

This chapter deals with kinematics, momentum, and energy of a system.


First we study a point-mass system model. Next, a rigid body model followed
by a flexible body model are discussed. For each of these models, position,
velocity, and acceleration are determined. The system linear and angular
momentums and potential and kinetic energies are derived.

2.1 Introduction
Understanding the kinematics of a system is vital for deriving its equations
of motion and designing a controller to achieve desired performance. The
system may comprise of a point mass, a rigid body or a flexible body or a
combination of these elements. Kinematics deals with the determination of
position, velocity, and acceleration of a point mass or an element on a rigid
or flexible body. In this chapter, we explain these aspects. In addition, the
system linear and angular momentums are derived. The potential energy due
to gravity gradient force as well as strain energy along with kinetic energy
are also explained.

2.2 Point Mass


We derive the inertial position, velocity and acceleration of the spacecraft.
Furthermore, the momentum as well as energy of the spacecraft are also
obtained.
40 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

2.2.1 Position, Velocity and Acceleration


In studying the motion of a spacecraft, we start with a point mass model
which is the simplest model. By point mass model, we mean the spacecraft
mass is uniformally distributed and therefore it can be represented by a point
which has no dimension. In fact, this point mass model is quite accurate for
studying the trajectory or orbital motion of the spacecraft. We next derive
the position, velocity, and acceleration of a spacecraft with reference to an
inertial reference frame and then for a non-inertial reference frame.
Inertial reference frame. The position of a spacecraft m is defined
~ (see Fig. 2.1).
with respect to a reference point, say O, in a vector form by R
~
The corresponding velocity, V and acceleration, ~a of the point mass m are

O
Y

Figure 2.1: Describing position of a spacecraft.

described by

~
~ = dR
V (2.1)
dt
~
dV ~
d2 R
~a = = 2 (2.2)
dt dt
where d()/dt denotes the differentiation of () with respect to time. In order
~ V~ , and ~a, we need to express these vectors in
to obtain physical values of R,
a particular coordinate system associated with a coordinate frame attached
to the point O. There are namely three coordinate systems (Fig. 2.2): Rect-
angular coordinates, Cylindrical coordinates, and Spherical coordinates.
The Rectangular Coordinates are also called Cartesian coordinates. The
corresponding coordinate frame attached with the point O is named as Carte-
sian coordinate frame. This frame, say O − XY Z, must have the following
properties:
2.2. POINT MASS 41

k K j1
V i1
k1
m
R
j Z
O φ J
Y
θ r
i
I
X

Figure 2.2: Coordinate systems.

(a) Orthogonal: All the axes should be orthogonal, mathematically stated


as,
Iˆ · Iˆ = Jˆ · Jˆ = K̂ · K̂ = 1 (2.3)
Iˆ · Jˆ = Jˆ · K̂ = K̂ · Iˆ = 0 (2.4)
ˆ J,
where I, ˆ and K̂ are unit vectors along X, Y , and Z axes, respectively.

(b) Right-handed: The coordinate axes should be right-handed, i.e.,


Iˆ × Jˆ = K̂, Jˆ × K̂ = I,
ˆ K̂ × Iˆ = Jˆ (2.5)
~ velocity V
In cartesian coordinates (X − Y − Z), the position R, ~ , and
acceleration ~a are expressed as
~ = X Iˆ + Y Jˆ + Z K̂
R (2.6)
~
~ = dR = Ẋ Iˆ + Ẏ Jˆ + Ż K̂
V (2.7)
dt
~
dV
~a = = Ẍ Iˆ + Ÿ Jˆ + Z̈ K̂ (2.8)
dt
where Ẋ = dX/dt, Ẍ = d2 X/dt2 with similar descriptions for other variables.
Note that the coordinate frame O − XY Z attached to the point 0 is a non-
rotating and non-accelerating frame and therefore, I= ˆ˙ J=
ˆ˙ K̂=0.
˙
This frame
is also called the Inertial reference frame and the position, velocity, and
acceleration with reference to this frame are named inertial position, inertial
velocity, and inertial acceleration.
~
In cylindrical coordinates (r − θ − Z), we can write the position vector R
as
~ = rî + Z K̂
R (2.9)
42 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

where O − îĵ k̂ is a rotating coordinate frame with rotation ω ~ =θ̇k̂. The unit
vector î is along the vector ~r, the unit vector ĵ is normal to ~r in the plane
containing ~r and θ, and the unit vector k̂ makes the right-handed triad,
i.e., î × ĵ = k̂, perpendicular to the plane containing î and ĵ (see Fig. 2.2).
Note in the case of the planar motion of a spacecraft, its position can be
defined by polar coordinates (r − θ).
The velocity V ~ of the spacecraft can be written as
~
dR
V~ = = ṙ î + rî˙ + Ż k̂ (2.10)
dt
where k̂ = K̂. Knowing î˙ = ω
~ × î and ω
~ = θ̇k̂, we get

î˙ = θ̇ k̂ × î = θ̇ ĵ (2.11)

Substituting î˙ in Eq. (2.10), we have the velocity vector as


~ = ṙ î + rθ̇ ĵ + Ż k̂
V (2.12)

Similarly, we can obtain acceleration ~a as


~
dV
~a = = r̈ î + ṙ î˙ + ṙ θ̇ĵ + rθ̈ ĵ + rθ̇ ĵ˙ + Z̈ k̂ (2.13)
dt
or
   
~a = r̈ − rθ̇2 î + rθ̈ + 2ṙθ̇ ĵ + Z̈ k̂ (2.14)

Note that in the case of orbital motion of a spacecraft, the orbital plane
remains fixed in the inertial space if there exists no external force; which will
be explained in the next section. And therefore, we have Z=Ż=Z̈=0. The
preceding equations reduce to
V~ = ṙ î + rθ̇ ĵ (2.15)
   
~a = r̈ − rθ̇2 î + rθ̈ + 2ṙθ̇ ĵ (2.16)

~ can be expressed as
In spherical coordinates (R − θ − φ), the position R
~ = Rî1
R (2.17)
~ The velocity and accel-
where î1 is along the direction of position vector R.
eration vectors are given by
V~ = Ṙî1 + Rθ̇cosφĵ1 + Rφ̇k̂1 (2.18)
 
  cosφ d 2
~a = R̈ − Rφ̇2 − Rθ̇2 cos2 φ î1 + (R θ̇) − 2Rθ̇φ̇sinφ ĵ1
R dt
 
1 d 2
(R φ̇) + Rθ̇2 sinφcosφ k̂1 (2.19)
R dt
2.2. POINT MASS 43

where unit vectors ĵ1 and k̂1 are along the directions of angles θ and φ,
respectively (see Fig. 2.2). For complete derivations of Eqs. (2.18) and
(2.19), refer to the textbook by Meriam1 .
Non-inertial reference frame. In general, most of the reference frames
are rotating or accelerating or both rotating and accelerating with respect
to an inertial reference frame. These frames are called non-inertial reference
frames. In order to apply Newton’s laws of motion, we need to obtain in-
ertial velocities and accelerations from non-inertial reference frames. Let us
consider an inertial reference frame, denoted by O − XY Z with I, ˆ J,
ˆ and K̂
unit vectors along X, Y , and Z axes, respectively (Fig. 2.3). A non-inertial
reference frame, denoted by S − xyz is specified with respect to the inertial
reference frame O − XY Z by the position vector ρ ~ and the rotation vector
~ . For a given position ~r, velocity ~r˙ , and acceleration ~r¨ of a body of mass m
ω
with respect to the S − xyz frame, we are required to determine its position
~ velocity R,
R, ~˙ and acceleration R ~¨ in the O − XY Z frame.

Z
z m
V
r
F
Inertial
y
Reference R S
Frame Non−inertial
ρ Reference
x Frame
O
Y

Figure 2.3: Inertial and non-inertial reference frames.

The position vector of the body of mass m with respect to S − xyz frame
can be written as

~rxyz = xî + y ĵ + z k̂ (2.20)

The position vector of the frame center S can be expressed with respect to
the inertial reference frame S − XY Z as

ρ~XY Z = ρX Iˆ + ρY Jˆ + ρZ K̂ (2.21)
1 Meriam, J. L., Dynamics, second edition, John Wiley & Sons, Inc., 1971, pp. 77-84
44 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

We can now write the position vector of the body of mass m with respect to
the O − XY Z frame as
~ XY Z = ρ~XY Z + ~rxyz = ρX Iˆ + ρY Jˆ + ρZ K̂ + xî + y ĵ + z k̂
R (2.22)

Differentiating with respect to time leads to

~˙ XY Z =ρ̇X Iˆ + ρ̇Y Jˆ + ρ̇Z K̂ + ρX Iˆ˙ + ρY Jˆ˙ + ρZ K̂˙


R
˙
+ ẋî + ẏ ĵ + ż k̂ + xî˙ + y ĵ˙ + z k̂ (2.23)

Knowing

˙ ˙ ˙
Iˆ = Jˆ = K̂ = 0 (as I, ˆ J,
ˆ and K̂ are unit vectors along
inertial reference frame)
(2.24)
˙
î˙ = ω
~ × î, ĵ˙ = ω
~ × ĵ, k̂ = ω
~ × k̂

yields

~˙ XY Z = ρ̇x Iˆ + ρ̇y Jˆ + ρ̇z K̂ + ẋî + ẏ ĵ + ż k̂ + ω


  
R ~ × xî + y ĵ + z k̂ (2.25)

~˙ XY Z = ρ̇x Iˆ + ρ̇y Jˆ + ρ̇z K̂ and ~r˙xyz = ẋî + ẏĵ + ż k̂, we have


Writing ρ

~˙ XY Z = ρ
R ~˙ XY Z + ~r˙xyz + ω
~ × ~rxyz (2.26)

This equation presents an inertial velocity from a given position and velocity
in a non-inertial frame.
Next we derive an expression for inertial acceleration. Differentiating Eq.
(2.25) with respect to time and then applying Eqs. (2.24) lead to

~¨ XY Z = ρ̈x Iˆ + ρ̈y Jˆ + ρ̈z K̂ + ẍî + ÿĵ + z̈ k̂


 
R
~ × ~r˙xyz + ω ~ × ~r˙xyz + ω
~˙ × ~rxyz + ω

+ω ~ × ~rxyz (2.27)

~˙ XY Z = ρ̈x Iˆ + ρ̈y Jˆ + ρ̈z K̂ and ~r¨xyz = ẍî + ÿ ĵ + z̈ k̂, we have


Writing ρ

~¨ XY Z = ρ~¨XY Z + ~r¨xyz + 2(~


R ω × ~r˙xyz ) + ω
~ × (~ ~˙ × ~rxyz
ω × ~rxyz ) + ω (2.28)

The above expression is the inertial acceleration of the body of mass m. The
term 2(~ω × ~r˙xyz ) is called coriolis acceleration. The term ω ~ × (~ ω × ~r) is
known as centripetal acceleration. Note that the motion of the body of mass
m is observed in the non-inertial reference frame S − xyz which has linear
acceleration, ρ~¨XY Z , angular velocity ω
~ , and angular acceleration ω ~˙ . If we take
¨
~¨XY Z =~
ρ ~˙ =0 in the above equation, then R
ω =ω ~ XY Z = ~r¨xyz (i.e., accelerations
2.2. POINT MASS 45

in both the frames are same). In other words, if the two frames are non-
accelerating and non-rotating, the accelerations observed in both the frames
will be the same and we call both the frames as the inertial reference frames.
We take another situation when the non-inertial reference frame S − xyz
coincides with the inertial reference frame 0 − XY Z (i.e., the frame center S
lies at the frame center 0). Then we have

ρ
~XY Z = ρ ~˙ XY Z = ρ~¨XY Z = 0 (2.29)
~rxyz = R~ XY Z (2.30)

Using these relations in Eqs. (2.22), (2.26) (2.28), we can express inertial
position, velocity and accelerations as
~ XY Z = R
R ~ xyz (2.31)
~˙ XY Z = R
R ~˙ xyz + ω
~ ×R~ xyz (2.32)
~¨ XY Z = R
R ~¨ xyz + 2(~
ω×R ~˙ xyz ) + ω
~ × (~
ω×R ~˙ × R
~ xyz ) + ω ~ xyz (2.33)

In general, we place inertial and non-inertial reference frames at the center of


mass of the body whose motion is under study. The non-inertial frame is fixed
to the body and rotating with it while the inertial reference frame is fixed
in space. In such cases these equations are very useful to find inertial states
of the motion and then we apply Newton’s laws of motion. For studying
the motion of spacecraft and aircrafts around the Earth, the reference frame
fixed to the Earth at its center of mass is treated as an inertial reference
frame. This frame is known as geocentric reference frame. However, the
Earth is not a fixed body and is rotating about its own axis with a period
23.94 hours. The error due to the Earth motion is found to be negligible for
our analysis of the motion of spacecraft in the Earth orbit as the magnitudes
of ω
~ and R ~ are quite small. But this error is significant for spacecraft in
planetary missions and even for the orbital motion of the Moon. Therefore,
for planetary missions, an inertial reference frame fixed at the center of the
Sun, called the heliocentric reference frame, or at the distant stars, is highly
accurate.
Example 2.1
A dumbbell satellite system comprising of two spacecraft of point masses
m1 and m2 connected by a massless rigid cable of length L is orbiting in
a circular orbit of radius R around the Earth (Fig. 2.4). The position of
the system center of mass S is defined by R, the distance from the center
of Earth to the system center of mass, and θ, an angle measured from the
reference line X. Upon application of a controller the masses m1 and m2
remain aligned along the Local vertical. Find the inertial position, velocity,
and acceleration vectors and their corresponding magnitudes for spacecraft
m1 and m2 .
46 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

Orbit

m2
L
S

Y R m1

Local Vertical

θ
E
X

Z
Figure 2.4: A dumbbell satellite system.

Solution.
The vectors R~ 1 and R~ 2 describe the positions of spacecraft m1 and m2
from the center of the inertial reference frame E − XY Z (Fig. 2.5). Let r1
and r2 be the positions of m1 and m2 from S.

Orbit

m2
R2 S
m1 L
Y R

R1
θ
E
X

Figure 2.5: A dumbbell satellite system showing the position vectors of space-
craft.
2.2. POINT MASS 47

The position vectors of spacecraft m1 and m2 can be expressed as

~1 = R
R ~ + ~r1 (2.34)
~2 = R
R ~ + ~r2 (2.35)

As the center of mass lies at S, we have

m1~r1 + m2~r2 = 0 (2.36)

~ from the mass m2 to the mass m1 as shown in Fig. 2.5.


Let us define L
~ as
Then we can write L

~ = ~r2 − ~r1
L (2.37)

~ as
Using Eqs. (2.36-2.37), we can write ~r1 and ~r2 in terms of L
m2 ~
~r1 = − L (2.38)
m1 + m2
m1 ~
~r2 = L (2.39)
m1 + m2

Substituting the above expressions in Eqs. (2.34) and (2.35), we have

~1 = R
~− m2 ~
R L (2.40)
m1 + m2
~2 = R
~+ m1 ~
R L (2.41)
m1 + m2

Next we fix a rotating coordinate frame, E − îĵ k̂ at the Earth center with
~ k̂ normal to the orbital plane, and ĵ completes the right-handed
î along R,
triad. This coordinate frame is rotating with ω ~ = θ̇k̂. Writing R ~ and L~ in
Eqs. (2.40) and (2.41) in terms of unit vectors of the frame E − îĵ k̂ yields
 
~1 = R − m2
R L î (2.42)
m1 + m2
 
~2 = R + m1
R L î (2.43)
m1 + m2

or the corresponding magnitudes as

~ 1| = R − m2
R1 = |R L (2.44)
m1 + m2
~ 2| = R + m1
R2 = |R L (2.45)
m1 + m2
48 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

The velocity vectors of spacecraft m1 and m2 are derived as follows:

~  
~1 = dR1 = R −
V
m2
L î˙ (2.46)
dt m1 + m2
~
dR2

m1

~
V2 = = R+ L î˙ (2.47)
dt m1 + m2

Upon substitution of the relation

î˙ = ω
~ × î = θ̇k̂ × î = θ̇ĵ (2.48)

in the preceding equations, we have


 
~1 = R − m2
V L θ̇ĵ (2.49)
m1 + m2
 
~2 = R + m1
V L θ̇ĵ (2.50)
m1 + m2

and the corresponding magnitudes as


 
~1 | = R − m2
V1 = |V L θ̇ (2.51)
m1 + m2
 
~2 | = R + m1
V2 = |V L θ̇ (2.52)
m1 + m2

The acceleration vectors of spacecraft m1 and m2 are next derived con-


sidering θ̇=constant and θ̈ = 0:

~1
dV

m2

~a1 = = R− L θ̇ĵ˙ (2.53)
dt m1 + m2
~2
dV

m1

~a2 = = R+ L θ̇ĵ˙ (2.54)
dt m1 + m2

Applying the relation

ĵ˙ = ω
~ × ĵ = θ̇k̂ × ĵ = −θ̇î (2.55)

in the preceding equations leads to

dV~1 
m2

~a1 = =− R− L θ̇2 î (2.56)
dt m1 + m2
dV~2 
m1

~a2 = =− R+ L θ̇2 î (2.57)
dt m1 + m2
2.2. POINT MASS 49

The corresponding magnitudes of acceleration vectors are


 
m2
a1 = |~a1 | = R − L θ̇2 (2.58)
m1 + m2
 
m1
a2 = |~a2 | = R + L θ̇2 (2.59)
m1 + m2

Remarks. We have special cases as follows:

(a) If the two spacecraft merge, i.e., L=0, the preceding relations (2.44)-
(2.59) reduces to the case of a single spacecraft, derived earlier Eqs.
(2.9)-(2.16).

(b) If the mass m1 happens to be a spacecraft and the mass m2 is a payload,


then for the case m1  m2 , i.e., (m2 /m1 ) ⇒ 0, the preceding equations
for position, velocity, and acceleration vectors reduce to
 
~1 = m2 /m1
R R− L î = Rî (2.60)
1 + m2 /m1
 
~2 1
R = R+ L î = (R + L)î (2.61)
1 + m2 /m1
 
~1 m2 /m1
V = R− L θ̇ĵ = Rθ̇ĵ (2.62)
1 + m2 /m1
 
~2 1
V = R+ L θ̇ĵ = (R + L)θ̇ĵ (2.63)
1 + m2 /m1
 
m2 /m1
~a1 =− R− L θ̇2 î = −Rθ̇2 î (2.64)
1 + m2 /m1
 
1
~a2 =− R+ L θ̇2 î = −(R + L)θ̇2 î (2.65)
1 + m2 /m1

Note that the system center of mass lies on the spacecraft m1 .

Rotation matrix. In the case when the two coordinate frames differ
by an angle, we can transform coordinates in one coordinate frame to an-
other frame using a transformation matrix called a Rotation matrix, usually
denoted by R. For example, let there be two coordinate frames O − XY Z
and O − xyz as shown in Fig. 2.6. In terms of unit vectors, these coordinate
frames can be described by O − IˆJˆK̂ and O − îĵ k̂, respectively. The frame
O − xyz is obtained by θ rotation of the frame O − XY Z about the x-axis.
Note that the x-axis and X-axis are the same, i.e., î = I.ˆ The unit vectors
in the frames O − xyz and O − XY Z are related by the rotation matrix as
50 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

follows:
   


 î 




 Iˆ 


   
ĵ = Rx (θ) Iˆ (2.66)

 
 
 

 k̂ 
  
 K̂ 

where Rx (θ) is given by


 
1 0 0
 
Rx (θ) = 0 cosθ sinθ (2.67)
 
 
0 −sinθ cosθ

Applying the preceding relation (2.66), the position of the spacecraft m given
with respect to the unit vectors î, ĵ, and k̂ in the frame O − xyz can be
expressed in terms of the unit vectors I,ˆ J,
ˆ and K̂ in the frame O − XY Z.
The expression Rx (θ) in the above equations denotes the rotation matrix
for θ rotation about the x-axis. More conveniently, Rx (θ) is also denoted
as R1 (θ) where 1 refers to x. Note that the angle θ is the rotation in the
anti-clockwise direction. We can derive the rotation matrix R about the y

z m
R

y
θ
O
Y

X, x

Figure 2.6: Two coordinate frames.

and z-axes as follows:


 
cosθ 0 −sinθ
 
Ry (θ) = R2 (θ) =  0 1 0 (2.68)
 

 
sinθ 0 cosθ
2.2. POINT MASS 51
 
cosθ sinθ 0
 
Rz (θ) = R3 (θ) = −sinθ cosθ 0 (2.69)
 
 
0 0 1

where 2 and 3 corresponds to y and z axes, respectively.


The rotation matrix R has the following properties:

(a) Orthogonal: As all the axes are orthogonal, we have

RRT = 1 or RT = R−1 (2.70)

where superscripts T and −1 denote transpose and inverse of the ma-


trix.
(b) Unit norm:

k R k= 1 (2.71)

Applying the property (2.70), we can express the unit vectors in the O −
XY Z frame in terms of the unit vectors of the O − xyz frame as

   


 Iˆ 




 î 


   

−1
= Rx (θ) ĵ (2.72)

 
 
 

 K̂ 
   k̂ 
 

where Rx−1 (θ) is given by


 
1 0 0
 
Rx−1 (θ) = RxT (θ) = 0 cosθ −sinθ (2.73)
 
 
0 sinθ cosθ

Example 2.2
For a given inertial position of a spacecraft m in a circular orbit,
~ = Rî
R

in term of a unit vector î along the x-axis in the O − xyz rotating frame
(see Fig. 2.7), determine the inertial position, velocity, and acceleration of
the spacecraft in terms of unit vectors associated with the O − XY Z inertial
frame.
52 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

R m
y
x
θ
O
X

Z, z

Figure 2.7: A spacecraft in orbit.

Solution.
The inertial position of the spacecraft m is given in term of a unit vector
î along the x-axis in the O − xyz rotating frame as
~ = Rî
R (2.74)

The coordinate frame O − xyz is obtained from the inertial coordinate


frame O − XY Z by anticlockwise rotation through an angle θ about the
Z − axis. The transformation between the two frames can be written as
   


 î 




 Iˆ 


   
ĵ = Rz (θ) Jˆ (2.75)

 
 
 

 k̂ 
   K̂ 
 

where
 
cosθ sinθ 0
 
Rz (θ) = −sinθ cosθ 0 (2.76)
 
 
0 0 1

From the preceding Eq.(2.75), the unit vector î can be expressed in terms
of unit vectors along the inertial coordinate axes as
î = cosθIˆ + sinθJˆ (2.77)

Substituting î from the preceding Eq. (2.77) into Eq. (2.74), we have the
inertial position vector R~ in terms of unit vectors along the inertial axes as
~ = RcosθIˆ + RsinθJˆ
R (2.78)
2.2. POINT MASS 53

The corresponding magnitude is given as


1/2
~ = R2 cos2 θ + R2 sin2 θ
R = |R| =R (2.79)

The inertial velocity vector is derived using the expression for position
vector, Eq. (2.78) as

~
~ = dR = −Rθ̇sinθIˆ + Rθ̇cosθJˆ
V (2.80)
dt

ˆ˙ J=0
Note that I= ˆ˙ as Iˆ and Jˆ are unit vectors along inertial axes.
The magnitude of the velocity vector is
 1/2
~ | = R2 θ̇2 sin2 θ + R2 θ̇2 cos2 θ
V = |V = Rθ̇ (2.81)

Using the velocity vector Eq. (2.80), the inertial acceleration vector is
obtained as
~
dV
~a = = −Rθ̇2 cosθIˆ − Rθ̇2 sinθJˆ (2.82)
dt
The corresponding magnitude is given by
 1/2
a = |~a| = R2 θ̇4 cos2 θ + R2 θ̇4 sin2 θ = Rθ̇2 (2.83)

Remark. The magnitudes of inertial position, velocity and acceleration


vectors derived here are the same as obtained previously when these vectors
are expressed in terms of a rotating coordinate frame. Thus, the magnitudes
of inertial position, velocity and acceleration vectors remain the same whether
they are expressed in terms of an inertial coordinate frame (i.e., O − XY Z)
or a rotating coordinate frame (i.e., O − xyz). These observations can be
attributed to the fact that the expressions for inertial position, velocity and
acceleration in both the frames O − XY Z and O − xyz are with respect to
the same inertial point O.
Example 2.3
For studying the motion of the Moon, if we fix an inertial reference frame
at the center of the Earth, then what would be the inertial acceleration of the
Moon? However, if the inertial reference frame is fixed at the center of the
Sun, what would be the changes in the inertial acceleration? Finally, where
should we fix the inertial reference frame in order to obtain a more accurate
inertial acceleration? The parameters for the Earth and Moon are given as
follows:
54 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

Celestial Mass Radius Graviational Mean distance Period


Body Constant (µ) from Sun/Earth around Sun/Earth
(kg) (km) (km3 /s2 ) (km) (days)
Earth 5.974 × 1024 6378.14 3.986 × 105 1.49599× 108 365.256
(1AU from Sun) (about Sun)
Moon 7.3483 × 1022 1738.2 4.903 × 103 3.844× 105 29.5

Note. Assume that the Moon is moving in a circular orbit around the
Earth. Furthermore, the Earth is also orbiting in a circular orbit around the
Sun. Neglect the motion of the Earth about its own axis as the Earth is
considered as a point mass.
Solution.
The inertial position of the Moon with respect to the Earth (considering
the inertial frame E − XY Z fixed with the Earth at its center of mass) is

~r = rî (2.84)

where the frame E − xyz is a rotating frame with î along the radial vector,
k̂ orbit normal, and ĵ completes the right-handed triad.
Assume that the Moon is moving in a circular orbit around the Earth,
the inertial acceleration of the Moon is obtained as

~a = ~r¨ = −θ̇2 rî (2.85)

where θ̇ is the angular velocity with respect to the inertial reference frame
E − xyz. The angle θ denotes the angle with respect to the reference line
X-axis.
The inertial position of the Moon with respect to the Sun (considering
the inertial frame S − XY Z fixed with the Sun at its center of mass) is

~m = R
R ~ + ~r (2.86)

where R~ is the radial vector between the Sun and the Earth while ~r is the
radial vector between the Earth and the Moon.
We represent the vectors R~ and ~r with respect to coordinate frames S −
xm ym zm and E − xyz, respectively. The frame S − xm ym zm is fixed at the
~ k̂m towards orbit
center of mass of the Sun with îm along the radial vector R,
normal, and ĵm completes the right-handed triad. The frame E − xyz is fixed
at the center of mass of the Earth with î along the radial vector ~r, k̂ towards
2.2. POINT MASS 55

orbit normal, and ĵ completes the right-handed triad. So, we can write the
inertial position of the Moon as
~ m = Rîm + rî
R (2.87)

The inertial acceleration of the Moon is obtained as

~¨ m = −θ̇2 Rîm − θ̇2 rî


~am = R (2.88)
m

The corresponding magnitude can be expressed as


2
am = [(θ̇m R)2 + (θ̇2 r)2 + 2(θ̇m
2 2
θ̇ Rr)î · îm ]1/2
2
= [(θ̇m R)2 + (θ̇2 r)2 + 2(θ̇m
2 2
θ̇ Rr)cos(θm − θ)]1/2 (2.89)

In the case the Sun, the Earth, and the Moon are aligned, (i.e., θm =θ),
then the magnitude of the inertial acceleration of the moon simplifies to
 2  2
2 2 2π 2π
am = θ̇m R + θ̇ r = R+ r (2.90)
Tm T

Note that Eq. (2.85) denotes the inertial acceleration of the moon fixing
the inertial frame at the Earth center of mass while Eq. (2.90) represents the
inertial acceleration of the moon considering the inertial frame at the Sun
center of mass. Taking the numerical data:
T (orbital period)=29.5 days=29.5×24×3600=2548800 sec; Tm (orbital
period)=365.25 days=31557600 sec; R=1.49599×108 km; r=3.844× 105 km
the inertial acceleration of the moon in the former case (given by Eq. (2.85))
is
 2

a= × 3.844 × 105 = 2.338 × 10−6 km/s2 (2.91)
2548800

while for the later case, the inertial acceleration of the moon with respect to
the Sun (given by Eq. (2.85)) is
 2  2
2π 8 2π
am = × 1.49599 × 10 + × 3.844 × 105
31557600 2548800
= 5.936 × 10−6 + 2.338 × 10−6 = 8.274 × 10−6 km/s2 (2.92)

From the preceding results, it is more accurate to fix an inertial frame at the
center of mass of the Sun for studying the motion of the Moon. Note that
the component of acceleration due to rotation of the Earth about the Sun
2
(i.e., θ̇m R=5.936× 10−6 km/s2 ) is larger than the component of acceleration
due to rotation of the Moon about the Earth (i.e., θ̇2 r=2.338× 10−6 km/s2 ).
It would be erroneous to fix an inertial frame at the Earth center of mass.
56 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

Example 2.4
For a given problem in Example 2.1, if the dumbbell system undergoes an
in-plane libration or oscillation of β about the local vertical (Fig. 2.8), then
determine the inertial position, velocity, and acceleration vectors and their
corresponding magnitudes for spacecraft m1 and m2 .

Orbit m2

β
L
S
R
Y
m1

θ
E
X

Figure 2.8: A dumbbell satellite system undergoing in-plane libration.

Solution.
As derived in Example 2.1, the position vectors of the spacecraft m1 and
m2 are
~1 = R
~− m2 ~
R L (2.93)
m1 + m2
~2 = R
~+ m1 ~
R L (2.94)
m1 + m2

In order to express R~ and L ~ in the preceding equations in terms of unit


vectors of coordinate frames, we consider four coordinate frames:

(a) The frame E − XY Z is an inertial coordinate frame with origin at


the the Earth’s center of mass E and the X and Y -axes are in the
orbit plane while the Z-axis is normal to the orbit plane (i.e., ⊥ to the
ˆ J,
paper). The unit vectors I, ˆ and K̂ are along the X, Y , and Z-axes,
respectively.
(b) The frame E − x0o yo0 zo0 is a reference coordinate frame with the x0o -axis
~ the y 0 -axis in the orbital plane is perpendicular to the
aligned with R, o
2.2. POINT MASS 57

Y
x
y R xo y L xo
o o
θ
β
θ β
y
E S
X
θ β
Z, zo zo , z

Figure 2.9: Orientation of coordinate frames.

x0o - axis, and the zo0 -axis completes the right-handed triad (i.e., normal
to the orbit plane). The unit vectors î0o , ĵo0 , and k̂o0 are along the x0o ,
yo0 , and zo0 -axes, respectively. The zo0 -axis coincides with the inertial
Z-axis.
(c) The frame S − xo yo zo is an orbital reference coordinate frame with the
xo -axis aligned with R, ~ the yo -axis in the orbital plane is perpendicular
to the xo - axis, and the zo -axis completes the right-handed triad. The
unit vectors îo , ĵo , and k̂o are along the xo , yo , and zo -axes, respectively.
(d) S − xyz, a coordinate frame with the x-axis aligned with L, ~ the y-axis
~ and the z-axis completes the
in the orbit plane is perpendicular to L,
right-handed triad. The unit vectors î, ĵ, and k̂ are along the x, y, and
z-axes, respectively.

Note that the frames E − x0o yo0 zo0 and S − xo yo zo differs by the origin of the
coordinate frame. The frame S−xo yo zo is obtained from the frame E−x0o yo0 zo0
by translation of a distance of R. The unit vectors in both the frames are
same.
The frames E − x0o yo0 zo0 , S − xo yo zo , and S − xyz are rotating coordinate
frames. Let us derive the angular velocity vectors for all these frames. The
rotation angle vector for the frame E − x0o yo0 zo0 is

d~θo0 = dθk̂o0 (2.95)

Differentiating the preceding expression with respect to time, we have

d~θo0
~ o0 =
ω = θ̇ k̂o0 (2.96)
dt
˙
Note here k̂o0 = 0 as the frame E − x0o yo0 zo0 rotates about the zo0 - axis.
58 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

As mentioned earlier, the unit vectors are the same in the frames E −
x0o yo0 zo0 and S − xo yo zo . So, we have the angular velocity vector of the frame
S − xo yo zo , denoted by ω ~ o0 of
~ o , is the same as the angular velocity vector ω
0 0 0
the frame E − xo yo zo , i.e.,
ω
~ o = θ̇k̂o (2.97)
Here k̂o =k̂o0 .
In the case of the frame S − xyz, the rotation angle vector is
~θ = θk̂o + β k̂ (2.98)
Differentiating the preceding expression with respect to time, we have
~ = d~
ω θ/dt = θ̇k̂o + β̇ k̂ = (θ̇ + β̇)k̂ (2.99)
˙
Note here k̂ = 0 as the frame S − xyz rotates about the z- axis. The unit
vectors k̂o and k̂ are normal to the orbit plane.
Now we can express R ~ and L~ in terms of unit vectors of the coordinate
frames S − xo yo zo and S − xyz as
~ = Rîo
R (2.100)
~ = Lî
L (2.101)

Substituting the preceding relations in Eqs. (2.93) and (2.94) and taking
m2
γ= (2.102)
m1 + m2
we have
~ 1 = Rîo − γLî
R (2.103)
~ 2 = Rîo + (1 − γ)Lî
R (2.104)
where 1 − γ = m1 /(m1 + m2 ).
To express î in terms of unit vectors in the frame S − xo yo zo , we consider
the transformation as
   
 î   î 
 o

 
 
 

  
ĵ = Rz (β) ĵo (2.105)

 
 
 

 k̂ 
   k̂
 

o

where Rz (β) is
 
cosβ sinβ 0
 
Rz (β) = −sinβ cosβ 0 (2.106)
 
 
0 0 1
2.2. POINT MASS 59

Using Eq. (2.105), we can write î as

î = cosβ îo + sinβ ĵo (2.107)

Substituting the preceding expression in Eqs. (2.103) and (2.104), we have


the position vectors as
~ 1 = (R − γLcosβ) îo − γLsinβ ĵo
R (2.108)
~ 2 = [R + (1 − γ)Lcosβ] îo + (1 − γ)Lsinβ ĵo
R (2.109)

The corresponding magnitudes are

~ 1 | = R2 + γ 2 L2 − 2γRLcosβ 1/2
 
R1 = |R (2.110)
~ 2 | = R2 + (1 − γ)2 L2 + 2(1 − γ)RLcosβ 1/2
 
R2 = |R (2.111)

The velocity vectors of spacecraft m1 and m2 are derived as

~1
dR
= γLβ̇sinβ îo + (R − γLcosβ) î˙ o
 
V~1 =
dt
− γLβ̇cosβ ĵ − γLsinβ ĵ˙
o o (2.112)
~2
dR
= −(1 − γ)Lβ̇sinβ îo + [R + (1 − γ)Lcosβ] î˙ o
h i
V~2 =
dt
+ (1 − γ)Lβ̇cosβ ĵ + (1 − γ)Lsinβ ĵ˙
o o (2.113)

Upon substitution of the relations

î˙ o = ω
~ × îo = θ̇k̂o × îo = θ̇ĵo (2.114)
˙ĵ = ω ~ × ĵo = θ̇ k̂o × ĵo = −θ̇îo (2.115)
o

in the preceding equations, we have


h i h i
V~1 = γ(θ̇ + β̇)Lsinβ îo + Rθ̇ − γ(θ̇ + β̇)Lcosβ ĵo (2.116)
h i h i
V~2 = −(1 − γ)(θ̇ + β̇)Lsinβ îo + Rθ̇ + (1 − γ)(θ̇ + β̇)Lcosβ ĵo (2.117)

and the corresponding magnitudes as


h i1/2
~1 | = R2 θ̇2 + γ 2 (θ̇ + β̇)2 L2 − 2γRθ̇(θ̇ + β̇)Lcosβ
V1 = |V (2.118)
h i1/2
~2 | = R2 θ̇2 + (1 − γ)2 (θ̇ + β̇)2 L2 + 2(1 − γ)Rθ̇(θ̇ + β̇)Lcosβ
V2 = |V
(2.119)
60 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

The acceleration vectors of spacecraft m1 and m2 are next derived as:


~1
dV
= γ β̈Lsinβ + γ(θ̇ + β̇)β̇Lcosβ îo + γ(θ̇ + β̇)Lsinβ î˙ o
h i h i
~a1 =
dt
+ −γ β̈Lcosβ + γ(θ̇ + β̇)β̇Lsinβ ĵo + Rθ̇ − γ(θ̇ + β̇)Lcosβ ĵ˙ o
h i h i

(2.120)
~2
dV h i
~a2 = = −(1 − γ)β̈Lsinβ − (1 − γ)(θ̇ + β̇)β̇Lcosβ îo
dt
+ −(1 − γ)(θ̇ + β̇)Lsinβ î˙
h i
o
h i
+ (1 − γ)β̈Lcosβ − (1 − γ)(θ̇ + β̇)β̇Lsinβ ĵo

+ Rθ̇ + (1 − γ)(θ̇ + β̇)Lcosβ ĵ˙ o


h i
(2.121)

Using Eqs. (2.114) and (2.115), we can derive the acceleration vectors and
the corresponding magnitudes.

2.2.2 Momentum
Momentum can be described as “mass in motion” i.e., if a spacecraft of mass
m is moving, we say it has momentum. There can be linear momentum and
angular momentum as described next.
Linear Momentum
The linear momentum of a spacecraft of mass m is defined as

p
~ = m~v (2.122)

where ~v is an inertial velocity vector, i.e., velocity vector relative to an inertial


point. For example, let a spacecraft in an orbit as shown in Fig. 2.10. Writing
~v in terms of polar coordinates, we have the linear momentum as

p~ = m(ṙ î + rθ̇ ĵ) (2.123)

where î is along ~r and, ĵ is ⊥ to ~r in the same plane. We can derive the same
result alternatively writing the linear momentum given by

p~ = m~r˙ = m[~r˙xyz + ω
~ × ~r] (2.124)

where ~r˙xyz is the spacecraft velocity with respect to the O − xyz frame with
the x-axis along ~r, the z-axis is normal to the orbit plane, and the y-axis
completes the right-handed triad. The vector ω ~ denotes the inertial angular
velocity of the O − xyz frame. In this example, ω ~ = θ̇k̂, ~r = rî, and ~r˙xyz = ṙ î,
and so substituting these in Eq. (2.124) leads to the same expression as given
by Eq. (2.123).
2.2. POINT MASS 61

v
m

r
θ
O
X

Figure 2.10: A spacecraft in orbit.

Example 2.5
Determine the linear momentum of the system described in Example 2.4.
Solution.
The linear momentum of the system denoted by p
~ is the sum of the linear
momentum of spacecraft m1 and m2 , i.e.,

p
~ = p~1 + p
~2 (2.125)

where p~1 and p~2 denote the linear momentum of spacecraft m1 and m2 ,
respectively. These are expressed as

p ~1
~ 1 = m1 V (2.126)
p ~2
~ 2 = m1 V (2.127)

The velocity vectors V1 and V2 are

~˙ + ~r˙1
V~1 = R (2.128)
~˙ + ~r˙2
V~2 = R (2.129)

~1 and V
Substituting the preceding expressions for V ~2 into Eqs. (2.126) and
(2.127), we can write the system linear momentum referring to Eq. (2.125)
as
~˙ + m1~r˙1 + m2~r˙2
p~ = (m1 + m2 )R (2.130)

Knowing that the system center of mass lies at S, we have

m1~r1 + m2~r2 = 0 (2.131)


62 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

Differentiating the above equation with respect to the time yields


m1~r˙1 + m2~r˙2 = 0 (2.132)

Note the spacecraft masses m1 and m2 are assumed to be constant. Applying


the above relation (2.132) into Eq. (2.130) leads to


p~ = M R (2.133)

where M = m1 + m2 , total system mass. The preceding equation states that


the linear momentum of the system is the product of the system mass and the
inertial velocity vector.
Angular Momentum
The angular momentum of a spacecraft of mass m is defined as
~ = ~r × m~v
H (2.134)

where ~r is an inertial position vector, i.e., position vector relative to an inertial


point. For the same example shown in Fig. 2.10, the angular momentum of
the spacecraft is
~ = mrî × (ṙ î + rθ̇ ĵ) = mr2 θ̇ k̂
H (2.135)

The above result can be derived more elegantly using the vector method as
explained next.
The spacecraft velocity vector (with respect to inertial frame O − XY Z)
can be expressed as

~v = ~r˙ = ~r˙xyz + ω
~ × ~r (2.136)

where ~r˙xyz represents velocity with respect to the rotating frame O −xyz and
ω
~ denotes the angular velocity with respect to the inertial frame O − XY Z.
Now we can express the angular momentum of the spacecraft as
~ = m~r × (~r˙xyz + ω
H ~ × ~r) = m[~r × ~r˙xyz + ~r × (~
ω × ~r)] (2.137)

Applying the triple vector product relation ~a × (~b ×~c) = (~a ·~c)~b − (~a · ~b)~c, and
using the vector relation ~a · ~a = a2 , we obtain
~ = m[~r × ~r˙xyz + (~r · ~r)~
H ~ )~r] = m[~r × ~r˙xyz + r2 ω
ω − (~r · ω ~ − (~r · ω
~ )~r]
(2.138)

The vector ~r and ~r˙xyz have the same direction vector (i.e., ~r = rî, ~r˙xyz =
ṙ î). For a planar motion of the spacecraft (i.e., in the r-θ plane), ω~ ⊥ ~r.
Considering these facts, we have

~r · ω
~ = 0, ~r × ~r˙xyz = 0 (2.139)
2.2. POINT MASS 63

Using these relations in Eq. (2.138), we obtain


~ = mr2 ω
H ~ = mr2 θ̇k̂ (2.140)

Thus, the preceding equation states that the angular momentum of the system
is the products of the system mass, the square of the distance from the system
center of mass, and the inertial angular velocity vector.
Note. The angular momentum per unit mass of the spacecraft is called
specific angular momentum denoted by h.
Example 2.6
Determine the angular momentum of the system described in Example
2.4.
Solution.
~ is the sum of the
The angular momentum of the system denoted by H
angular momentum of spacecraft m1 and m2 , i.e.,
~ =H
H ~1 + H
~2 (2.141)

where H~ 1 and H
~ 2 denote the angular momentum of spacecraft m1 and m2 ,
respectively. These are expressed as
~1 = R
H ~ 1 × m1 V
~1 (2.142)
~2 = R
H ~ 2 × m2 V
~2 (2.143)

Knowing

~j = R
R ~ + ~rj , j = 1, 2; ~˙ j = R
~j = R
V ~˙ + ~r˙j , j = 1, 2

we can write the system angular momentum referring to Eq. (2.141) as

~ =m1 (R
H ~˙ + ~r˙1 ) + m2 (R
~ + ~r1 ) × (R ~ + ~r2 ) × (R~˙ + ~r˙2 )

=(m1 + m2 )(R ~˙ + R
~ × R) ~ × (m1~r˙1 + m2~r˙2 )

+ (m1~r1 + m2~r2 ) × R ~˙ + m1 (~r1 × ~r˙1 ) + m2 (~r2 × ~r˙2 ) (2.144)

As the system center of mass lies at S, we have

m1~r1 + m2~r2 = 0, m1~r˙1 + m2~r˙2 = 0 (2.145)

Note the spacecraft masses m1 and m2 are constant. Applying the above
relation (2.145) to Eq. (2.144) yields

~ = (m1 + m2 )(R
H ~˙ + m1 (~r1 × ~r˙1 ) + m2 (~r2 × ~r˙2 )
~ × R) (2.146)
64 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

Here the velocity vectors are given by


~˙ = R
R ~˙ xyz + ω ~
~ o × R, ~r˙j = (~r˙j )xyz + ω
~ × ~rj , j = 1, 2
where ω~ o and ω
~ are the angular velocity vectors (with respect to the inertial
frame E−XY Z) of the coordinate frames S−xo yo zo and S−xyz, respectively.
Knowing that the system is rotating in a circular orbit (i.e., R ~˙ xo yo zo =0)
and the length connecting the two spacecraft (i.e., (~r˙1 )xyz =(~r˙2 )xyz =0) is con-
stant, we have
~ = (m1 + m2 )[R
H ~ × (~ ~ + m1 [~r1 × (~
ωo × R)] ω × ~r1 )] + m2 [~r2 × (~
ω × ~r2 )]
(2.147)

Applying the triple vector product relation ~a × (~b × ~c) = (~a · ~c)~b − (~a · ~b)~c
yields
~ =(m1 + m2 )[(R
H ~ · R)~
~ ωo − (R ~ ·ω ~ + m1 [(~r1 · ~r˙1 )~
~ o )R] ω − (~r1 · ω
~ )~r1 ]
+ m2 [(~r2 · ~r˙2 )~
ω − (~r2 · ω
~ )~r2 ] (2.148)

As ω ~ ω
~ o ⊥ R, ~ ⊥ ~r1 , and ω
~ ⊥ ~r2 , we have
~ ·ω
R ~ o = 0, ~r1 · ω
~ = 0, ~r2 · ω
~ =0 (2.149)
Substituting the above relations into Eq. (2.148), we have
~ = (m1 + m2 )R2 ω
H ~ o + [m1 (~r1 · ~r1 ) + m2 (~r2 · ~r2 )]~
ω (2.150)
Knowing
m2 ~ m1 ~
~r1 = − L, r~2 = L
m1 + m2 m1 + m2
and using the dot product relation ~a · ~a = a2 , we obtain
~ = M R2 ω
H ~ o + Me L2 ω
~ (2.151)
where M = m1 +m2 denotes total system mass and Me = (m1 m2 )/(m1 +m2 )
is an equivalent system mass. Note that the first term in the preceding
equation denotes the angular momentum due to the system orbital motion
while the second term is the angular momentum of the system libration about
the local vertical.

2.2.3 Energy
The energy of a system, denoted by E is comprised of its potential energy,
U and its kinetic energy, T , i.e.,
E =U +T (2.152)
2.2. POINT MASS 65

Potential Energy
The potential energy of a spacecraft m orbiting about a planet (Fig. 2.11)
due to its gravity is given by
µm
U =− (2.153)
r
where r is the inertial position of the spacecraft and µ is the gravitational
constant of the planet around which the spacecraft is orbiting. As this po-
tential energy is due to the gravitational force of the planet, it is also called
gravitational potential energy.

r
θ
O
X
Planet

Figure 2.11: A spacecraft orbiting around a planet.

Kinetic Energy
The kinetic energy of a spacecraft m is given by
1 1  
T = m(~v .~v ) = m ṙ2 + θ̇2 r2 (2.154)
2 2
where ~v is the inertial velocity of the spacecraft.
Note. Potential or kinetic Energy is a scalar quantity function.
Example 2.7
Determine the potential and kinetic energies of the system described in
Example 2.4.
Solution.
Let us first derive the potential energy of the system followed by the
system kinetic energy.
Potential Energy
The potential energy U of the system is the sum of the potential energy
due to the spacecraft m1 , U1 and the potential energy due to the spacecraft
66 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

m2 , U2 , i.e.,

U = U1 + U2 (2.155)

The potential energy due to the spacecraft m1 can be expressed as


µm1
U1 = − (2.156)
~ 1|
|R
where R1 is the position vector of the spacecraft m1 given by
~1 = R
R ~ + ~r1 (2.157)

or
~ 1 | = [R2 + 2R.~
|R ~ r1 + r2 ]1/2 (2.158)
1

~ 2 = R2 and ~r2 = r2 . Substituting |R


as R ~ 1 | from Eq.(2.158) into Eq.(2.156),
we have
µm1
U1 = −
(R2 + 2R~ · ~r1 + ~r2 )1/2
1
" #−1/2
µm1 2R ~ · ~r1 r12
=− 1+ + 2 (2.159)
R R2 R

Applying Binomial series expansion,


n(n − 1) 2 n(n − 1)(n − 2) 3
(1 + x)n = 1 + nx + x + x + ···
1·2 1·2·3
∞  
X n k
= x if |x| < 1 (2.160)
k
k=0

where
 
n n n(n − 1) · · · (n − k + 1)
= = (2.161)
k k!(n − k)! 1 · 2···k
leads to
" #
µm1 ~ · ~r1
R 1 r12 ~ · ~r1 )2
3 (R
U1 = − 1− − + + ··· (2.162)
R R2 2 R2 2 R4

Now, considering the fact r1  R and expanding upto O(1/R3 ), we get


" #
µm1 R~ · ~r1 1 r12 ~ · ~r1 )2
3 (R
U1 = − 1− − + (2.163)
R R2 2 R2 2 R4
2.2. POINT MASS 67

Note that the above potential energy is approximate. However, this expres-
sion is accurate for most practical applications.
Following a procedure similar to that described above and knowing the
~2 = R
position vector R ~ + ~r2 , we can determine the potential energy of the
system due to spacecraft m2 as
" #
µm2 R~ · ~r2 1 r22 ~ · ~r2 )2
3 (R
U2 = − 1− − + (2.164)
R R2 2 R2 2 R4

Thus, we can write the system potential energy U as per Eq.(2.155) using
Eqs.(2.163-2.164) as
"
µ ~ · ~r1 ) + m2 (R
m1 (R ~ · ~r2 ) 1 m1 r2 + m2 r2
1 2
U =− (m1 + m2 ) − 2

R R 2 R2
#
~ · ~r1 )2 + m2 (R
3 m1 (R ~ · ~r2 )2
+ (2.165)
2 R4

As the center of mass lies at S, we get

m1~r1 + m2~r2 = 0 (2.166)

Let us consider the distance between m1 and m2 be L. Then

~ = ~r2 − ~r1
L (2.167)

~ as
Using Eqs. (2.166-2.167), we can write ~r1 and ~r2 in terms of L
m2 ~
~r1 = − L (2.168)
m1 + m2
m1 ~
~r2 = L (2.169)
m1 + m2

Now, we define R ~ and L


~ with respect to the orbital coordinate frame
S − xo yo zo and the dumbbell fixed coordinate frame S − xyz, respectively as

~ = Rîo ;
R ~ = Lî
L (2.170)

Substituting Eq. (2.166) and Eqs. (2.168-2.169) into Eq.(2.165), we have


µM µ
U =− + Me [1 − 3(~io · ~i)2 ]L2 (2.171)
R 2R3
68 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

where M and Me denote system mass and equivalent system mass, respec-
tively. They are

M = m1 + m2 (2.172)
m1 m2
Me = (2.173)
m1 + m2

The transformation from the frame S − xo yo zo to the frame S − xyz is


obtained by a rotation of β about xo -axis. Using transformation relation
(2.69), we have
   


 î 




 î o



   
ĵ = Rz (β) ĵo (2.174)

 
 
 

 k̂ 
   k̂ 
 
o

where
 
cosβ sinβ 0
 
Rz (β) = −sinβ cosβ 0
 
 
0 0 1

Using Eq.(2.174), we can write ĵ in terms of the orbital frame axes as

î = cosβ îo + sinβ ĵo (2.175)

Thus, we obtain (îo · î) as

îo · î = cosβ (2.176)

and substituting this into Eq. (2.171), we get the system potential energy
µM µ
U =− + Me (1 − 3cos2 β)L2 (2.177)
R 2R3
Note in the above potential energy expression, the first term is due to the
orbital motion of the system while the second term is because of the libration
of the two spacecraft about the system center of mass.
Kinetic Energy
The system kinetic energy T is

T = T1 + T2 (2.178)

Here, T1 and T2 are kinetic energies due to spacecraft m1 and spacecraft m2 ,


respectively.
2.2. POINT MASS 69

The kinetic energy of spacecraft m1 is given by


m1 ~˙ ~˙
T1 = (R1 · R1 ) (2.179)
2

Using Eq. (2.157), Eq.(2.179) can be written as


m1 ~˙ 2 ˙ 2 ~˙ · ~r˙1 ]
T1 = [R + ~r1 + 2R (2.180)
2

Similarly, we can derive kinetic energy T2 due to the spacecraft m2 as


m2 ~˙ 2 ˙ 2 ~˙ · ~r˙2 ]
T2 = [R + ~r2 + 2R (2.181)
2

Using expressions for T1 and T2 from Eq. (2.180) and Eq. (2.181), we
find the system kinetic energy T to be
(
1 ~˙ 2 + m1~r˙ 2 + m2~r˙ 2
T = (m1 + m2 )R 2 2
2
)
˙~ ˙ ˙~ ˙
+ 2[m1 (R · ~r1 ) + m2 (R · ~r2 )]
( )
1 ˙~ 2
= (m1 + m2 )R + m1~r˙12 + m2~r˙22 ) (2.182)
2

Now, we write ~r˙1 and ~r˙2 using Eqs.(2.168-2.169) as


m2 ~˙
~r˙1 = − L (2.183)
m1 + m2
m1 ~˙
~r˙2 = L (2.184)
m1 + m2

Knowing
!
~
dR
~˙ =
R ~
+ ω~o × R (2.185)
dt
xyz
!
~
dL
~˙ =
L +ω ~
~ ×L (2.186)
dt
xyz

and using Eq. (2.170), we have


! !
~
dR ~
dL
= Ṙîo ; = 0; ω
~ o = θ̇k̂o ; ω
~ = (θ̇ + β̇)k̂ (2.187)
dt dt
xyz xyz
70 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

Substituting Eqs. (2.183-2.184) into Eq. (2.182) and considering that


the system is moving in a circular orbit, Ṙ = 0, and the rod connecting two
masses is rigid, i.e., L̇ = 0, we have the system kinetic energy as
1 1
T = M (θ̇2 R2 ) + Me (θ̇ + β̇)2 L2 (2.188)
2 2
As stated previously, note that in the above expression the first term is due to
the orbital motion of the system and the second term is due to the libration
of the two spacecraft about the system center of mass.
Note that the total energy of a single point mass system (ı.e., sum of
system potential and kinetic energy) remains constant if there are no external
forces or torques. However, for an orbiting rigid-body system including point
masses connected by a cable or rod, the total energy does not remain constant,
instead, the Hamiltonian of the system defined as

H = T2 − T0 + U (2.189)

remains constant. Here Tk , k = 0, 2 denotes kinetic energy terms with power


k. The kinetic energy of a system can be written in terms of Tk as

T = T0 + T1 + T2 (2.190)

In the case that the kinetic energy T is proportional to the square of the
velocity of the coordinate (i.e., , T0 = T1 = 0), then the Hamiltonian of the
system equals to the system total energy E, as

H = T2 + U = E (2.191)

which remains constant as stated previously. Referring to Eq. 2.188 for the
system (Example 2.7), we have
1 1 1
T0 = M (θ̇2 R2 ) + Me θ̇2 L2 , T2 = Me α̇2 L2 (2.192)
2 2 2
H =T2 − T0 + U (2.193)

For example, consider pitch motion of a satellite in a circular orbit:


1 1
T = Ix ω 2 = Ix (θ̇ + α̇)2 (2.194)
2 2
1 1
T0 = Ix θ̇ , T2 = Ix α̇2
2
(2.195)
2 2

2.3 Rigid Body


In Section 2.2, we have discussed systems comprised of spacecraft assumed
to be point masses. These spacecraft can only have translational motion or
2.3. RIGID BODY 71

orbital motion (i.e., three dimensional motion along X, Y , and Z directions


in cartesian coordinates; R, θ, and Z in cylindrical coordinates; R, θ, and
φ in spherical coordinates) and there exists no rotational motion about its
fixed coordinate frame. However, the system can have rotational motion with
respect to another coordinate frame.
A rigid body is a system of point masses, fixed relative to each other (see
Fig. 2.12, the distance ρ remains fixed, i.e., ρ̇=ρ̈=0). It can have translational
or orbital motion, as if it is a point mass, as well as rotational motion about its
fixed coordinate frame, called attitude motion. As we did in the the previous
Section 2.2, here we first define the attitude motion and then we extend
the analysis presented for the point mass spacecraft to derive the inertial
position, velocity, and acceleration as well as the momentum and energy of
the rigid-body spacecraft.

m
dm
ρ
dm

Figure 2.12: A rigid-body spacecraft.

2.3.1 Defining Attitude


The orientation of a spacecraft with respect to a reference frame is called
the attitude of a spacecraft and the associated motion is termed as attitude
motion. In order to define the orientation, a coordinate frame is attached to
the spacecraft, generally at its center of mass, and it is called a body-fixed
coordinate frame. The orientation of this body-fixed frame, say S − xyz, with
respect to a reference frame, say O−XY Z defines the attitude of a spacecraft
(Fig. 2.13). The attitude can be represented using various methods such as
direction cosines, Euler angles, Quaternions, and others. These methods are
described below.
Direction cosines.
An orientation of an axis, say, ρ̂, with respect to the coordinate frame
O − XY Z (see Fig. 2.14) can be written as
ρ̂ = CX Iˆ + CY Jˆ + CZ K̂ (2.196)
ˆ J,
where I, ˆ and K̂ are unit vectors along X, Y , and Z axes, respectively.
Note ρ̂ is a unit vector.
72 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

Y m
S
x

z
R

O
X

Figure 2.13: Attitude of a spacecraft.

γ ρ

α
O
φ X

Figure 2.14: Orientation of an axis.

In Eq. (2.196), CX , CY , and CZ define the direction of the axis ρ̂ with


respect to axes X, Y , and Z, respectively, as
ρ̂ · Iˆ = CX , ρ̂ · Jˆ = CY , ρ̂ · K̂ = CZ (2.197)
In fact, CX , CY , and CZ are the cosines of the angles, say, α, γ, and φ, that
the axis ρ
~ makes with the axes X, Y , Z, respectively (Fig. 2.14). Thus, as
CX , CY , and CZ define directions as well as cosines of the angles, we call
them direction cosines. These direction cosines are not independent and they
follow the relation
2
CX + CY2 + CZ2 = 1 (2.198)
as the axes I, J, and K are orthogonal, i.e., ρ̂ · ρ̂=1.
2.3. RIGID BODY 73

For defining the attitude of a spacecraft, we have three axes (x, y, and
z) and therefore, we generalize the preceding Eq. (2.196) and express unit
vectors along three axes as

î = CxX Iˆ + CxY Jˆ + CxZ K̂


ĵ = CyX Iˆ + CyY Jˆ + CyZ K̂
k̂ = CzX Iˆ + CzY Jˆ + CzZ K̂ (2.199)

or it can written as
   


 î 

 

 ˆ
I 

   
ĵ =R Jˆ (2.200)

 
 
 

 k̂ 
  
 K̂ 

where R is called an attitude matrix or rotation matrix or direction cosine


matrix and is given by
 
CxX CxY CxZ
 
R = CyX CyY CyZ  (2.201)
 
 
CzX CzY CzZ

As discussed in Section 2.2, the rotation matrix R has the following prop-
erties:

(a) Orthogonal: All the axes should be orthogonal, expressed mathemati-


cally as

RRT = 1 or RT = R−1 (2.202)

(b) k R k=1

These properties lead to the following constraints:


X
2
Cij = 1, i = x, y, z (2.203)
j=X,Y,Z
X
Cij Ckj = 0, i, k = x, y, z; i 6= k (2.204)
j=X,Y,Z

The preceding equations result in 6 constraints (i.e., 3 constraints from Eqs.


(2.203) and 3 constraints from Eqs. (2.204)). In fact, the attitude requires
9 direction cosines (i.e., Cij , i=x, y, z; j=X, Y, Z). Therefore, we can say
that 3 direction cosines (i.e., 9-6=3) are sufficient to define the attitude of a
spacecraft.
74 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

Euler Angles.
The orientation of a spacecraft can be specified completely by a sequence
of three consecutive rotations about different spacecraft body axes. The first
and the last rotations about the same body axes are possible and thus, we
have 12 such possible combinations, i.e., 1-2-3, 1-3-2, 2-1-3, 2-3-1, 3-1-2, 3-2-
1, 2-1-2, 3-1-3, 1-2-1, 3-2-3, 1-3-1, and 2-3-2. Here, 1, 2, and 3 correspond
to x, y, and z axes, respectively. For example, using a 3-2-1 Euler angle
rotation sequence, and applying rotating matrix relations Eqs. (2.66)-(2.69)
we obtain the rotation matrix, R321 (α, φ, γ) as

R321 (α, φ, γ) =R1 (γ)R2 (φ)R3 (α)


 
cosαcosφ sinαcosφ −sinφ
 
= cosαsinφsinγ − sinαcosγ sinαsinφsinγ + cosαcosγ cosφsinγ 
 
 
cosαsinφcosγ + sinαsinγ sinαsinφcosγ − cosαsinγ cosφcosγ
(2.205)

Thus, the transformation from the frame S − io jo ko to the body fixed frame
S − ijk using 3-2-1 Euler angle rotation sequence is
   


 î 




 iˆo



   
ĵ = R321 (α, φ, γ) jˆo (2.206)
   
 kˆ 
   
 k̂ 
   
o

The transformation from the body fixed frame S −ijk to the frame S −io jo ko
is
   


 îo 




 î 


   
ĵo = R123 (−γ, −φ, −α) ĵ (2.207)

 
 
 

 k̂ 
   k̂ 
 
o

where the rotation matrix R123 (−γ, −φ, −α) by taking the inverse of the
−1
transformation R321 . However, from the properties of a rotation matrix ex-
−1 T
plained earlier, R321 is the transpose of R321 (i.e., R321 ). Thus, we find
R123 (−γ, −φ, −α) as
 
cosαcosφ cosαsinφsinγ − sinαcosγ cosαsinφcosγ + sinαsinγ
 
R123 = sinαcosφ sinαsinφsinγ + cosαcosγ sinαsinφcosγ − cosαsinγ 
 
 
−sinφ cosφsinγ cosφcosγ
(2.208)
2.3. RIGID BODY 75

In addition to rotation matrices, we need to determine the inertial angular


velocity ω
~ of the spacecraft. The angular velocity ω~ of the spacecraft can be
expressed as

ω
~ = ωx î + ωy ĵ + ωz k̂ = α̇k̂o + φ̇ĵ1 + γ̇ î (2.209)

where α̇ is about the ko axis in the frame S − xo yo zo , followed by φ̇ about


the j1 axis in the intermediate frame S − x1 y1 z1 and finally γ̇ about the î
axis in the frame S − xyz.

       
 î   î   î   î 
 1  o  2  1

 
 
 
 
 
 
 

   
ĵ1 = R3 (α) ĵo , ĵ2 = R2 (φ) ĵ1 ,

 
 
 
 
 
 
 

 k̂
 
  k̂   k̂ 
     k̂
 

1 o 2 1
   
 î   î 
 2 

 
 
 
  
ĵ = R1 (γ) ĵ2 (2.210)

 
 
 

 k̂ 
   k̂ 
 
2

Substituting ko and j1 from Eq. (2.210) and using Eqs. (2.69) and (2.68),
we have

ωx = −α̇sinφ + γ̇, ωy = α̇cosφ sin γ + φ̇cosγ, ωz = α̇cosφ cos γ − φ̇sinγ


(2.211)

or we can write
    
 ω  −sinφ 0 1  α̇ 
 x

 
  
   

ωy =  cosφ sin γ cosγ 0 φ̇ (2.212)
 

 
   

 ω
 
 cosφ cos γ −sinγ 0  γ̇ 
 
z

It should be noted that the Euler angle representation for the spacecraft
attitude motion suffers from a major problem of singularity. For example,
using a 3-2-1 Euler angle sequence considering α, φ, γ be the successive
rotation angles, the singularity occurs at φ = ±π/2. In a 3-1-3 Euler angle
sequence, we have singularity at φ=0 or π, whereas, the singularity occurs at
φ = ±π/2 in the case of a 3-1-2 Euler angle sequence.
Example 2.8
Consider the sequence of body rotations that generates the orientation of
coordinate frame B with respect to coordinate frame A: α about the x-axis,
then φ about the new z-axis.
76 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

a) Determine the rotation Matrix for each simple rotation.


b) Determine the complete rotation matrix R13 .
c) Show that if α and φ are very small then the order of rotations does
not matter.
Solution.

(a) The rotation matrix for α about the x-axis is


 
1 0 0
 
R1 (α) = 0 cosα sinα
 
 
0 −sinα cosα

The rotation matrix for φ about the z-axis is


 
cosφ sinφ 0
 
R3 (φ) = −sinφ cosφ 0
 
 
0 0 1

(b) The complete rotation matrix R13 is


  
cosφ sinφ 0 1 0 0
  
R13 (α, φ) = R3 (φ)R1 (α) = −sinφ cosφ 0 0 cosα sinα
  
  
0 0 1 0 −sinα cosα
 
cosφ sinφcosα sinφsinα
 
= −sinφ cosφcosα cosφsinα
 
 
0 −sinα cosα

(c) We obtain the rotation matrix R31 (φ, α) = R1 (α)R3 (φ):

  
1 0 0 cosφ sinφ 0
  
R31 (φ, α) = 0 cosα sinα −sinφ cosφ 0
  
  
0 −sinα cosα 0 0 1
 
cosφ sinφ 0
 
= −cosαsinφ cosαcosφ sinα
 
 
sinαsinφ −sinαcosφ cosα
2.3. RIGID BODY 77

If α and φ are assumed to be very small, i.e., cosα = cosφ=1, sinα=α,


sinφ=φ, and αφ=0, then R13 and R31 are
 
1 φ 0
 
R13 = −φ 1 α = R31
 
 
0 −α 1

Thus, the order of rotation does not matter if α and φ are very small.

2.3.2 Position, Velocity and Acceleration


A rigid body spacecraft can be assumed to be made up of many point masses.
Let a point mass dm be located at a distance of ρ from the center of mass
of the spacecraft (Fig. 2.15). The position of mass dm with respect to an
inertial coordinate frame O − XY Z is
~m = R
R ~ +ρ
~ (2.213)

where R~ defines the position of the center of mass of the spacecraft with
respect to the inertial reference frame, O − XY Z.

Y dm m
ρ S
Rm

θ
O
X

Figure 2.15: Point mass dm of a rigid body spacecraft.

The velocity of mass dm can be expressed as


~˙ m = R
~m = R
V ~˙ + ρ
~˙ (2.214)

The velocity of the spacecraft center of mass is


1 ~m = 1
Z Z
V~cm = V ~˙ + ρ~˙ )dm
(R (2.215)
m m m m
78 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

As the spacecraft center of mass lies at S, we have


Z
ρ
~dm = 0 (2.216)
m

Furthermore, differentiating the preceding relation, we have


Z
~˙ dm = 0
ρ (2.217)
m

Substituting this relation in Eq. (2.215), we have

~˙ = R~˙
 
~cm = R
V +ω ~
~o × R (2.218)
xo y o zo

~˙ o )xo yo zo is the velocity of the spacecraft with respect to S − xo yo zo


where (R
frame and ωo is the inertial angular velocity of the S − xo yo zo frame. For
example, a spacecraft is moving in an orbit with ω ~ o = θ̇k̂ (see Fig. 2.15).
The velocity of the center of mass of the spacecraft, V~cm can be obtained
applying the above relation as
~cm = Ṙîo + θ̇ k̂o × Rîo = Ṙîo + θ̇Rĵo
V (2.219)

The preceding result shows that the velocity of the center of mass of the space-
craft is the same as obtained previously for the case of point mass spacecraft.
Thus, for determining the orbital motion of the rigid-body spacecraft, defined
by the translational motion of its center of mass, we can consider it as a point
mass spacecraft.

2.3.3 Momentum
The linear and angular momentum of the spacecraft are derived as follows.
Linear Momentum
The linear momentum of the spacecraft is
Z Z
p~ = d~
p= ~m dm
V (2.220)
m m

~˙ + ρ
Knowing Vm = R ~˙ , we write
Z
p
~= ~˙ + ρ
(R ~˙ )dm (2.221)
m

Applying the center of mass relation (2.217), we obtain

~˙ = mV
p~ = mR ~cm (2.222)
2.3. RIGID BODY 79

Thus, the linear momentum of the rigid-body spacecraft is the product of the
mass of the spacecraft and the velocity of its center of mass.
Angular Momentum
The angular momentum of the spacecraft is given by
Z Z
H~ = dH~ = ~m × V
R ~m dm (2.223)
m m

Knowing
~ m =R
R ~ +ρ ~
~m =R
V ~˙ m = R
~˙ + ρ

we can write the spacecraft angular momentum as


Z
~
H = (R ~ +ρ ~˙ + ρ
~) × (R ~˙ )dm
m
Z Z Z
=m(R ~˙ + R
~ × R) ~× ~˙ dm +
ρ ρ ~˙ +
~dm × R ρ × ρ~˙ )dm
(~ (2.224)
m m m

Applying the center of mass relations (2.216) and (2.217) yields


Z
˙
~ ~ ~
H = m(R × R) + (~
ρ×ρ ~˙ )dm (2.225)
m

~˙ can be expressed as
Here ρ

~˙ = ρ
ρ ~˙ xyz + ω
~ ×ρ
~ (2.226)

where ρ~˙ xyz represents the velocity of the mass dm with respect to the center
of mass and ω ~ is its inertial angular velocity. The spacecraft is assumed to
be a rigid body and therefore,

~˙ xyz = 0
ρ (2.227)

Substituting Eqs. (2.226) and (2.227) in Eq. (2.225) and applying the triple
cross product relation, we have
Z Z
~ ~ ˙
~ 2
H = m(R × R) + ρ ω~ dm − (~
ρ·ω
~ )~
ρdm (2.228)
m m

Now we express ρ
~ and ω
~ in terms of unit vectors along the rotating coordinate
frame S − xyz as

ρ
~ = xî + y ĵ + z k̂ (2.229)
ω
~ = ωx î + ωy ĵ + ωz k̂ (2.230)
80 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

Substituting ρ
~ and ω ~ from the preceding equations into Eq. (2.228), we have
Z
~ ~ ˙
~
H =m(R × R) + (x2 + y 2 + z 2 )(ωx î + ωy ĵ + ωz k̂)dm
m
Z h
− (x2 ωx + xyωy + xzωz )î + (y 2 ωy + yxωx + yzωz )ĵ
m
i
+ (z 2 ωz + zxωx + zyωy )k̂ dm (2.231)

We have the following mass moment of inertia relations:


Z Z Z
Ixx = (y 2 + z 2 )dm; Iyy = (z 2 + x2 )dm; Izz = (x2 + y 2 )dm;
m m m
Z Z Z
Ixy = xydm; Iyz = yzdm; Ixz = xzdm (2.232)
m m m

Referring to these relations, we can also obtain


Iyy + Izz − Ixx Izz + Ixx − Iyy
Z Z
x2 dm = ; y 2 dm = ;
2 2
Zm m
Ixx + Iyy − Izz
z 2 dm = (2.233)
m 2

Using Eqs. (2.232) and (2.233) into Eq. (2.231), we have the angular
momentum as

~ = m(R
H ~˙ + Hx î + Hy ĵ + Hz k̂
~ × R) (2.234)

where

Hx =Ixx ωx − Ixy ωy − Ixz ωz (2.235)


Hy =Iyy ωy − Iyx ωx − Iyz ωz (2.236)
Hz =Izz ωz − Izx ωx − Izy ωy (2.237)

Alternatively, we can write

~ = m(R
H ~˙ + I~
~ × R) ω (2.238)

where I, called the inertia tensor or inertia matrix and ω


~ , the angular velocity
vector are
   
Ixx −Ixy −Ixz ωx
   
I = −Iyx Iyy −Iyz  , ω ~ =  ωy  (2.239)
   
   
−Izx −Izy Izz ωz
2.3. RIGID BODY 81

In the case the rigid body possesses symmetrical mass distributions, Ixy =
Iyx , Ixz = Izx , and Iyz = Izy .
Furthermore, we can write the angular momentum of the spacecraft as
~ =H
H ~o + H
~b (2.240)

where
~ o = m(R
H ~˙
~ × R), ~ b = Hx î + Hy ĵ + Hz k̂
H (2.241)

~ o is the orbital angular momentum of the spacecraft while H


Note H ~ b is the
attitude angular momentum of the spacecraft. Thus, the angular momentum
of a spacecraft is the sum of its orbital and attitude angular momentum.
For example, a spacecraft is moving in an orbit with ω~ o = θ̇k̂ (see Fig.
2.16). Assuming the body fixed axes x, y, and z be the principal axes, then
the product of inertia terms vanishes, i.e., (Ixy = Ixz = Iyz = 0). The
angular momentum of this spacecraft applying Eq. (2.234) can be obtained
as

~ = mθ̇R2 k̂o + Ixx ωx î + Iyy ωy ĵ + Izz ωz k̂


H (2.242)

Y m
S
x

z
R

O θ
X
.
θ

Figure 2.16: A rigid-body spacecraft in orbit.

Note in the preceding equation, to obtain the magnitude of the angular


momentum, |H|, ~ we need to write the unit vector k̂o in terms of unit vectors
î, ĵ, and k̂ or vice versa applying a transformation matrix as explained in
Section 2.3.1.
82 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

Example 2.9
The system shown in Fig. 2.17 is comprised of a rigid-body spacecraft
m1 and an auxiliary mass m2 (a point mass) orbiting about the Earth. The
auxiliary mass m2 is attached to the spacecraft m1 by a rigid cable of length
L at an offset of ~a = aî from the center of mass of the spacecraft, where î is a
unit vector in the spacecraft body-fixed frame Sxyz . The auxiliary mass m2
undergoes an in-plane libration or oscillation of β about the local vertical.
The mass of the spacecraft is very large in comparison to the auxiliary mass
m2 (i.e., m1  m2 ). Assume the cable is massless and take I as the mass
moment of inertia of m1 about the body-fixed z-axis. Derive the linear and
angular momentum of the system. Note. In the figure, ~a makes an angle of
α with the local vertical.
m2

Orbit L

β
a
S m1
Y
R

θ
E
X

Figure 2.17: A dumbbell satellite system undergoing in-plane libration.

Solution.
The linear momentum of the system is
p~ = p~1 + p
~2 (2.243)
where p~1 and p~2 represent the linear momentum of the spacecraft m1 and
auxiliary mass m2 , given by

p ~˙ 1
~1 = m1 V~1 = m1 R (2.244)
p ~˙ 2
~2 = m2 V~2 = m2 R (2.245)

Here R ~ 1 and R
~ 2 denote the inertial position vectors of m1 and m2 , respec-
tively. Note that the velocity V~1 corresponds to the velocity of the center of
2.3. RIGID BODY 83

mass of m1 .
The inertial velocity vectors of m1 and m2 are

~˙ 1 = R
R ~˙ + ~r˙1 , ~˙ 2 = R
R ~˙ + ~r˙2 (2.246)

where R~ denotes the position vector of the system center of mass with respect
to the Earth center while ~r1 and ~r2 specify the relative position vectors of
m1 and m2 , respectively with respect to the system center of mass.
From the center of mass relation, we have

m1~r1 + m2~r2 = 0 (2.247)

Using Eqs. (2.244)-(2.247), we obtain the linear momentum of the system


as

~˙ = (m1 + m2 )ωo Rĵo


p~ = (m1 + m2 )R (2.248)

Now we consider the given case of m1  m2 . As per Eq. (2.247), we have


~r1 = 0. Thus the linear momentum of the system simplifies to

p
~ = m1 ωo Rĵo (2.249)

The angular momentum of the system is


~ =H
H ~1 + H
~2 (2.250)

where H~ 1 and H
~ 2 represent the angular momentum of the spacecraft m1 and
auxiliary mass m2 and are expressed as

~ 1 = m1 (R
H ~˙ 1 ) + I~
~1 × R ω, ~ 2 = m2 (R
H ~˙ 2 )
~2 × R (2.251)

Here I denotes the moment of inertia of the spacecraft m1 about the body-
fixed z axis.
Using the preceding equations and considering center of mass relation
(2.247), we obtain the angular momentum of the system as

~ = (m1 + m2 )(R
H ~˙ + m2 (~r2 × ~r˙2 ) + I~
~ × R) ω (2.252)

Here the second term is written applying the triple vector product formula
~a × (~b ×~c) = (~a ·~c)~b − (~a ·~b)~c and knowing (L ~ ·ω
~ )=0, (~a · ω
~ L )=0, and (~a · ω
~ )=0
(since ω ~ ω
~ ⊥ L, ~ L ⊥ ~a, and ω ~ ⊥ ~a) as follows.

~r2 × ~r˙2 = (~a + L) ~˙ = L2 ω


~ × (~a˙ + L) ~ · ~a)~
~ L + (L ~ ωL + (~a · ~a)~
ω + (~a · L)~ ω
84 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

Using the coordinate transformations


       
 î L
  î o
  î   î 
 o

 
 
 
    
      
  

ĵL = Rz (β) ĵo , ĵ = Rz (α) ĵo

 
 
 
 
 
 
 

 k̂ 
   k̂   k̂ 
     k̂
 

L o o

the angular momentum of the system is obtained as


~ =(m1 + m2 )R2 ω
~ o + m 2 a2 ω ωL + L2 ω
 
H ~ + aLcos(α − β)~
ω + aLcos(α − β)~ ~L
+ Iω k̂ (2.253)
where ωo = θ̇, ω = θ̇ + α̇, and ωL = θ̇ + β̇.
Remark. For the case where the system center of mass is assumed to
lie on one of the bodies of the system, never consider m1  m2 or ~r1 = 0
before arriving at the total linear momentum of the system (i.e., Eq. (2.247))
otherwise we will get erroneous terms involving R in the final expression for
linear or angular momentum of the system.

2.3.4 Energy
In this section, we derive the potential and kinetic energies of a rigid body
spacecraft (Fig. 2.15).
Potential Energy
The potential energy of a spacecraft is given by
Z
U= dU (2.254)
m
where dU is the potential energy of the mass dm (see Fig. 2.15) expressed as
µdm
dU = − (2.255)
~ m|
|R
Here the position of the mass dm is
~m = R
R ~ + ρ~ (2.256)
or
~ m | = |R
|R ~ +ρ ~ ·ρ
~| = [R2 + 2R ~ + ρ~2 ]1/2 (2.257)
~ m | in Eq. (2.255) and using Eq. (2.255), we have
Substituting |R
µdm
Z
U =−
2 ~
m (R + 2R · ρ ~+ρ ~2 )1/2
" #−1/2
µ
Z
2R~ ·ρ
~ ρ2
=− 1+ + 2 dm (2.258)
R m R2 R
2.3. RIGID BODY 85

Applying Binomial series expansion,


∞  
n
X n n
(1 + x) = x if |x| < 1 (2.259)
k
k=0

where
 
n n n(n − 1) · · · (n − k + 1)
= = (2.260)
k k!(n − k)! 1 · 2···k

for the term inside the integration, we get


Z " ~ ·ρ ~ ·ρ
#
µ R ~ 1 ρ2 3 (R ~)2
U =− 1− − + + · · · dm (2.261)
R m R2 2 R2 2 R4

Now, considering the fact ρ  R and carrying out expansion upto O(1/R3 ),
we get
Z " ~ · ρ~ 1 ρ2 ~ ·ρ
#
µ R 3 (R ~)2 1
U =− 1− − + + O( 3 ) dm
R m R2 2 R2 2 R4 R
" #
µ 1 1 3
Z Z Z
=− m− 2 ~
(R · ρ
~)dm − 2
ρ dm + ~
(R · ρ 2
~) dm
R R m 2R2 m 2R4 m
(2.262)
As the center of mass lies at S, we have
Z
ρ~dm = 0 (2.263)
m

~ and ρ
R ~ are defined with respect to the orbital coordinate frame S − xo yo zo
and the satellite body-fixed coordinate frame S − xyz, respectively as
~ = Rîo
R (2.264)
ρ~ = xî + y ĵ + z k̂ (2.265)

~ and ρ
Substituting R ~ in Eq.(2.262) and applying Eq. (2.263), we obtain
(
µ 1
Z
U =− m− (x2 + y 2 + z 2 )dm
R 2R2 m
3
Z
+ [x2 (îo · î)2 + y 2 (îo · ĵ)2 + z 2 (îo · k̂)2
2R2 m
)
+ 2xy(îo · î)(îo · ĵ) + 2xz(îo · î)(îo · k̂) + 2yz(îo · ĵ)(îo · k̂)]dm

(2.266)
86 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

Substituting mass moment of inertia Eqs. (2.232) and (2.233)) into Eq.
(2.266), we obtain the potential energy of the spacecraft as
(
µm µ
U =− − (Iyy + Izz − Ixx )[3(îo · î)2 − 1]
R 4R3
+ (Izz + Ixx − Iyy )[3(îo · ĵ)2 − 1]
+ (Ixx + Iyy − Izz )[3(îo · k̂)2 − 1]
+ 12Ixy (îo · î)(îo · ĵ) + 12Ixz (îo · î)(îo · k̂)
)
+ 12Iyz (îo · ĵ)(îo · k̂) (2.267)

In the case that the spacecraft body-fixed axes x, y, and z are the principal
axes, then the product of inertia terms vanish, i.e., (Ixy = Ixz = Iyz = 0).
The potential energy of the spacecraft can be expressed as
(
µm µ
U =− − (Iyy + Izz − Ixx )[3(îo · î)2 − 1]
R 4R3
+ (Izz + Ixx − Iyy )[3(îo · ĵ)2 − 1]
)
2
+ (Ixx + Iyy − Izz )[3(îo · k̂) − 1] (2.268)

For a particular case of the Euler angle rotation sequence 3-2-1, we can find
(îo · î), (îo · k̂), and (îo · k̂) as described in Section 2.3.1. The potential energy
of the spacecraft is
(
µm µ
U =− − (Iyy + Izz − Ixx )[3(cosαcosφ)2 − 1]
R 4R3
+ (Izz + Ixx − Iyy )[3(cosαsinφsinγ − sinαcosγ)2 − 1]
)
+ (Ixx + Iyy − Izz )[3(cosαsinφcosγ + sinαsinγ)2 − 1] (2.269)

Kinetic Energy
The kinetic energy of the spacecraft is
Z
T = dT (2.270)
m

where dT is the kinetic energy of the mass dm given by


1 ~˙ 2
dT = R dm (2.271)
2 m
2.3. RIGID BODY 87

~˙ m = R
Using R ~˙ + ρ
~˙ , we can write the kinetic energy of the spacecraft as

1
Z
T = ~˙ 2 + ρ~˙ 2 + 2R
(R ~˙ · ρ
~˙ )dm (2.272)
2 m

~˙ dm=0), we
R
Knowing the center of mass of the spacecraft lies on S, (i.e., m
ρ
have
1 1 ~˙ 2 1
Z Z
T = ~˙ 2 + ρ
(R ~˙ 2 )dm = mR + ρ~˙ 2 dm (2.273)
2 m 2 2 m

~˙ is
where ρ

~˙ = ρ
ρ ~˙ xyz + ω
~ ×ρ
~ (2.274)

As the spacecraft is a rigid body,

~˙ xyz = 0
ρ (2.275)

Substituting the preceding relations in Eq. (2.273) and taking

~ = Rîo
R (2.276)
ρ~ = xî + y ĵ + z k̂ (2.277)

we obtain
1 ~˙ 2 1
Z
T = mR + {(−yωz + zωy )2 + (−zωx + xωz )2 + (−xωy + yωx )2 ]dm
2 2 m
1 ~˙ 2 1
Z
= mR + [(y 2 + z 2 )ωx2 + (z 2 + x2 )ωy2 + (x2 + y 2 )ωz2
2 2 m
− 2xyωx ωy − 2yzωy ωz − 2zxωz ωx ]dm (2.278)

Carrying out the integration, we have

1 ~˙ 2 1 h i
T = mR + Ixx ωx2 + Iyy ωy2 + Izz ωz2 − 2(Ixy ωx ωy + Iyz ωy ωz + Izx ωz ωx )
2 2
(2.279)

or in matrix form
1 ~˙ 2 1 T
T = mR + ω ~ I~
ω (2.280)
2 2
where I is the inertia tensor and ω~ is angular velocity vector (Eq. 2.239).
Note in the preceding equation the first term is the orbital kinetic energy and
the next term is the attitude kinetic energy.
88 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

In the case of the spacecraft body-fixed axes x, y, and z being the principal
axes, then the product of inertia terms vanish, i.e., (Ixy = Ixz = Iyz = 0).
The kinetic energy of the spacecraft is
1 ~˙ 2 1
T = mR + [Ixx ωx2 + Iyy ωy2 + Izz ωz2 ] (2.281)
2 2

For a particular case of the Euler angle rotation sequence 3-2-1, we can
determine the inertial angular velocity ω
~ as

ωx = −(θ̇ + α̇)sinφ + γ̇, ωy = (θ̇ + α̇)cosφ sin γ + φ̇cosγ


ωz = (θ̇ + α̇)cosφ cos γ − φ̇sinγ (2.282)

~˙ can be expressed as
The velocity R

~˙ = (Ṙ2 + θ̇2 R2 )1/2


|R| (2.283)

Substituting Eqs. (2.282) and (2.283) into Eq. (2.281), we can obtain the
kinetic energy of the spacecraft.
Example 2.10
Determine the potential and kinetic energies of the system described in
Example 2.9.
Solution.
Using the Summary Sheet (System: One Rigid Body and One Point Mass,
Section 2.5 and taking ~r1 = 0, we can write the kinetic and potential energies
of the system as
1 ~˙ 2 + 1 m2~r˙22 + 1 Iz ω 2
T = (m1 + m2 )R (2.284)
2 2 2
µ(m1 + m2 ) µ 2 3µ ~ · ~r2 )2
U =− + m r −
3 2 2
m (R
5 2
R 2R 2R
µ n o
+ (Ix + Iy + Iz ) − 3[Iz + (I y − Ix )cos2α] (2.285)
4R3
where ω = θ̇ + α̇ is the angular velocity of the spacecraft m1 . The term ~r2
denotes the position vector of m2 with respect to the system center of mass
S and can be written as
~
~r2 = ~a + L, ~˙
~r˙2 = ~a˙ + L (2.286)

With respect to coordinate frames, we have

~˙ = Ṙîo + Rθ̇ĵo ,
R ~a˙ = ωaĵ, ~˙ = u̇îL + ωL LĵL
L (2.287)

where ωL = θ̇ + β̇ and the frames îo ĵo k̂o , îĵ k̂, and îL ĵL k̂L denote orbital refer-
ence frame, satellite body-fixed frame (with î along ~a), and cable fixed frame
2.3. RIGID BODY 89

(with îL along cable length), respectively. Thus, the kinetic and potential
energies can be rewritten as
1 1 n o
T = M [Ṙ2 + θ̇2 R2 ] + m2 ω 2 a2 + u̇2 + ωL 2 2
L + 2ωaĵ · [u̇îL + ωL LĵL ]
2 2
1 2
+ Iz ω (2.288)
2
µM µ n
2 2 2
o
U =− + m 2 a + L + 2aL(î · î L ) − 3[a(i o · î) + L(i o · î L )]
R n 2R3
µ o
+ (Ix + Iy + Iz ) − 3[Iz + (Iy − Ix )cos2α] (2.289)
4R3
where M = m1 + m2 . Using the coordinate transformations:
       
 îL   îo   î   î 
 o

 
 
 
 
 
 
 

      
ĵL = Rz (β) ĵo , ĵ = Rz (α) ĵo

 
 
 
 
 
 
 

 k̂ 
   k̂   k̂ 
     k̂
 

L o o

or

î = cosαîo + sinαĵo , ĵ = −sinαîo + cosαĵo (2.290)


îL = cosβ îo + sinβ ĵo , ĵL = −sinβ îo + cosβ ĵo (2.291)
îL = cos(α − β)î − sin(α − β)ĵ, ĵL = sin(α − β)î + cos(α − β)ĵ (2.292)

lead to the kinetic and potential energies as


1 1 n
T = M [Ṙ2 + θ̇2 R2 ] + m2 ω 2 a2 + u̇2 + ωL 2 2
L + 2ωa[−u̇sin(α − β)
2 2
o 1
+ ωL Lcos(α − β)] + Iz ω 2 (2.293)
2
µM µm2 n 2 2 2
o
U =− + a + L + 2aLcos(α − β) − 3[acosα + Lcosβ]
R n 2R3
µ o
+ (Ix + Iy + I z ) − 3[Iz + (Iy − Ix )cos2α] (2.294)
4R3
Here u̇ = 0 as the cable is rigid.
Example 2.11
Fig. (2.18) shows a spacecraft with a two-link manipulator undergoing
in-plane motion. The system center of mass is denoted by S. The spacecraft
center of mass is represented by S1 while S2 denotes the center of mass of the
payload. In Fig. (2.19), the center of masses of Links 1 and 2 are denoted
by SL1 and SL2 , respectively. Sj1 and Sj2 are revolute joints while Sj3 is a
fixed joint. The masses of the joints and the motors attached to them are
assumed to be negligible and therefore, not considered in this problem. The
various coordinate frames are shown in Fig. (2.20).
90 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

Link 2
S2 (mL2 )
β2
Payload L2 Link 1
(m2 ) (mL1 )
L1
β1
Orbit

S1 Spacecraft
R S ( m1)
Y

θ
E
X

Figure 2.18: Spacecraft with two-link manipulator undergoing in-plane mo-


tion.

S2 a2 Link 2
Sj (mL2 )
3
SL2 L2
Payload
(m2 ) Sj
2

Link 1
L1 (mL1 )
SL 1
r2 rL2
Sj
1

rL1
a1
S1 Spacecraft
( m1)
r1
S

Figure 2.19: System center of mass locations.

Determine the following:


2.3. RIGID BODY 91

(a) position vector of the Link-2 with respect to the system center of mass.

(b) kinetic energy of the Link-2 assuming the mass of the spacecraft is much
larger than the masses of other bodies in the system and thereby, the
system center of mass S coincides with the center of mass of the satellite,
S1 . The moment of inertia of the link about the axis perpendicular to
the plane of motion is IL2 . Further assume the offsets on the satellite
and payloads are ~a1 = a1 î1 , and ~a2 = −a2 î2 where î1 and î2 are unit
vectors of the coordinate frames attached on the satellite and payload
center of masses S1 and S2 , respectively. The unit vector ~i2 k ~iL2 .

xL2 Link 2
Sj3 (mL2 )
S2
a2 xL1
SL2 β2
Payload L2
(m2 ) Sj2
Link 1
(mL1 )
L1
β x1
yL SL 1 1
2
S j1 x1
α
xo
yL a1
1
yo y1 S1
xo
S Spacecraft
( m1)
Y
R

θ
E
X

Figure 2.20: System various coordinate frames.

Solution.

(a) Let ~r1 , ~rL1 , ~rL2 , and ~r2 denote the position vectors of the center of
mass of spacecraft m1 , and links L1 and L2 , respectively. We can write
92 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

the center of mass relation as

m1~r1 + mL1 ~rL1 + mL2 ~rL2 + m2~r2 = 0 (2.295)

Knowing
1~ 1~ 1~
~rL1 =~r1 + ~a1 + L 1, ~rL2 = ~rL1 + L 1 + L2 (2.296)
2 2 2
1~
~r2 =~rL2 + L2 − ~a2 (2.297)
2
we have
1~
(m1 + mL1 + mL2 + m2 )~r1 + (mL1 + mL2 + m2 )(~a1 + L 1)
2
1 ~ 2 ) + m2 ( 1 L
~1 + L ~ 2 − ~a2 ) = 0 (2.298)
+ (mL2 + m2 )(L
2 2
or
1 h 1 ~1
~r1 = − (mL1 + mL2 + m2 )~a1 + (mL1 + 2mL2 + 2m2 )L
M 2
1 ~ 2 − m2~a2
i
+ (mL2 + 2m2 )L
2
(2.299)

Taking
mL 1 + mL 2 + m2 1 mL1 + 2mL2 + 2m2
γ1 = , γ2 =
M 2 M
1 mL2 + 2m2 m2
γ3 = , γ4 = , M = m1 + mL 1 + mL 2 + m2
2 M M
we have
h i
~ 1 + γ3 L
~r1 = − γ1~a1 + γ2 L ~ 2 − γ4~a2 (2.300)

Thus, the position vector of the link-2, ~rL2 is


1~ 1~ ~1 + 1L~2
~rL2 = ~rL1 + L 1 + L2 = ~
r1 + ~a1 + L (2.301)
2 2 2
or

~ 1 + ( 1 − γ3 )L
~rL2 = (1 − γ1 )~a1 + (1 − γ2 )L ~ 2 + γ4~a2 (2.302)
2

(b) The kinetic energy of the link-2 is the sum of the orbital kinetic energy
(TL2o )and the rotational kinetic energy (TL2b ), i.e.,
1 ~˙ L
2 1 2
TL2 = TL2o + TL2b = mL 2 R + IL2 ωL (2.303)
2 2
2 2
2.3. RIGID BODY 93

where
~˙ L2 =R
R ~˙ + ~r˙L2

=R ~˙ 1 + ( 1 − γ3 )L
~˙ + (1 − γ1 )~a˙ 1 + (1 − γ2 )L ~˙ 2 + γ4~a˙ 2 (2.304)
2

~˙ 1 , L
The vectors ~a˙ 1 , L ~˙ 2 , and ~a˙ 2 are given by
 
~a˙ 1 = ~a˙ 1 +ω~ 1 × ~a1 (2.305)
x1 y 1 z1

~˙ 1 = L ~˙ 1
 
L +ω~ L1 × L~1 (2.306)
xL1 y L1 zL1

~˙ 2 = L ~˙ 2
 
L +ω~ L2 × L~2 (2.307)
xL2 y L2 zL2
 
~a˙ 2 = ~a˙ 2 +ω~ 2 × ~a2 (2.308)
x2 y 2 z2
 
Knowing ~a˙ 1 = 0, ω
~ 1 = (θ̇ + α̇)k̂1 , and ~a1 = (a1 î1 + b1 ĵ1 + c1 k̂1 ),
x1 y 1 z1
we have

~a˙ 1 =~
ω1 × ~a1 = (θ̇ + α̇)k̂1 × (a1 î1 + b1 ĵ1 + c1 k̂1 )
=(θ̇ + α̇)(a1 ĵ1 − b1 î1 ) (2.309)

Similarly, using the relations


ω
~ L1 = (θ̇ + α̇ + β̇1 )k̂L1 , ω
~ L2 = (θ̇ + α̇ + β̇1 + β̇2 )k̂L2
ω
~ 2 = (θ̇ + α̇ + β̇1 + β̇2 )k̂2 (2.310)
we have
~˙ 1 =(θ̇ + α̇)L1 ĵL1 , L
L ~˙ 2 = (θ̇ + α̇ + β̇1 + β̇2 )L2 ĵL1 (2.311)
~a˙ 2 =(θ̇ + α̇ + β̇1 + β̇2 )k̂2 × (a2 î2 + b2 ĵ2 + c2 k̂2 )
=(θ̇ + α̇ + β̇1 + β̇2 )(a2 ĵ2 − b2 î2 ) (2.312)

Knowing R~˙ = Ṙîo + Rθ̇ĵo , we rewrite R~˙ L2 as


"
˙
~
RL2 = Ṙîo + Rθ̇ĵo + (1 − γ1 )(θ̇ + α̇)(a1 ĵ1 − b1 î1 ) + (1 − γ2 )(θ̇ + α̇)L1 ĵL1

1
+ ( − γ3 )(θ̇ + α̇ + β̇1 + β̇2 )L2 ĵL1 + γ4 (θ̇ + α̇ + β̇1
2 #
+ β̇2 )(a2 ĵ2 − b2 î2 ) (2.313)
94 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

Thus, the kinetic energy of link-2 is


"  2
1 L2
TL2 = mL2 (ωo R)2 + (ω1 a1 )2 + (ωL1 )2 + ωL2
2 2
+ 2ω1 ωL1 a1 L1 cosβ1 + ω1 ωL2 a1 L2 cos(β1 + β2 )
#
1 2
+ ωL1 ωL2 L1 L2 cosβ2 + IL2 ωL (2.314)
2 2

where

ωo = θ̇, ω1 = (θ̇ + α̇), ωL1 = θ̇ + α̇ + β̇1 ,


ωL2 = θ̇ + α̇ + β̇1 + β̇2 , ω2 = ωL2

2.4 Flexible Body


In the previous Section 2.3, we have discussed the kinematics of a rigid-body
system wherein the relative distance between any two points on the rigid
body remains fixed. However, in practice, there exist no such ideal rigid
bodies.
A flexible body is a system of point masses, not fixed relative to each
other (see Fig. 2.21, the distance ρ does not remain fixed, i.e., ρ̇ 6= 0, ρ̈ 6= 0)
and consequently, the body undergoes deformation and its elements oscillate.
There are several types of deformation depending upon the direction and lo-
cation of the external forces or torques acting or acted on the body. In the
case where there is no permanent or residual deformation when the external
load is entirely removed, the deformation is called elastic deformation. How-
ever, in certain situations the deformation remains permanent and we call
this plastic deformation. In this section, we will be mainly concerned with
elastic deformation.

m
dm
ρ
dm

Figure 2.21: A flexible body spacecraft.


2.4. FLEXIBLE BODY 95

2.4.1 Defining Deformation


Space systems can consist of different bodies such as rods, strings, beams,
plates, shells or columns. When acted upon by forces or torques, all of
these bodies undergo deformation differently. We will focus our study on
the deformation of rods, strings and beams. The deformation of plates or
columns will not be discussed.
Deformation in Rods or Strings
A body such as rod or string can undergo axial deformation (i.e., along
the axis of the body-fixed frame). Figure 2.22 shows a rod undergoing de-
formation along the longitudinal axis (i.e., the longest axis or minimum area
moment of inertia axis). Such deformation is called longitudinal deformation.
Here Lo is undeformed length while u is the longitudinal deformation. The
position vector of right side of the rod is
~r = (Lo + u)î (2.315)
where î is the unit vector along the x-axis. The corresponding stress and
Y

O
X

Lo u

Figure 2.22: A rod undergoing longitudinal deformation.

strain induced in the rod are given by


Force F
σ= = (2.316)
Area of Cross-Section A
Deformation u
= = (2.317)
Initial length Lo
The above stress and strain are known as Normal stress and Normal or
Longitudinal strain, respectively. They are related by a coefficient called
Young’s modulus of elasticity E, defined as
Normal Stress F/A F Lo
E= = = (2.318)
Normal Strain u/Lo Au
96 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

The above coefficient is known for various materials of the rod (refer to
Handbook on Materials). The preceding expression is called Hooke’s Law
and is only valid within the elastic limit of the material of the rod. Note the
force is acting normal to the cross-section of the body (Fig. 2.22).
A body can undergo three-dimensional axial deformation (i.e., along all
three axes of the body-fixed frame). The deformation along transverse axes
are called transverse deformation. Let us derive the strain of the body in this
situation (Fig. 2.22). We consider an element length dx of a body undergoing
deformation of du along the x-axis (i.e., longitudinal direction) while dv and
dw along the y and z axes, respectively (i.e., transverse directions). The final
length of the element, denoted by d~s, is

d~s = (dx + du)î + dv ĵ + dwk̂ (2.319)

where the unit vectors î, ĵ, and k̂ are along x, y, and z axes, respectively.
The corresponding magnitude is
1/2
ds = (dx + du)2 + dv 2 + dw2

(2.320)

dw
ds C dv
O
x A dx B du X

Figure 2.23: Three-dimensional axial deformation.

The strain in the element dx is expressed as

ds − dx ds 1/2
− 1 = (1 + ux )2 + vx2 + wx2

= = −1 (2.321)
dx dx

where ux = du/dx, vx = dv/dx, and wx = dw/dx.


Assuming the deformations in all directions are very small ( i.e., ux  1,
2.4. FLEXIBLE BODY 97

vx  1, zx  1) and applying Binomial series expansion, we have


1/2
 = 1 + 2ux + u2x + vx2 + wx2

−1
"
1 1 1 1
= 1 + ux + u2x + vx2 + wx2 − (2ux + u2x + vx2 + wx2 )2
2 2 2 8
#
1
+ (2ux + u2x + vx2 + wx2 )3 · · ·
8
" #
1 2 1 2 3 3 1 2 1 2
= 1 + ux + vx + wx − ux − ux vx − ux wx + · · · (2.322)
2 2 2 2 2

Considering upto the second order terms of the deformation (ux , vx and wx ),
we express the strain as
 2  2
∂u 1 ∂v 1 ∂w
= + + (2.323)
∂x 2 ∂x 2 ∂x
Note here we use partial differentials instead of exact differentiation for the
deformation variables as the deformations u, v and w not only depend upon
position x on the body but also on time t (i.e., u = u(x, t), v = v(x, t), w =
w(x, t)). The lateral and longitudinal strains follow a relationship defined by
Poisson’s ratio as
Lateral strain
Poisson’s ratio (ν) = (2.324)
Longitudinal strain
For most metals the value of ν lies in the range 0.25 to 0.35.
The deformations of the body discussed above can be discretized using
assumed modes for each component of the body, i.e.,
n
X
u(x, t) = φi (x)Ui (t) (2.325)
i=1
Xn
v(x, t) = ϑi (x)Vi (t) (2.326)
i=1
Xn
w(x, t) = ψi (x)Wi (t) (2.327)
i=1

where φi , ϑi , and ψi are admissible functions or shape functions while Ui , Vi ,


and Wi are generalized coordinates. The term n denotes number of modes.
The shape functions φi , ϑi , and ψi , are assumed depending upon geometric
and natural spatial boundary conditions. The geometric boundary conditions
refer to deflection or rotation experienced by the body where as the natural
boundary conditions refer to shear force or bending moment acting on the
98 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

body. A list of shape functions for cables, strings, and beams are given in
Section. Summary.
The rod can undergo torsional deformation (i.e., angular twist of elastic
bodies) as illustrated in Fig. 2.24. This deformation results in shear stress
(τ ) and shear strain (γ). These are related by Hooke’s law in shear (linear
elastic region) given by
Shear Stress (τ )
G= (2.328)
Shear Strain (γ)
where G is the shear modulus of elasticity or modulus of rigidity. The moduli
of elasticity in tension and shear are related as
E
G= (2.329)
2(1 + ν)

A B
O C
φ X

Figure 2.24: Rod or Beam undergoing torsion deformation.

In most studies with slender beams, the out-of-plane (transverse) shear


induced warping is usually neglected, but the torsion induced warping is
used to account for its influence on the torsional rigidity. Poisson’s effect is
generally very small in slender beams and hence is also neglected.
Deformation in Beams
A beam undergoes bending deformation when acted upon by forces or
moments that lie in a plane containing its longitudinal axis (see Fig. 2.25).
The forces are assumed to act perpendicular to the longitudinal axis and the
plane containing the forces is considered to be a plane of symmetry of the
beam. If couples are applied to the ends of the beam and no forces act on
the bar, then the bending is termed as pure bending. The stress developed in
the beam is known as bending stress. The bending produced by forces that
do not form couples is called ordinary bending.
2.4. FLEXIBLE BODY 99

Y
S

Undeformed Beam
ρ
M M
O
X
v
Neutral Surface
Deformed Beam
Z

Neutral Axis

Beam Cross-section

Figure 2.25: Beam undergoing bending deformation.

In the pure bending situation, the beam is subjected to only normal


stresses with no shear stresses and negligible rotary inertia. The Euler-
Bernoulli beam theory is used to explain the deformation of the beam. There
always exists one surface in the beam containing fibers that do not undergo
any deformation. This surface is called the neutral surface or neutral plane
of the beam. The intersection of the neutral surface with any cross section
of the beam perpendicular to the longitudinal axis is called the neutral axis.
Note that the neutral axis passes through the centroid of the area of the
cross-section. All fibers on one side of the neutral axis are in a state of ten-
sion, while those on the opposite side are in compression. A plane section of
the beam normal to its longitudinal axis prior to loading remains plane after
the forces and couples have been applied.
From elementary beam theory the moment-curvature equation is
M 1 d2 v σ
= = 2 = (2.330)
EI ρ dx y
where M denotes the bending moment acting at a particular cross section
of the beam. E is the Young’s modulus of the beam. I denotes the area
moment of inertia of the cross section about the centroidal axis. σ is the
normal stress along y-axis on the beam cross-section at y distance from the
neutral axis. The radius of curvature ρ of the neutral surface of the beam is
given by
1 d2 y/dx2
= (2.331)
ρ [1 + (dy/dx)2 ]3/2

Vertical force at any cross section of the beam is called shear force and is
100 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

expressed as

dM
Fs = (2.332)
dx

The equation of beam flexure (v in the y direction) in the absence of


external loading is given by

∂2v ∂2v
 

2
EIz 2
+ρ 2 =0 (2.333)
∂x ∂x ∂t

where ρ represents the beam mass per unit length and EIz is the flexural
rigidity in the y direction. An analogous expression can be written for beam
deflection in the z direction.
The natural boundary conditions are

∂2v
My (Bending Moment) = EIz (2.334)
∂x2
∂3v
Fy (Shear Force) = EIz 3 (2.335)
∂x
where directions are indicated by the subscripts. Similar expressions can be
written for flexure in the z direction.
In the case of transverse vibration of a long, thin beam, it is valid to
ignore shear deformation and rotatory inertia. However, for shorter, stubbier
beams, such effects are significant. The beam flexure including shear de-
formation and rotatory inertia is described by Timoshenko beam Theory[1].
Figure 2.26 shows the kinematics of deformation of a beam which undergoes
shear deformation in addition to pure bending. α(x, t) is the total transverse
displacement of the neutral axis of the beam.

Y
φ α
v
M Neutral Axis
Fs
O
X

Figure 2.26: Beam undergoing bending and shear deformation.


2.4. FLEXIBLE BODY 101

The shear angle is


∂v
β =α− (2.336)
∂x

From elementary beam theory the moment-curvature equation is

M = EIα0 (2.337)

2.4.2 Energy
In this section, we derive the potential and kinetic energies of a flexible sys-
tem. In addition to potential energy due to gravitational force (or gravita-
tional potential energy) that was discussed previously, there exists potential
energy due to the elastic strain energy stored in the flexible body . Let us
denote gravitational potential energy by Ug and elastic potential energy by
Ue . Thus, the potential energy of the system is

U = Ug + Ue (2.338)

Note that in some cases, the contribution of Ug to the system potential en-
ergy is much smaller in comparison to Ue and therefore, it can be assumed
negligible.
The kinetic energy of a flexible system can be written as

T = Tr + Te (2.339)

where Tr is kinetic energy due to the rigid part of the flexible body whereas
Te refers to the kinetic energy due to the flexible part the system.
Under various deformations the potential and kinetic energies due to flex-
ibility are as follows:
Axial deformation (Fig. 2.23)

1 1
Ue = mean force × deformation = F u = E(x)A(x)Lo 2 (2.340)
2 2
Here the force F is applied on the area of cross-section A of a flexible body of
undeformed length Lo and as it varies linearly from 0 to F , the deformation
of the body changes linearly from 0 to x within the elastic limit of the body.
For an element dx, the system potential and kinetic energies are

1 L F2 1 L
Z Z
Ue = dx = E(x)A(x)2 dx (2.341)
2 0 E(x)A(x) 2 0
1 L 2
Z
Te = ρu̇ dx (2.342)
2 0
102 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

where dx is undeformed length. u(x, t) is the displacement of the cross-section


along the axial direction. A(x) denote the cross-sectional area at x distance
with respect to the body-fixed coordinate frame. ρ is mass per unit length
of the rod. L denotes the length of the rod.
Torsional or shear deformation (Fig. 2.24)

L 2
T2 1 L

1 ∂φ(x, t)
Z Z
Ue = dx = GIp dx (2.343)
2 0 G(x)Ip (x) 2 0 ∂x
2
1 L

∂φ(x, t)
Z
T = ρ(x) dx (2.344)
2 0 ∂t

where G is modulus
RR 2 of torsional rigidity, and Ip is polar area moment of
inertia (i.e., r dA) about the centroidal x-axis.
Bending deformation (Fig. 2.25)

L  2 2
1 M2 1 L ∂ v
Z Z
Ue = dx = E(x)I(x) 2
dx (2.345)
2 0 E(x)I(x) 2 0 ∂x
2
1 L

∂v(x, t)
Z
T = ρ(x) dx (2.346)
2 0 ∂t

where I(x) is the cross-sectional area moment of inertia of the variable geome-
try beam about the neutral axis. ρ(x) is mass per unit length. u(x, t) denotes
the transverse deformation (displacement) from the neutral axis. EI(x) is
called flexural rigidity or bending stiffness.
Bending deformation (v and w directions) and torsional deformation

L 2 2
∂ 2 v(x, t) 1 L
  2
1 ∂ w(x, t)
Z Z
Ue = EI dx + EI dx
2 0 ∂x2 2 0 ∂x2
2
1 L

∂φ(x, t)
Z
+ GIp dx (2.347)
2 0 ∂x

Bending deformation and shear deformation (Fig. 2.26


The bending strain energy is
L
1
Z
Ub = EI(α0 )2 dx (2.348)
2 0

The shear strain energy can be expressed as


L
1
Z
Us = kGA(β)2 dx (2.349)
2 0
2.4. FLEXIBLE BODY 103

where the shear coefficient k for a rectangular beam is k = 5/6.


Thus, the potential energy due to flexibility or elastic strain energy is

Ue = Ub + Us (2.350)

The kinetic energy of the beam is

1 L 1 L
Z Z
2
Te = ρA(v̇) dx + ρI(α̇)2 dx (2.351)
2 0 2 0

Example 2.12
In Example 2.4, the cable connecting m1 and m2 is considered to be a
flexible rod undergoing longitudinal deformation u(x, t) with its rigid length
of L. The cable has a mass of ρ per unit length. Consider the fundamental
mode and assuming the rigidity of the cable as EA and m2  m1 , derive the
kinetic and potential energies of the flexible rod.
Solution.
The kinetic and potential energies of the flexible rod are

1 1 L  ˙ ˙
Z   Z
Te = ˙ ˙
~r · ~r dm = ρ ~r · ~r dx (2.352)
2 m 2 0
"  2 #
1 L 1 L ∂u
Z Z
2
Ue = EA dx = EA dx (2.353)
2 0 2 0 ∂x

where ~r denotes the position vector of mass dm of the rod given as

~r = u(x, t)î (2.354)

The longitudinal deformation of the rod with the fundamental mode is


x
u(x, t) = φ1 (x)U = U (2.355)
L

Knowing L̇ = 0,
x
~r˙ (x, t) = U̇ î (2.356)
L

Thus, the kinetic and potential energies of the flexible rod are

1 L x2 2 1
Z
Te = 2
U̇ ρdx = mL U̇ 2 (2.357)
2 0 L 6
Z L 2

1 U 1 EA 2
Ue = EA 2 dx = U (2.358)
2 0 L 2 L
104 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

where mL = ρL.
Example 2.13
In Example 2.4, the cable connecting m1 and m2 is considered to be a
flexible rod undergoing transverse deformation v(x, t) with its rigid length
of L. The cable has a mass of ρ per unit length. Consider the fundamental
mode and assuming the rigidity of the cable as EA and m2  m1 , derive the
kinetic and potential energies of the flexible rod.
Solution.
The kinetic and potential energies of the flexible rod is
1 L  ˙ ˙
Z
Te = ρ ~r · ~r dx (2.359)
2 0
1 L
Z
Ue = EA2 dx (2.360)
2 0
where ~r denotes the position vector of mass dm of the rod given as
~r = xî + v(x, t)ĵ (2.361)
The transverse deformation of the rod with the fundamental mode is
 πx 
v(x, t) = ϑ1 V (t) = sin V (2.362)
L

Knowing L̇ = 0,
 πx 
~r˙ (x, t) = sin V̇ ĵ (2.363)
L
The strain  is given by
 2
1 ∂v
= (2.364)
2 ∂x
where
∂v πV πx
= cos (2.365)
∂x L L
Thus, the kinetic and potential energies of the flexible rod is
1 L πx 1
Z
T = ρV̇ 2 sin2 dx = mL V̇ 2 (2.366)
2 0 L 4
1 L1 π2 V 2 πx
Z
Ue = EA 2 cos2 dx
2 0 4 L L
2 2 Z L 1 + cos 2πx
1 EAπ V L 1 EAπ 2 V 2
= 2
dx = (2.367)
8 L 0 2 16 L
2.4. FLEXIBLE BODY 105

where mL = ρL.
Example 2.14
The system is comprised of a two link flexible manipulator undergoing
planar rigid rotation and bending deformation of Link-1 (L1 ) and Link-2
(L2 ). β1 and β2 describe rigid body rotations while bending deformations
are denoted by v1 (x1 , t) and v2 (x2 , t) on Link-1 and Link-2, respectively. The
coordinate frames S1 − î1 ĵ1 and S2 − î2 ĵ2 are fixed with rigid parts of Link-1
and Link-2 with their lengths L1 and L2 along î1 and î2 , respectively. ρ1 and
rho2 denote the mass per unit length of Link-1 and Link-2. Each link is mod-
eled as a uniform flexible beam. Assuming the Euler-Bernoulli assumptions
of negligible shear deformation and negligible distributed rotatory inertia,
derive the kinetic and potential energies of the system due to flexibility.
Solution.
The kinetic energy of the system is
1 L1 ˙ 2 1 L2 ˙ 2
Z Z
T = Te1 + Te2 = ρ1~rm1 dx1 + ρ2~rm2 dx2 (2.368)
2 0 2 0
where the position vectors of elemental masses are
~rm1 =x1 î1 + v1 (x1 , t)ĵ1 (2.369)
~rm2 =L1 î1 + v1 (L1 , t)ĵ1 + x2 î2 + v2 (x2 , t)ĵ2 (2.370)
Furthermore, we have
~r˙m1 =[−ω1 v1 (x1 , t)]î1 + [ω1 x1 + v̇1 (x1 , t)]ĵ1 (2.371)
~r˙m2 =[−ω1 v1 (L1 , t)]î1 + [ω1 L1 + v̇1 (L1 , t)]ĵ1 +
+ [−ω2 v2 (x2 , t)]î2 + [ω2 L2 + v̇2 (x2 , t)]ĵ2 (2.372)
where ω1 = β̇1 , ω2 = β̇2 .
Squaring the position vectors lead to
~r˙m1
2
=[−ω12 v1 (x1 , t)]2 + [ω1 x1 + v̇1 (x1 , t)]2 (2.373)
~r˙m2
2
=[−ω1 v1 (L1 , t)]2 + [ω1 L1 + v̇1 (L1 , t)]2 + [−ω2 v2 (x2 , t)]2
+ [ω2 L2 + v̇2 (x2 , t)]2 + 2[−ω1 v1 (L1 , t)][−ω2 v2 (x2 , t)]î1 · î2
+ 2[−ω1 v1 (L1 , t)][ω2 L2 + v̇2 (x2 , t)]î1 · ĵ2
+ 2[ω1 L1 + v̇1 (L1 , t)][−ω2 v2 (x2 , t)]ĵ1 · î2 +
+ 2[ω1 L1 + v̇1 (L1 , t)][ω2 L2 + v̇2 (x2 , t)]ĵ1 · ĵ2 (2.374)

Knowing
î1 · î2 =cos(β2 − β1 ), î1 · ĵ2 = −sin(β2 − β1 )
ĵ1 · î2 =sin(β2 − β1 ), ĵ1 · ĵ2 = cos(β2 − β1 )
106 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

we have
~r˙m1
2
=[−ω12 v1 (x1 , t)]2 + [ω1 x1 + v̇1 (x1 , t)]2 (2.375)
~r˙m2
2
=[−ω1 v1 (L1 , t)]2 + [ω1 L1 + v̇1 (L1 , t)]2 + [−ω2 v2 (x2 , t)]2
+ [ω2 L2 + v̇2 (x2 , t)]2 + 2[−ω1 v1 (L1 , t)][−ω2 v2 (x2 , t)]cos(β2 − β1 )
− 2[−ω1 v1 (L1 , t)][ω2 L2 + v̇2 (x2 , t)]sin(β2 − β1 )
+ 2[ω1 L1 + v̇1 (L1 , t)][−ω2 v2 (x2 , t)]sin(β2 − β1 )+
+ 2[ω1 L1 + v̇1 (L1 , t)][ω2 L2 + v̇2 (x2 , t)]cos(β2 − β1 ) (2.376)
Upon substitution of these variables into Eq. (2.368), we find the system
kinetic energy.
The system potential energy is given by
2 2
1 L1 1 L2
 2  2
∂ v1 (x1 , t) ∂ v2 (x2 , t)
Z Z
Ue = E1 I1 dx1 + E2 I2 dx2
2 0 ∂x21 2 0 ∂x22
(2.377)

Example 2.15
In Example 2.9, m1 and m2 are connected by a flexible beam (instead of
a cable) of mass ρ per unit length with rigid length Lo . The beam undergoes
bending deformation of v(x, t). Assume a Euler-Bernoulli Beam. Iz is the
mass moment of inertia of m1 about the body-fixed z-axis and EIL denotes
the modulus of rigidity of the beam. Derive the kinetic energy (system)
and potential energy due to bending deformation. Assume in-plane system
motion and express the solution in terms of bending deformation v(x, t). Note
that α is the rigid-body rotation (pitch angle) of m1 .
Solution.
The system kinetic and potential energies are
1 ~˙ 1 L0  1 n
Z
ρ (ωv)2 + (v̇ + ωx)2 dx + m2 [ωv(L0 , t)]2

T = MR +
2 2 a 2
o 1  2
∂v
+ [v̇(L0 , t) + ωL0 ]2 + Iz ω + (2.378)
2 ∂x x=L
2
1 L0
 2
∂ v(x, t)
Z
Ue = EIL dx (2.379)
2 a ∂x2
where mL = ρL0 , M = m1 + m2 + mL
Example 2.16
For the given problem in Example 2.4, the dumbbell system undergoes
an in-plane libration of β about the local vertical along with longitudinal
deformation u of the cable. Assume the rigidity of the cable as EA, derive
the following:
2.4. FLEXIBLE BODY 107

(a) inertial position, velocity, and acceleration of m1 and m2 .

(b) kinetic and potential energies of the system.

Solution.

(a) As described in Example 2.4, the position vectors of spacecraft m1 and


m2 are

~1 = R
R ~ − γL
~ (2.380)
~2 = R
R ~ + (1 − γ)L
~ (2.381)

where γ = m2 /(m1 + m2 ) and L is given by

~ = (L0 + u)î
L (2.382)

where L0 remains fixed while the deformation u changes (i.e., u̇ 6= 0


and ü 6= 0).
The velocity vectors of spacecraft m1 and m2 are

~˙ − γ L
~1 = R
V ~˙ (2.383)
~˙ + (1 − γ)L
~2 = R
V ~˙ (2.384)

~˙ = Ṙîo + ωo Rĵo and L


where R ~˙ = u̇î + ωLĵ. Here ωo = θ̇ and ω = θ̇ + β̇.
The acceleration vectors of spacecraft m1 and m2 are

~¨ − γ L
~a1 = R ~¨ (2.385)
~¨ + (1 − γ)L
~a2 = R ~¨ (2.386)

where R~¨ = R̈îo + 2ωo Ṙĵo − ω 2 Rîo + ω̇o Rĵo and L


~¨ = üî + 2ω u̇ĵ − ω 2 Lî +
o o
ω̇Lĵ.

(b) Following the procedure outlined in Example 2.7, the kinetic and po-
tential energies of the system are obtained as

1 1
T = M (Ṙ2 + θ̇2 R2 ) + Me [u̇2 + (L0 + u)2 ω 2 ] (2.387)
2 2
µM µ 1 EA 2
U =− + Me (1 − 3cos2 β)L2 + u (2.388)
R 2R3 2 L0

where M = m1 + m2 and Me = m1 m2 /(m1 + m2 ).


108 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

Example 2.17
In Example 2.9, consider the cable connecting m1 and m2 to be flexible
with rigid length Lo and longitudinal deformation u (i.e., L = Lo + u). As-
sume the rigidity of the cable as EA, derive the kinetic and potential energies
of the given system. Take Iz as the mass moment of inertia of m1 about the
body-fixed z-axis.
Solution.
Using Summary Sheet (System: One Rigid Body and One Point Mass),
Section 2.5 and taking ~r1 = 0, we can write the kinetic and potential energies
of the system as

1 ~˙ 2 + 1 m2~r˙ 2 + 1 Iz ω 2
T = (m1 + m2 )R 2 (2.389)
2 2 2
µ(m1 + m2 ) µ 3µ ~ · ~r2 )2
U =− + m2 r22 − m2 (R
R 2R3 2R5
µ n o 1 EA
+ (Ix + Iy + Iz ) − 3[I z + (Iy − I x )cos2α] + u2 (2.390)
4R3 2 L0

where ω = θ̇ + α̇ is the angular velocity of the spacecraft m1 . Note that the


last term in the potential energy expression is due to flexibility of the cable.
The term ~r2 denotes the position vector of m2 with respect to the system
center of mass S and can be written as

~
~r2 = ~a + L, ~˙
~r˙2 = ~a˙ + L (2.391)

With respect to coordinate frames, we have

~˙ = Ṙîo + Rθ̇ĵo ,
R ~a˙ = ωaĵ, ~˙ = u̇îL + ωL LĵL
L (2.392)

where ωL = θ̇ + β̇ and the frames îo ĵo k̂o , îĵ k̂, and îL ĵL k̂L denote orbital refer-
ence frame, satellite body-fixed frame (with î along ~a), and cable fixed frame
(with îL along cable length), respectively. Thus, the kinetic and potential
energies can be rewritten as

1 1 n o
T = M [Ṙ2 + θ̇2 R2 ] + m2 ω 2 a2 + u̇2 + ωL 2 2
L + 2ωaĵ · [u̇îL + ωL LĵL ]
2 2
1 2
+ Iz ω (2.393)
2
µM µ n o
U =− + 3
m2 a2 + L2 + 2aL(î · îL ) − 3[a(io · î) + L(io · îL )]2
R 2R
µ n o 1 EA
+ (Ix + I y + I z ) − 3[Iz + (Iy − Ix )cos2α] + u2 (2.394)
4R3 2 L0
2.5. SUMMARY 109

where M = m1 + m2 . Using the coordinate transformations:

î = cosαîo + sinαĵo , ĵ = −sinαîo + cosαĵo (2.395)


îL = cosβ îo + sinβ ĵo , ĵL = −sinβ îo + cosβ ĵo (2.396)
îL = cos(α − β)î − sin(α − β)ĵ, ĵL = sin(α − β)î + cos(α − β)ĵ (2.397)

lead to the kinetic and potential energies as


1 1 n
T = M [Ṙ2 + θ̇2 R2 ] + m2 ω 2 a2 + u̇2 + ωL 2 2
L + 2ωa[−u̇sin(α − β)
2 2
o 1
+ ωL Lcos(α − β)] + Iz ω 2 (2.398)
2
µM µm2 2 n
2 2
o
U =− + a + L + 2aLcos(α − β) − 3[acosα + Lcosβ]
R 2R3
µ n o 1 EA
+ (I x + I y + Iz ) − 3[Iz + (Iy − Ix )cos2α] + u2 (2.399)
4R3 2 L0

2.5 Summary
In this chapter, we have discussed kinematics of point masses, rigid bodies
and flexible bodies. The energy and momentum of these bodies have been
derived. The important results discussed in this chapter are summarized as
follows.
110 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

2.5.1 Kinematics: Point Mass

System: Single Point Mass

Inertial velocity ~r˙XY Z = ~r˙xyz + ω


~ × ~rxyz
Inertial acceleration ~r¨XY Z = ~r¨xyz + 2(~
ω × ~r˙xyz ) + ω
~ × (~
ω × ~rxyz )
~˙ × ~rxyz

 
1 0 0
 
Rotation Matrix Rx (θ) = 0 cosθ sinθ
 
 
0 −sinθ cosθ

 
cosθ 0 −sinθ
 
Ry (θ) =  0 1 0
 

 
sinθ 0 cosθ

 
cosθ sinθ 0
 
Rz (θ) = −sinθ cosθ 0
 
 
0 0 1
Triple Vector product ~a × (~b × ~c) = (~a · ~c)~b − (~a · ~b)~c
Linear Momentum p
~ = mR ~˙

Angular Momentum ~ = m(R


H ~˙
~ × R)

Kinetic Energy T = 21 mR ~˙ 2
µm
Potential Energy U =− R

System: Two Point Masses(N =2)

Center of Mass Equation m1~r1 + m2~r2 = 0


~ :
If ~r2 = ~r1 + L ~
~r1 = −γ L,
~r2 = (1 − γ)L; ~ γ = m2 /(m1 + m2 )
Linear Momentum p~ = (m1 + m2 )R~˙

Angular Momentum H~ = (m1 + m2 )(R ~ × R) ~˙ + m1 (~r1 × ~r˙1 ) + m2 (~r2 × ~r˙2 )

Kinetic Energy T = 12 (m1 + m2 )R ~˙ 2 + 1 m1~r˙ 2 + 1 m2~r˙ 2


2 1 2 2
µ(m1 + m2 ) µ  2 2

Potential Energy U =− + 3 m 1 r1 + m 2 r2
hR 2R

i
− 5 m1 (R ~ · ~r1 )2 + m2 (R ~ · ~r2 )2
2R
2.5. SUMMARY 111

System: N-Point Masses

PN
Center of Mass Equation i=1mi~ri = 0
~ i−1 , PN −1 ~
If ~ri = ~ri−1 + L ~r1 = − i=1 γi Li
Pk−1 ~ i − PN −1 γi L
~ i,
i = 2, 3, · · · , N : ~rk = i=1 (1 − γi )L i=k

k = 2, 3, · · · , N
PN
mi
γk = M
i=k+1
, k = 1, 2, · · · , N − 1,
PN
M = i=1 mi

PN
Linear Momentum p~ = i=1 mi R
Angular Momentum ~ = PN mi (R
H ~ × R)~˙ + PN mi (~ri × ~r˙i )
i=1 i=1
˙~ 2 1 PN
T = 2 i=1 mi R + 2 i=1 mi~r˙i2
1
PN
Kinetic Energy
µ PN µ PN 2
Potential Energy U = −R i=1 mi + 2R3 i=1 mi ri
3µ P N ~ ri )2
− 2R 5 i=1 mi (R · ~
112 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

2.5.2 Kinematics: Rigid Body

System: Single Rigid Body

Linear Momentum ~˙
p~ = mR
~ = m(R
Angular Momentum H ~˙ + Ix ωx î + Iy ωy ĵ + Iz ωz k̂
~ × R)

3-2-1 Euler Sequence (α, φ, γ) with θ̇ = orbital rate:


ωx = −(θ̇ + α̇)sinφ + γ̇
ωy = (θ̇ + α̇)cosφ sin γ + φ̇cosγ
ωz = (θ̇ + α̇)cosφ cos γ − φ̇sinγ
Kinetic Energy T = 12 mR~˙ 2 + 1 [Ixx ωx2 + Iyy ωy2 + Izz ωz2 ]
2 n

Potential Energy U = − µmR − µ 2


4R3 (Iyy + Izz − Ixx )[3(îo · î) − 1]

+(Izz + Ixx − Iyy )[3(îo · ĵ)2 − 1]


o
+(Ixx + Iyy − Izz )[3(îo · k̂)2 − 1]
3-2-1 Euler Sequence:
n
U = − µm
R −
µ
4R3 (Iyy + Izz − Ixx )[3(cosαcosφ)2 − 1]
+(Izz + Ixx − Iyy )[3(cosαsinφsinγ − sinαcosγ)2 − 1]
o
+(Ixx + Iyy − Izz )[3(cosαsinφcosγ + sinαsinγ)2 − 1]

Planar Motion
Kinetic Energy ~˙ 2 + 1 Iz ωz2 , ωz = θ̇ + α̇
T = 12 mR 2 n
µm µ
Potential Energy U = − R + 4R 3 (Ix + Iy + Iz )
o
−3[Iz + (Iy − Ix )cos2α]

System: Two Rigid Bodies (Planar Motion)

Kinetic Energy ~˙ 2 + 1 m1~r˙ 2 + 1 m2~r˙ 2


T = 12 (m1 + m2 )R 2 1 2 2

+ 12 Iz1 ωz21 + 12 Iz2 ωz22 ; ωz1 = θ̇ + α̇1 , ωz2 = θ̇ + α̇2


Potential Energy
h i
µ(m1 +m2 ) µ 3µ ~ · ~r1 )2 + m2 (R
~ · ~r2 )2
m1 r12 + m2 r22 −
 
U− R + 2R3 2R5 m1 (R
n o
µ
+ 4R 3 (Ix1 + Iy1 + Iz1 ) − 3[Iz1 + (Iy1 − Ix1 )cos2α1 ]
n o
µ
+ 4R 3 (Ix2 + Iy2 + Iz2 ) − 3[Iz2 + (Iy2 − Ix2 )cos2α2 ]
2.5. SUMMARY 113

System: N-Rigid Bodies (Planar Motion)

Kinetic Energy T = 1
PN ~˙ 2 +
mi R 1
PN
mi~r˙i2 + 1
PN 2
2 i=1 2 i=1 2 i=1 Izi ωzi

Potential Energy
µ PN µ PN 2 3µ PN ~ ri )2
U = −R i=1 mi + 2R3 i=1 mi ri − 2R5 i=1 mi (R · ~
n o
µ P N
+ 4R 3 i=1 (Ix i + Iy i + Iz i ) − 3[Iz i + (Iy i − Ix i )cos2αi ]

System: One Rigid Body and One Point Mass (Planar Motion)

Kinetic Energy ~˙ 2 + 1 m1~r˙ 2 + 1 m2~r˙ 2 + 1 Iz1 ω 2


T = 12 (m1 + m2 )R 2 1 2 2 2 z1

Potential Energy
h i
U = − µ(m1R+m2 ) + 2R µ 2 2 3µ
3 [m1 r1 + m2 r2 ] − 2R5 m1 (R~ · ~r1 )2 + m2 (R
~ · ~r2 )2
n o
µ
+ 4R 3 (Ix1 + Iy1 + Iz1 ) − 3[Iz1 + (Iy1 − Ix1 )cos2α1 ]

System: M-Rigid Bodies(m1 , m2 , · · · , mM ) and N-Point


Masses(mN +1, mN +2 , · · · , mM+N )


PM+N
Linear Momentum p~ = i=1 mi R
Angular Momentum ~ =
H
PM+N
mi (R~ ~˙ + PM+N mi (~ri × ~r˙i )
× R)
i=1 i=1
PM h i
+ i=1 Ixi ωxi îi + Iyi ωyi ĵi + Izi ωzi k̂i
Kinetic Energy
PM+N
T = 12 i=1 mi R~˙ 2 + 1 PN +M mi~r˙ 2
2 i=1 i
PM h i
1
+ 2 i=1 Ixi ωxi + Iyi ωyi + Izi ωz2i
2 2

Potential Energy
P M +N
µ mi µ PM+N 3µ PM+N ~ · ~ri )2
mi ri2 − 2R
 
U =− i=1
R + 2R3 i=1 5 i=1 mi (R
n h
µ PM
+ 4R3 i=1 (Ix i + Iy i + I z i ) − 3 (Iyi + Izi − Ixi )(îo · î)2
io
+(Izi + Ixi − Iyi )(îo · ĵ)2 + (Ixi + Iyi − Izi )(îo · k̂)2
Potential Energy
(Planar Motion)
P M +N
µ mi µ PM+N 3µ
 PM+N
~ · ~ri )2
mi ri2 −

U =− i=1
R + 2R3 i=1 2R5 i=1 mi (R
µ PM n o
+ 4R 3 i=1 (Ixi + Iyi + Izi ) − 3[Izi + (Iyi − Ixi )cos2αi ]
114 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

2.5.3 Kinematics: Flexible Body

Deformation in Rod
Axial deformation: longitudinal and transverse deformation;
Torsional deformation
Deformation in Beam
Bending deformation; Shear deformation
Moment-curvature equation for Pure Bending Deformation
M 1 d2 v σ
EI = ρ = dx2 = y
M =bending moment acting at a particular cross section of the beam;
E=Young’s modulus of the beam; I= area moment of inertia of the cross
section about the centroidal axis; σ=normal stress; y=distance from the
neutral axis
Deformations discretized using Assumed Modes
Pn Pn
u(x, t) = i=1 φi (x)Ui (t), v(x, t) = i=1 ϑi (x)Vi (t),
w(x, t) = ni=1 ψi (x)Wi (t)
P

φi , ϑi , and ψi =admissible functions or shape functions assumed depending


upon geometric and natural spatial boundary conditions; Ui , Vi , and
Wi =generalized coordinates; n=number of modes
Admissible or Shape Functions
Tether/Cable
 2i−1
φi (x) = Lx0 Longitudinal
√  
iπx
ϑi (x) = 2sin L0 Transverse
√  
ψi (x) = 2sin iπx L0 Transverse
p
L0 =undeformed or rigid length; (2)normalizingf actor
Beam: Bending deformation (fixed-free end)
h i
cosh(λi L)+cos(λi L)
φi (x) = cosh(λi x) − cos(λi x) − sinh(λ i L)+sin(λi L)
[sinh(λi x) − sin(λi x)]
L=length of the beam; λi =eigenvalue from the characteristic equation:
cos(λi L) cosh(λi L) + 1 = 0
eigenvalues:λ1 L = 1.8751, λ2 L = 4.6941, λ3 L = 7.8548, λ4 L = 10.996
i+1 iπx 2
φi (x) = 1 − cos iπx 1

L + 2 (−1) L
1 ∂v 2
2
 = ∂u + 12 ∂w

Strain ∂x + 2 ∂x ∂x

u = u(x, t)=longitudinal deformation,


v = v(x, t),w = w(x, t)=transverse deformation
RL
Strain energy(axial) Ue = 12 0 E(x)A(x)2 dx
dx=undeformed length
RL  2
Strain energy(torsional) Ue = 12 0 GIp ∂φ(x,t)
∂x dx,
G=modulus of rigidity
RL  2 2
Bending strain energy Ub = 12 0 EI ∂ ∂x
v(x,t)
2 dx
L
Us = 12 0 kGA(β)2 dx, k=shear coefficient
R
Shear strain energy
2.5. SUMMARY 115

Deflectionof a cantilever beam


F x3f  2  3 
y= 2 xx − xx
6EI f f
116 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

References

1. W. Jr. Weaver, S. P. Timoshenko, D. H. Young, Vibration Problems


in Engineering, Fifth Edition, John Wiley & Sons, New York, 1990.

2. Peter W. Likins, Elements of Engineering Mechanics, McGraw-Hill,


New York, 1973.

3. C. Truesdell, Essays in the History of Mechanics, Springer-Verlag, New


York, 1968.

4. John L. Synge and Byron A. Grifth. Principles of Mechanics. McGraw-


Hill, New York, third edition, 1959. edition, 1980.

5. W. T. Thomson. Introduction to Space Dynamics, Dover, New York,


1986.

6. Thomas R. Kane, Peter W. Likins, and David A. Levinson, Spacecraft


Dynamics, McGraw-Hill, New York, 1983.

7. Michael D. Gri and James R. French, Space Vehicle Design, AIAA


Education Series, American Institute of Aeronautics and Astronautics,
Washington, D.C., 1991.

8. Bong Wie, Space Vehicle Dynamics and Control, AIAA, Reston, Vir-
ginia, 1998.

Problem Set 2

2.1 What are different coordinate frames used in studying the dynamics of
a spacecraft? Show these frames with sketches.

2.2 What do you understand by an inertial frame? Why is it so relevant?


Is an Earth-fixed frame an inertial frame?
If an Earth-fixed frame is taken as an inertial frame, what would be an
order of magnitude error in the calculations.

2.3 For the given problem in Example 2.1, the dumbbell satellite system
undergoes an in-plane libration or oscillation of β about the local ver-
tical and the distance L between the satellites m1 and m2 varies with a
constant speed of v m/s. Determine the inertial position, velocity, and
acceleration vectors and their corresponding magnitudes for satellites
m1 and m2 .
2.5. SUMMARY 117

2.4 For the given problem in Example 2.4, if the dumbbell satellite system
undergoes a three-dimensional libration motion with in-plane libration
β about the local vertical followed by an out-of-plane libration of η,
then determine the inertial position, velocity, and acceleration vectors
and their corresponding magnitudes for satellites m1 and m2 .

2.5 For a given system of a spacecraft m1 and two bodies m2 and m3


connected through rigid massless cables of length L1 and L2 (Fig. 2.27),
determine the inertial position, velocity, and acceleration vectors and
their corresponding magnitudes for the spacecraft m1 and two bodies
m2 and m3 if the system undergoes in-plane motion. The masses of the
two bodies are very small in comparison to the mass of the spacecraft
and therefore, the system center of mass is considered to coincide with
the center of mass of the spacecraft m1 .

m3
β2
Orbit L2
Local Vertical
m2
L1 β1
Local Vertical
S m1
Y R

θ
E
X

Figure 2.27: System undergoing in-plane libration.

In case m1 is comparable to m2 or m3 then what would be the positions


of spacecraft m2 and m3 .

2.6 In Problem 2.5, consider the case that the masses of the two bodies
are not small in comparison to the mass of the spacecraft. Then, the
system center of mass does not coincide with the center of mass of the
spacecraft m1 . Determine the position vectors of the spacecraft and
two bodies with respect to the system center of mass. Write the answer
in terms of the respective body-fixed frame.
118 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

2.7 Using Matlab, determine the position, velocity and acceleration of a


spacecraft using a polar coordinate frame with the following details:
orbit: circular
orbital altitude: 400 km (Low-Earth orbit)
If the orbital altitude is changed to 36000 km (≈ Geostationary or-
bit), then what would be the position, velocity and acceleration of the
spacecraft.
p
Note. θ̇ = µ/a3 , where µ=Earth’s gravitational constant=3.986×
105 km3 /s2 ; a=semi-major axis; Earth’s radius=6378 km.

2.8 Solve Problem 2.7 using Matlab Simulink.

2.9 If the spacecraft in Problem 2.7 is orbiting in an elliptic orbit of ec-


centricity (e = 0.02), then what would be the position, velocity and
acceleration of the spacecraft. Show your results using Simulink.
Note. In an elliptic orbit, the position R and the angle θ with respect
to a reference line follow the following relation:

a(1 − e2 )
= 1 + ecosθ
R

2.10 For the given problem in Example 2.4, it is desired to catch an object
orbiting in a circular orbit of orbital period P . Determine the required
increase in length of the rigid cable L so that the satellite m2 catches
the object orbiting in the same orbital plane as of the given dumbbell
satellite system. Also find the inertial position, velocity, and accelera-
tion of the m2 at the time of capture. Hint. At the time of capture,
the inertial position, velocity and acceleration of the satellite m2 should
be equal to that of the orbiting object.

2.11 A system comprises of three identical satellites located at the vertices


of a triangle and connected through three identical rigid cables (sides of
a triangle) of length L. The system undergoes inplane orbital motion
as well as rotational motion (attitude motion). The orbital motion is
assumed to be circular and described by R (distance of system center
of mass from the Earth center) and θ (with respect to a reference line).
The system is rotating about its center of mass with an angular velocity
of β̇ with respect to orbital motion. Assuming the mass of each satellite
to be m, determine the inertial position, velocity, and acceleration of
the satellite.

2.12 A dumbbell satellite system comprising of satellites m1 and m2 con-


nected through a rigid cable of length L is librating with β in-plane
2.5. SUMMARY 119

libration in a circular orbit around the Earth. The following data are
given:
R(orbital radius)=7378 km, m1  m2 , m2 =100 kg, L=1 km
Using a Matlab program, plot the inertial position, velocity and accel-
eration of √
the satellite m1 verses orbit for β=0 deg, 45 deg, and 90 deg.
Take β̇ = 3θ̇cosβ. Here θ̇ denotes system orbital angular velocity.
The system described above is to be applied to catch a satellite m3 =3
kg orbiting in a circular orbit of orbital radius 7379 km. Plot the inertial
position, velocity and acceleration of this satellite and find the required
increase in length of the cable of the dumbbell satellite system in order
to capture the satellite. What would be the libration angle β and its
rate β̇ at the time of capture. Show in the plot when the capture occurs.
Hint. At the time of capture, the inertial position, velocity, and ac-
celeration of the satellite m2 should be equal to that of the orbiting
satellite.
2.13 Derive the inertial acceleration of body m1 moving along the cable in
the system undergoing inplane libration β as shown in Fig. 2.28. The
cable connecting the bodies has a mass of ρ per unit length. If the mass
m2 is much larger than m1 (i.e., m1  m2 ), then what would be the
inertial acceleration of m1 .
State at which values of inplane angle β the inertial acceleration of m1
would be maximum and minimum.

Orbit m2
β
L
S
R
Y m1
x

θ
E
X

Z
Figure 2.28: Space elevator.
120 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

2.14 The system shown in Fig. 2.29 is comprised of a rigid-body spacecraft


m1 and an auxiliary mass m2 (a point mass) orbiting the Earth. The
auxiliary mass m2 is attached to the spacecraft m1 by a rigid cable of
length L at an offset of ~a = aî + bĵ from the center of mass of the
spacecraft, where î and ĵ are unit vectors in the spacecraft body-fixed
frame Sxyz . The auxiliary mass m2 undergoes an in-plane libration or
oscillation of β about the local vertical. The mass of the spacecraft
is very large in comparison to the auxiliary mass m2 (i.e., m1  m2 ).
Assume the cable has a mass of ρ per unit length. Derive system linear
and angular momentums.

m2

Orbit L

β
a
S m1
Y
R

θ
E
X

Figure 2.29: System undergoing in-plane librational motion.

2.15 Derive the potential energy of a dumbbell system in Problem 2.5 un-
dergoing a three-dimensional libration motion with in-plane libration
β about the local vertical followed by the out-of-plane libration of η.
Assume the cable connecting the two bodies are massless.
State at which values of inplane angle β the potential energy would be
maximum and minimum assuming η = 0.

2.16 For the given system described in Problem 2.5 derive the kinetic and
potential energies of the system.

2.17 A system is comprised of a rigid-body platform m1 , trolley m2 , and the


payload m3 (Fig. 2.30). The payload m3 is deployed using a cable and
moves as defined by ~x. The trolley moves on the platform defined by
d~ = aî + bĵ, where î and ĵ are platform-fixed coordinate axes.
2.5. SUMMARY 121

Derive the kinetic and potential energies of the given system undergo-
ing in-plane motion characterised by the platform rotation, α and the
~ The cable and pulleys are
payload in-plane swing, β with respect to L.
assumed to be massless and no swing or rotation of the trolley about
its center of mass is considered. As the platform is much larger than
the masses of other bodies, the system center of mass lies on the center
of mass of the platform.

Orbit
11111111
00000000
Platform
00000000
11111111
00000000
11111111
00000000
11111111
j
x
00000000
11111111 m3
00000000
11111111
00000000
11111111
000
111
d 111
000
000
111
000
111
00000000
11111111
S
R11111111
000
111
000
111
L
00000000
00000000
11111111
000
111
000
111
m2
Y
00000000
11111111
00000000
11111111m1
i

θ
E
X

Figure 2.30: Platform with trolley and payload.

2.18 Determine the rotation matrix R132 (α, φ, γ) and angular velocity ω ~ of
the spacecraft (in body-fixed frame) for a 1-3-2 Euler angle rotation
sequence. State the singularity problem for this sequence. Using the
rotation matrix R132 (α, φ, γ) find the rotation matrix for a 2-3-1 Euler
angle rotation sequence.
2.19 Consider the sequence of body rotations that generates the orientation
of B with respect to A: γ about the y-axis, then φ about the new z-axis,
and finally, α about the new x-axis
(a) Determine the rotation Matrix for each simple rotation.
(b) Determine the complete rotation matrixes R231 (γ, φ, α) and R132 (α, φ, γ).
(c) Show that if α, φ and γ are very small then the order of rotations
does not matter.
2.20 Answer the following:

(a) What is the minimum number of quantities needed to specify the


orientation of one coordinate system with respect to another?
122 CHAPTER 2. KINEMATICS, MOMENTUM AND ENERGY

(b) According to Euler’s equations of rigid body motion, is it possible


have an object spinning around one of the axes, say b̂1 axis at
constant rate ω0 , with no spin around the other two axes (i.e., ω1 =
ω0 , ω2 = ω3 = 0) if no torque is applied? Explain your answer.
(c) Can you prove that no matter what sequence is taken for the
Euler angle rotations, the angle of the second rotation displays
singularity at either zero or ±90 deg? To avoid the singularity
problem, what are the methods that can be applied ?.
(d) Is it possible to have the first and the last rotations about the
same body axes in the Euler angle rotations?

2.21 A rigid satellite orbiting about the Earth is undergoing three-dimensional


attitude motion. Derive the kinetic and potential energies of the satel-
lites considering 1-3-2 Euler angle rotation sequence.

2.22 In Problem 2.17 if the cable has a mass of ρ per unit length, determine
kinetic and potential energies of the system.
2.23 For the given problem in Problem 2.3, if the dumbbell satellite system
undergoes a three-dimensional libration motion with in-plane libration
β about the local vertical followed by the out-of-plane libration of η,
determine the kinetic and potential energies of the system. Assume the
cable has a mass of ρ per unit length.
2.24 Determine the kinetic energy of the system in Problem 2.3 and potential
energy due to a flexible cable with fundamental mode of longitudinal
oscillation during in-plane motion of the system. Assume the system
center of mass does not lie on m1 or m2 . Consider the shape function
(for fundamental mode)
x
φ(x) =
L0
where L0 denotes undeformed length of the cable.
Chapter 3

Forces and Torques

After understanding the kinematics of space systems as discussed in the pre-


vious chapter, the next logical topic is forces and torques affecting the motion
of space systems. This chapter deals with these forces and torques with a fo-
cus on environmental disturbances due to gravity, aerodynamic forces, solar
radiation, and magnetic fields. The gravitational force is explained first, fol-
lowed by discussion on aerodynamic force. Next, the solar radiation pressure
is explained. The last section is on magnetic torques.

3.1 Introduction
A satellite may experience internal and external disturbances while in or-
bit. Internal disturbances are mainly due to uncertainty in center of gravity,
thruster misalignment, mismatch of thruster outputs, dynamics of flexible
bodies, and thermal shocks on flexible appendages. External disturbances
are due to environmental affects including gravity, aerodynamic forces, solar
radiation, magnetic fields, and free molecular reaction. Solar pressure and
gravitational torques are of the same order of magnitude for geostationary
satellites (altitude=36,000 km) while aerodynamic torques are dominant for
near-Earth satellites. Under the influence of these torques, the satellite de-
viates in time from its preferred position and orientation, leading to mission
failure. However, to precisely determine the effects of disturbances on satel-
lite dynamics and thereby, apply control forces and torques (using onboard
thrusters and momentum (reaction) wheels) to restore the satellite position
and orientation, we must model these disturbances mathematically. The ac-
curacies of these mathematical models in comparison with the real models
(based on onboard satellite data) leads to a better understanding of the dy-
namics of space systems and allows for more accurate control.
124 CHAPTER 3. FORCES AND TORQUES

In this chapter, we focus primarily on environmental disturbances due


to gravity, aerodynamic forces, solar radiation, and magnetic fields. The
mathematical models of forces or torques experienced by the spacecraft due
to these disturbances are presented. We begin with gravitational force.

Figure 3.1: Environmental torques on a geostationary(GEOS)-A satellite [8].

3.2 Gravitational Force

In this section, we consider two gravitational force models: simplified gravita-


tional force model and generalized gravitational force model. The first model
is based on Newton’s law of gravitation while the second model is based on
potential energy using longitude and latitude of the spacecraft and Legendre
polynomials.
3.2. GRAVITATIONAL FORCE 125

3.2.1 Simplified Gravitational Force Model

The simplified gravitational force model is based on Newton’s law of gravita-


tion that states that every pair of bodies in the Universe attract each other
with a force aligned along the line joining their mass centers. This force is
directly proportional to the product of their masses and inversely propor-
tional to the square of the distance between them. Considering two bodies
of masses m1 and m2 separated by a distance of r (Fig. 3.2), the force of
gravitation between them can be expressed mathematically as

Gm1 m2 ~r
F~ = − (3.1)
r2 r

where G is the proportionality constant called the gravitational constant or


Universal gravitational constant and its recommended value as per the 2002
CODATA (Committee on Data for Science and Technology) is 6.6742×10−11
m3 kg−1 s−2 . The (-) sign in Eq. (3.1) signifies the attractive force between
the bodies.

r
m2
F
−F
m1

Figure 3.2: Gravitational force of attraction between two bodies.

Remarks. 1. This law strictly applies to point masses or spherically-


symmetric mass bodies.
2. This law is sufficiently accurate for most purposes that deal with weak
gravitational fields (for example, motion of planets in the solar system or
motion of spacecraft around planets). Therefore, it is also called the Universal
Law of Gravitation.
3. This law is unable to predict the anomalous behavior of the perihe-
lion of Mercury. However, the theory of general relativity by Albert Einstein
explains this behavior more accurately and therefore, for such special cases,
Newton’s law of gravitation has been replaced by the theory of general rela-
tivity.
Based on Eq. (3.1), the gravitational force acting on mass dm of a space-
craft orbiting a planet M is

~
GM dm R
dF~ = − 2
(3.2)
R R
126 CHAPTER 3. FORCES AND TORQUES

the resulting torque on the satellite is


Z m
T~ = ~ × dF~
ρ (3.3)
0

where ρ
~ is the distance of mass dm from the satellite center of mass.

3.2.2 Generalized Gravitational Force Model


The gravitational force can be derived from the system potential energy.
Knowing the fact that the force is gradient of the system potential energy U ,
the gravitational force is given by

F~ = −∇U (3.4)

where ∇ denotes the gradient of a function f and is defined mathematically


as
∂f ∂f ∂f
∇f (x, y, z) = î + ĵ + k̂ (3.5)
∂x ∂y ∂z

where x, y, z are the components along unit vectors î, ĵ, and k̂, respectively.
These unit vectors represent the axes of a right-handed orthogonal coordinate
frame. Note the variable f is the function of x, y, and z, only.
The gradient of the potential energy U can be written as
∂U ∂U ∂U
∇U = î + ĵ + k̂ (3.6)
∂x ∂y ∂z

Here î, ĵ, and k̂ are the unit vectors of the body-fixed coordinate frame
S − xyz.
A general expression of the potential energy of a spacecraft (m) orbiting
a planet (M ) due to gravity is
( ∞
"  n
GM m X Re
U =− 1− Cn0 Pn0 (sinδ)
r n=2
r
n  n #)
X Re
+ (Cnm cosmλ + Snm sinmλ)Pnm (sinδ)
m=1
r
(3.7)

where r is the radial distance of the spacecraft from the center of mass of
the planet and Re defines equatorial radius or characteristic length of the
planet. The nomenclature λ and δ describe the longitude and latitude of the
spacecraft measured in the planet-fixed frame. G denotes the universal grav-
itational constant. The terms Pnm (sinδ)cosmλ and Pnm (sinδ)sinmλ are
3.2. GRAVITATIONAL FORCE 127

called tesseral harmonics of nth degree and mth order and the correspond-
ing Cnm and Snm are known as tesseral harmonic coefficients. The tesseral
harmonics coefficients for which n = m 6= 0, are called sectorial harmonic
harmonic coefficients, while tesseral harmonics of order zero, Pn0 = Pn , are
zonal harmonics and the corresponding Cn0 are known as zonal harmonic
coefficients of order 0, which are also denoted by Jn . The coefficients Cn0
specify oblateness of the planet while Cnm characterize the ellipticity of the
planet’s equator. The terms Pn0 (sinδ) denote Legendre polynomial of degree
n and order 0 given by
1
P2 (sinδ) = (3sin2 δ − 1)
2
1
P3 (sinδ) = (5sin3 δ − 3sinδ) (3.8)
2
1
P4 (sinδ) = (35sin4 δ − 30sin2 δ + 3)
8
············

The terms Pnm are associated Legendre polynomial of degree n and order m
given as
k
(1 − sin2 δ)m/2 X (−1)i (2n − 2i)!(sinδ)n−m−2i
Pnm (sinδ) = (3.9)
2n i=0
i!(n − m − 2i)!(n − i)!

Here the integer k is



 n−m n − m even
k= 2 (3.10)
 n−m−1 n − m odd
2

For a spacecraft orbiting an asteriod, the potential energy of the spacecraft


simplifies to
"  2  3
GM m Re Re
U =− 1 − C20 P2 (sinδ) − C30 P3 (sinδ)
r r r
 2 #
Re
−3 C22 cos2 δcos(2λ) (3.11)
r

Note that Snm , C11 , C33 and other coefficients are zero.
In the case of a spacecraft orbiting the Earth (assuming the Earth as
symmetrically mass bodies), the potential energy of the spacecraft is

GM m
U =− (3.12)
r
128 CHAPTER 3. FORCES AND TORQUES

Note that the coefficients Snm and Cnm are zero.


Example 3.1
Derive an expression of gravity force acting on a spacecraft (mass m)
orbiting the Earth (mass M ) with an orbital distance of r. Assume the
spacecraft and Earth are symmetrically mass distributed bodies.
Solution.
Applying Newton’s Law of gravitation, the gravity force acting on a space-
craft (mass m) orbiting the Earth (mass M ) is
µm
F~ = − 2 îr (3.13)
r
where µ = GM is a gravitational parameter and îr is an unit vector along ~r,
directed away from the Earth.
Alternatively, the gravity force can be derived from the potential energy
of the spacecraft that is expressed as
µm
U =− (3.14)
r
Thus, the gravitation force is
∂U µm
F~ = −∇U = − îr = − 2 îr (3.15)
∂r r

3.3 Aerodynamic Forces


The aerodynamic forces or torques experienced by satellites vary with their
altitudes. Based on altitude, the Earth’s atmosphere is divided into four lay-
ers: Troposphere, Stratosphere, Mesosphere, and Thermosphere. Variations
of temperature and pressure along these layers of atmosphere are shown in
Figs. 3.3 and 3.4. Aerodynamic forces due to the motion of atmosphere
around satellites must be mathematically modelled. Mathematical models
that accurately match real data and are simple to model result in more accu-
rate predictions of satellite motion. Here we discuss a simplified aerodynamic
force model and a more sophisticated free-molecular force model.

3.3.1 Simplified Aerodynamic Force Model


The aerodynamic force on the satellite surface is given by
1
F~L = − CL ρAVrel
2
n̂ (3.16)
2
1
F~D = − CD ρAVrel
2
V̂rel (3.17)
2
3.3. AERODYNAMIC FORCES 129

Figure 3.3: Earth’s atmosphere showing variation of temperature.(Courtesy


of National Weather Service, Texas)

where CL and CD are the lift and drag coefficients, ρ is the air density, and
~rel or V̂rel (i.e., V̂rel =V
A is the area of the satellite surface normal to V ~rel /Vrel ).
~rel is the velocity vector of the satellite surface with respect to the atmo-
V
sphere while n̂ is the unit vector normal to V ~rel (Note V~rel × n̂ is normal
outward to the plane V̂ − n̂). The aerodynamic force acts at the center of
pressure of the satellite surface under consideration. Note the drag force F~D
on a body acts in the direction opposite to the relative velocity V ~rel .
Assuming the atmosphere is stationary (rotating with the same angular
velocity as the Earth), the V~rel can be expressed as

V~rel = ~vs − ~ve = ~vs − ω ~


~e × R (3.18)

where ~vs is the orbital velocity vector of the satellite.


The density of atmosphere ρ at the height h assuming an exponential
atmospheric model is given by


h − h0 −g
h − h0
ρ = ρ0 e H = ρ0 e 0 RT (3.19)
130 CHAPTER 3. FORCES AND TORQUES

Figure 3.4: Earth’s atmosphere showing variation of pressure and tempera-


ture.

where ρ0 is the density of the atmosphere at reference height h0 , g0 is the


gravitational acceleration at sea level (g0 =9.8 m/s2 ), R is the gas constant
(R=287 N-m/kg-K (1,715 ft-lb/slug-R)), T is the air temperature at height
h in degrees Kelvin, and H is the scale height. These parameters are taken
as ρ0 =1.585× 10−12 , H=62.2 km, and h0 =450 km. The atmospheric density
ρ, temperature, and pressure vary with altitude and these data are listed in
the Appendix along with scale height H.
The coefficients CL and CD are dependent on the geometry of the satellite
and vary with altitude, Mach number, and angle of attack as

CL ≡ CL (M, ζ, Vrel ), CD ≡ CD (M, ζ, Vrel ) (3.20)

where M is Mach number defined as the ratio of the speed of the spacecraft,
3.3. AERODYNAMIC FORCES 131

vs to the speed of sound in the surrounding medium, c, given by


vs
M= (3.21)
c
Here c is the acoustic velocity and is determined using the relation

c = kRT (3.22)
where k is the specific heat ratio (k=1.40 for air), R is the gas constant, and
T is the air temperature in degrees Kelvin. The coefficients CL and CD are
generally determined by wind tunnel experiments. For many common rocket
shapes, CD at M = 0.1 lies in the range of 0.35 to 1.0 with the average of
approximately 0.75. Earth orbiting satellites typically have CD varying from
2 to 4.

3.3.2 Free-Molecular Aerodynamic Force Model


Based on the free-molecular aerodynamic force model [1], the force on a flat
surface of area A (Fig. 3.5) is
h  τ i
F~ = A −n̂p + (n̂sinα − V̂rel ) (3.23)
cosα
~rel /Vrel is the
where p is the total pressure, τ is the shearing stress, V̂rel = V
velocity vector of the surface element with respect to the atmosphere, and n̂
is the outward-pointing unit normal vector. The angle of attack α is given
by
sinα = n̂T V̂rel (3.24)

The total pressure p and shearing stress τ are given by


(  r )
p 2 − σn σn Ts
= √ sinα +
q∞ π 2s Ta
1 −s2 sin2 α √
 
× e + π[1 + erf (ssinα)]sinα
s
 
2 − σn
+ [1 + erf (ssinα)] (3.25)
2s2
 
τ 1 2 2
= σt √ e−s sin α + [1 + erf (ssinα)]sinα (3.26)
q∞ cosα s π
where σn and σt are normal and tangential accommodation coefficients, Ts
is the absolute temperature of spacecraft surface, Ta is the atmospheric tem-
perature, q∞ is the dynamic pressure given by
1 2
q∞ = ρV (3.27)
2 rel
132 CHAPTER 3. FORCES AND TORQUES

Surface
Element of
Spacecraft
111
000
000
111
α111
000
V
000
111
000
111
000
111
000
111
000
111 p
n
000
111
000
111
000
111
τ
000
111
Outer 111
000
Surface
Figure 3.5: Geometry of free-molecular aerodynamic force model.

erf () is the error function defined by


Z x
2 2
erf (x) = √ e−y dy (3.28)
π 0

and s is the air speed, nondimensionalized by the mean molecular speed


of the atmosphere
s
2
Ma Vrel
s= (3.29)
2R Ta

Here Ma is the mean molar mass of the atmosphere and R∗ is the uni-
versal gas constant. Note that Ma and Ta vary with altitude and can be
obtained from the standard atmospheric model (1976 U.S. Standard Atmo-
sphere [2]). Note 1976 U.S. Standard Atmosphere model does not include
diurnal perturbations.
The accommodation coefficients σn and σt , represent the fact that some
impinging air molecules come to thermal equilibrium with the spacecraft and
then get remitted thermally while others reflect specularly. Here the specu-
lar reflection means that the angle of incidence equals the angle of reflection
and that the relative speed remains unchanged. σn =σt =0 implies specular
reflection of all the impinging air molecules, while σn =σt =1 implies full ther-
mal accommodation of all the impinging air molecules. Typical values of
σn and σt lie between 0.85 to 1. For example, altitude H=400 km, surface
temperature Ts =300oK, σn =0.85, and σt =0.9. In another example [3], alti-
tude H=500 km, mean Ta =997.3oK, mean density ρ=6.967× 10−13 kg/m3 ,
molecular speed s=1, σn =0.5, and σt =0.5.
3.4. SOLAR RADIATION PRESSURE FORCE 133

Note that if the outer surface of a spacecraft is nonconvex, the effects


shading and reincidence of emitted air molecules must be considered in the
model. Here shading means that one aerodynamic surface lies directly up-
stream of another aerodynamic surface.

3.4 Solar Radiation Pressure Force


Solar radiation pressure (SRP) forces are due to photons impinging on the
satellite surface. Modelling of these forces differ depending upon the satellite
surface and its configuration. Some configurations proposed for solar sails
are shown in Fig. 3.6. Here we discuss two SRP force models.

Figure 3.6: Solar sail configurations [11].

3.4.1 Solar Radiation Pressure Force Model I


The force due to the SRP on a flat surface (3.7) is given by
   
2
F~ = pAH(cosζ)(ŝ · n̂) (1 − ρs − ρτ )ŝ + 2ρs (ŝ · n̂) + ρd n̂ (3.30)
3
134 CHAPTER 3. FORCES AND TORQUES

where
ŝ=unit vector of the incoming light from the sun on the plate;
n̂= unit vector along the surface normal;
p=nominal solar radiation pressure; p = p0 /r2 ; p0 = 1.04 × 1017 Newton;
r is the distance from the Sun. For a satellite orbiting the
Earth (1 astronomical unit from the Sun), p is nearly constant
and its value is 4.563×10−6 N/m2 ;
A=surface area exposed to impinging photons;
H(cosζ)=1 for cosζ ≥ 0, 0 for cos ζ < 0 (assuming front surface is illuminated);
H(cosζ)=1 (assuming front and back surfaces are illuminated);
ρs =a fraction of impinging photons specularly reflected;
ρτ =a fraction of impinging photons transmitted;
ρa =a fraction of impinging photons absorbed;
ρd =a fraction of impinging photons diffusely reflected;
ρa + ρs + ρτ + ρd = 1.

Surface
Element of
Spacecraft
Incoming 01
Photons 1010
1010
ζ 1010
ζ 1010 ξ n
10
1010 F
Specularly 10 s
Absorbed
Reflected Front Photons
Photons Surface t
Figure 3.7: Solar radiation pressure force model of a nonperfect flat surface.

The orbital motion of satellite along with the direction of the Sun is
3.4. SOLAR RADIATION PRESSURE FORCE 135

illustrated in Fig. 3.8. Referring to this figure we can express the unit
vector of the incoming light from the sun, ŝ with respect to Iˆn − Jˆn − K̂n
coordinate frame (where Iˆn in the direction towards the ascending node; K̂n
is perpendicular to the orbit plane along ~h, and K̂n × Iˆn =Jˆn ) as
ŝ = −cos(ψ − Ω)Iˆn − sin(ψ − Ω)cos(i − )Jˆn + sin(ψ − Ω)sin(i − )K̂n
(3.31)
where ψ is the Sun angle with respect to the Vernal equinox, and  is the
angle between the equatorial and the ecliptic plane (=± 23 deg 27 min).

O r b it p la n e

S a t e llit e

N¢ R
T o Sun

E a rth C e n te r q y E c lip t ic
p la n e
i e s

L in e o f N o d e s
N
E q u a t o r ia l
p la n e

Figure 3.8: Geometry of satellite motion under solar radiation pressure.

For an ideal case of a perfect mirror with ρs =1 (i.e., ρd =ρa =ρτ =0) and
front and back surfaces illuminated, the preceding Eq. (3.30) for SRP force
reduces to
F~ = 2pAcos2 ξn̂ (3.32)

For an ideal case of a black body with ρa =1 (i.e., ρs =ρd =ρτ =0) and
front and back surfaces illuminated, the preceding Eq. (3.30) for SRP force
simplifies to
F~ = p(Acosξ)~s (3.33)
where Acosξ is called the projected area of the surface under consideration.
Alternatively, the SRP force can be approximately expressed as
F = ηpAcos2 ζ (3.34)
136 CHAPTER 3. FORCES AND TORQUES

where η is the overall sail thrust coefficient varying from 0 to 2. For a real
solar sail with sail wrinkles and billowing, η ≈ 1.8.
For satellites undergoing small pitch librations, the SRP force per unit
mass is modelled as
f~ = p(1 + ρ)(A/m)ŝ (3.35)
where ρ is the overall surface reflectance (0 for a black body and 1 for a mirror)
and A/m is the sail area-to-mass ratio. The magnitude of the acceleration
in ms−2 can be approximated as
−4.5 × 10−8 A
f= (3.36)
m
where A is the cross-sectional area of the satellite exposed to the Sun and
m is the mass of the satellite in kilograms. For satellites above 800 km,
acceleration from SRP is greater than that from aerodynamic drag while
below 800 km altitude, acceleration from atmospheric drag is greater.
Example 3.2
Derive the SRP force on a solar sail orbiting the Sun (Fig. 3.9) with re-
spect to local vertical and local horizontal (LVLH) coordinate frame. Assume
ρd ≈ 0 and both sides of the solar sail are illuminated.

j n
ζ i
s
Sail

r
Incident Orbit
Radiation

θ Inertial Reference
Sun
Figure 3.9: Satellite under solar-radiation pressure.

Solution.
The SRP force due to the solar radiation pressure (SRP) on a flat surface
is
   
2
F~ = pAH(cosζ)(ŝ · n̂) (1 − ρs − ρτ )ŝ + 2ρs (ŝ · n̂) + ρd n̂ (3.37)
3
3.4. SOLAR RADIATION PRESSURE FORCE 137

Here n̂ can be written in terms its components along î and ĵ as

n̂ = cosζ î + sinζ ĵ (3.38)

Substituting n̂ in Eq. (3.37) and assuming ρd = 0 and both sides of the


sail can be illuminated, we have
p0 h i
F~ = 2 A|cosζ| (1 + ρs cos2ζ − ρτ )î + (ρs sin2ζ)ĵ (3.39)
r

3.4.2 Solar Radiation Pressure Force Model II


This model includes optical and thermal properties of the surface of the
satellite [Chapter 2 of Ref. [6] and Appendix A of Ref. [5]]. The SRP force
acting on a flat surface is given by

F~ =Fn î + Ft ĵ (3.40)
Ft
q
(in magnitude and direction) F = Fn2 + Ft2 , tanξ = (3.41)
Fn
where
Fn ef Bf − eb Bb
=(1 + ρr ρs )cos2 ζ + Bf ρr (1 − ρs )cosζ + (1 − ρr )cosζ
pA ef + eb
(3.42)
Ft
=(1 − ρr ρs )cosζsinζ (3.43)
pA

Here Bf and Bb are non-Lambertian coefficients for front and back surfaces,
ef and eb are front and back surface emission coefficients, ρr is the reflectivity
of front surface, ρs is the specular reflection coefficient, and ξ is the angle of
SRP force vector from surface normal. For a square solar sail (Fig. 3.6), the
values of the optical properties (Appendix A of [5]) are: Bf =0.79, Bb =0.55,
ef =0.05, eb =0.55, ρr =0.88, ρs =0.94.
Note that the surface of a solar sail is curved and as a result the pressure
distribution (p) is not uniform across the surface. A numerical integration
of Eqs. (3.42) and (3.43) is carried to determine the pressure distribution.
This procedure is an iterative process because the pressure distribution is a
function of the shape of the solar sail, and vice versa. For the square solar sail
(Fig. 3.6), a parameterized SRP force model based on the iterative process
(Appendix B of [5]) is

F = ηpA(0.349 + 0.662cos2γ − 0.011cos4γ) (3.44)

where η=1.816 and γ=ζ − ξ.


138 CHAPTER 3. FORCES AND TORQUES

3.4.3 Earth shadow


The SRP does not fall on a satellite all the time while it orbits the Earth.
In some part of the orbit called eclipse, the SRP is blocked by the Earth
which comes in between the Sun and the satellite, and therefore, the SRP
force on the satellite is null during eclipse. The Models I and II discussed in
preceding sections should take this effect into consideration. There are two
approaches to handle Earth shadow. The first approach considers the Earth
as a perfect sphere and the Sun is assumed to be at an infinite distance from
the Earth. The second assumption implies that all rays coming from the Sun
are parallel to each other. In the second approach, the actual size of the Sun
is considered and solar rays from different parts of the Sun are not assumed
to be parallel to each other. The fist approach is simpler and is described
here (see Fig. 3.10).

Satellite
Sun Rays
Orbit
R

χ
χ R
s e
s Shadow
Apogee Perigee

Earth

Figure 3.10: Earth shadow.

Based on the first approach, the conditions for the satellite to be in the
Earth’s shadow are
cosχ < 0 and Re − R(1 − cos2 χ)1/2 > 0 (3.45)
where
a(1 − e2 )
cosχ = cosψcosθ + sinψcos(i − s )sinθ, R=
1 + ecosθ
Here a and e are orbit semi-major axis and eccentricity.
The size of the eclipse region (χs ) for the satellite in a circular orbit is
derived as
π − χs
Rcos( ) = Re (3.46)
2
3.5. MAGNETIC FIELD TORQUE 139
   
π Re Re
⇒ χs = − 2cos−1 or χs = 2sin−1 (3.47)
2 R R
For example, a satellite in the low-earth orbit, say 600 km altitude (R=6978.14
km), remains in eclipse for 132.1 deg out of 360 deg or 37% of the orbital
period. For a satellite in geostationary orbit (R=442241km), it remains in
eclipse for 5% of the orbital period.

3.5 Magnetic Field Torque


The Earth is a giant magnet and the interaction of its magnetic field with
a spacecraft causes a torque on the spacecraft. This magnetic torque is
~ and the magnetic dipole
dependent on the Earth’s magnetic field strength (B)
~ ), and can
moment of the spacecraft or the magnetic rod of the spacecraft (M
be mathematically expressed as
T~m = M
~ ×B
~ (3.48)
Typically, a spacecraft has a magnetic dipole moment of 1 amp-turn-m2 (A-
m2 ). For a magnetic rod, the magnetic dipole moment is given by
~ = N IAn̂
M (3.49)
where I is the current in Amperes, N is the number of coil turns, A is the
coil cross-sectional area, and n̂ is the unit vector normal to the coil area.
The magnetic field of the Earth, B, ~ is distributed around the planet as
shown in Figs. 3.11 and 3.12. The strength of the field is approximately
30000 nT at the equator and 60000 nT at the poles on the surface of the
Earth [9]. The magnetic dipole axis, denoted by m̂ or îm (Fig. 3.12) is
currently inclined to the equatorial plane by 11.5o and is drifting westward
at about 0.2o per year, and the strength is decreasing by 0.05% per year.
Note that the dipole is located at 79.8o N latitude and 107.0oW longitude
(as per year 1999) and this location is near the Ellef Rignes Island in Canada,
which is approximately 700 miles from the geographic North Pole.
The Earth’s magnetic field B ~ can be expressed as the negative gradient
of a scalar potential function V ,
~ = −∇V
B (3.50)
where V is described by a spherical harmonic expansion, a series of special
spherical functions of latitude/longitude and their associated coefficients 1 :
k  n+1 X n
X Re
V (R, θ, φ) = Re (gnm cosmφ + hm m
n sinmφ)Pn (θ) (3.51)
n=1
R m=0
1 This series was proposed by the German mathematician and magnetician Frederick

Gauss in 1838.
140 CHAPTER 3. FORCES AND TORQUES

Figure 3.11: The Magnetosphere.

Figure 3.12: Earth’s Magnetic Field.

where Re is the equatorial radius of the Earth; gnm and hm n are Gaussian coef-
ficients; R, θ, and φ are the geocentric distance, coelevation (or latitude), and
East longitude from Greenwich, and Pnm is the associated Legendre function
of degree n and order m, that resembles distorted sine waves. The Gaussian
3.5. MAGNETIC FIELD TORQUE 141

Table 3.1: Magnetic Field Models depending Upon Number of First


Terms in Eq. 3.51
.

Number of Magnetic Remarks


First Terms Field Model
One Centered dipole model coincident with
the Earth’s spin-axis
and the least accurate
Three Tilted dipole model Most common
Eight Quadrapole field model
Ten Octopole field model

coefficients gnm and hm n are determined through a least-squares analysis of a


world-wide distribution of magnetic observations. Any shape of the actual
magnetic field can be approximated to the desired resolution and accuracy by
selecting appropriate number of terms, k in the series Eq. 3.51. For example,
the most commonly used model, International Geomagnetic Reference Field
(IGRF) model2 uses k=10. The magnetic field originating within the core of
the Earth requires k = 15 while crustal anomalies visible in magnetic data
at satellite altitudes requires k=80.
The B ~ field in tangential coordinates is derived as follows

∂V 1 ∂V 1 ∂V
BR = − , Bθ = − , Bφ = − (3.52)
∂R R ∂θ Rsinθ ∂φ

Several models of the Earth’s magnetic field listed in Table 3.1 can be
derived from the spherical harmonic model described by previous Eqs. (3.50)
and (3.51), depending upon the number of first terms assumed in the series
Eq. 3.51. The titled dipole model, explained next, is commonly used.
A tilted dipole model of the Earth’s magnetic field can be derived by
considering only the first degree (n = 1) and all orders (m = 0, 1) in Eq.

2 The latest data on Gaussian coefficients for IGRF is listed in Appendix. These coeffi-

cients are updated every five years using new data, and named a Definitive Geomagnetic
Reference Field (DGRF). The DGRF data is also listed in Appendix.
142 CHAPTER 3. FORCES AND TORQUES

(3.51). The scaler potential V becomes


Re3  0 0
g P (θ) + (g11 cosφ + h11 sinφ)P11 (θ)

V (R, θ, φ) =
R2 1 1
1
= 2 g10 Re3 cosθ + g11 Re3 cosφsinθ + h11 Re3 sinφsinθ

(3.53)
R
where the first degree Gaussian coefficients (as per year 2000) are

g10 = −29615nT, g11 = −1728nT, and h11 = 5186nT (3.54)


The total dipole strength is
1/2
Re3 H0 = Re3 (g10 )2 + (g11 )2 + (h11 )2

(3.55)

which leads to a value of H0 =30115 nT.


The coelevation and the East longitude of the dipole (as per year 2000)
are
 
g10
θm = cos−1 H0 = 196.540 (3.56)
 
h11
φm = tan−1 g11
= 108.430 (3.57)

The magnetic field in local tangential coordinates with a tilted dipole


model is obtained as
 2
∂V Re  o
g1 cosθ + (g11 cosφ + h11 sinφ)sinθ

BR = − =2 (3.58)
∂R R
 2
1 ∂V Re  o
g1 sinθ − (g11 cosφ + h11 sinφ)cosθ

Bθ = − = (3.59)
R ∂θ R
 2
1 ∂V Re  1
g1 sinφ − h11 cosφ

Bφ = = − = (3.60)
Rsinθ ∂φ R

Assuming the Earth’s magnetic field a vector dipole with strength and
pole direction as given by Eqs. (3.56)-(3.60), we can describe the magnetic
~ in vector form as
field B
~ µf
B(R) = 3 [3(îm · îR )îR − îm ] (3.61)
R
where µf is the Earth’s magnetic dipole strength, îR is the unit vector joining
the Earth’s dipole center to the center of mass of the spacecraft, and îm is the
unit vector in the direction of the Earth’s magnetic field towards magnetic
south pole.
The Earth’s magnetic dipole strength µf is given by
µf = µ0 µm or Re3 H0 (3.62)
3.5. MAGNETIC FIELD TORQUE 143

Here µ0 is the permeance of vacuum or permeability constant ( 4π× 10−7


Weber/(Amp-m)), µm is the Earth’s magnetic dipole moment (8.06× 1022
Amp-m2 ), and Re is the equatorial radius of the Earth. The value of µf can
be taken as 8.1×1015 Wb-m or tesla-m3 (as in year 1962) or 7.96×1015 Wb-m
or tesla-m3 (as in year 1975).
The dipole is centered in the Earth and its orientation is defined by the
longitude of αm with respect to the vernal equinox and the latitude of (90o -
θm ) with respect to the equatorial plane. Note θm is the coelevation of the
dipole. We can express ~im in terms of the Earth centered coordinate frame
E − XY Z with X-axis directed towards the vernal equinox, Z-axis along
the Earth’s spin axis, and Y -axis makes the right-handed tried. Using the
transformation Ry (90o −θm )Rz (αm ) from the frame E−XY Z to E−îm ĵm k̂m ,
the ~im is expressed as
  
 sinθm cosαm 

 
  Iˆ 
 

  
îm = sinθm sinαm Jˆ (3.63)

 
 
 

cosθm   K̂ 

   

where αm is
αm = αg0 + ωe t + φm (3.64)
Here αg0 is the right ascension of the Greenwich meridian at some reference
time called the Greenwich sidereal time, 98.8279 deg at 0 hour UT, December
31, 1979; ωe is the average rotation rate of the Earth, 360.9856469 deg/day
or 7.2921152×10−5 rad/sec; t is the time since reference; φm is the East
longitude of the dipole, 108.43 deg (year 2000); θm is the coelevation of the
dipole, 196.54 deg (year 2000).
Substituting îm from Eq. (3.63) into Eq. (3.61) and taking R̂ = Rx Iˆ +
Ry Jˆ+ Rz K̂ (where Rx , Ry , and Rz are the geocentric direction cosines of R̂),
the Earth’s magnetic field with respect to the geocentric inertial coordinate
frame becomes
  

 3(î m · R̂)R x − sinθ m cosα 
m 
 
 Iˆ 

3
 
~ R e H 0
  
B= 3( î m · R̂)R y − sinθ m cosαm Jˆ (3.65)
R3   
  
 
3(îm · R̂)Rz − cosθm K̂

  
 

Note the preceding Eq. (3.65) assumes that the magnetic dipole direction
îm is towards magnetic south pole. For the orbital reference frame xo yo zo
considered as
“the xo -axis points along the spacecraft velocity vector, the yo -axis is
taken along normal to the orbital plane, and the zo -axis represents the third
axis of this right handed frame taken pointing along the nadir”
144 CHAPTER 3. FORCES AND TORQUES

then îm = ĵo (unit vector along the yo -axis pointing down towards mag-
netic south pole).
For the orbital reference frame xo yo zo taken as
“the xo -axis along the local vertical pointing towards zenith, the zo -axis is
taken along normal to the orbital plane, and the yo -axis represents the third
axis of this right handed frame taken pointing along the spacecraft velocity
vector”
then îm = −k̂o (unit vector along the negative zo -axis pointing down
towards magnetic south pole). Considering this reference frame and assuming
~ with respect to the orbital
no Earth rotation and no orbit precession, the B
reference frame can be expressed as
  


 cosi m





 î 
o 

~ µ0 µm    
B= −2sin(ω + θ)sini m ĵ o (3.66)
R3   
  
 
cos(ω + θ)sinim k̂o

  
 

where im is the orbit inclination of the satellite with respect to the magnetic
equator.
Let M ~ = Mx îo + My ĵo + Mz k̂o and B~ = Bx îo + By ĵo + Bz k̂o , then the
magnetic torque referring to Eq. 3.48 can be expressed as

T~ =M~ ×B
~
  
 My Bz − Mz By

 
  îo 
 
 




= Mz Bx − Mx Bz ĵo (3.67)

 
 
 

 B M −B M
 
 k̂ 
x y y x o

or

    
 T  0 Bz −By  Mx 
 x

 
 
   

Ty = −Bz 0 Bx  My (3.68)
 

 
   

 T
 
 By −Bx 0  M
 

z z

When a satellite is in a magnetic equatorial or nearly equatorial orbit,


i.e., im =0, as per Eq. (3.66), Bx = By =0, and B ~ will only have a component
along the k̂o -axis. So, as per Eq. (3.67), there exists no torque about k̂-axis
(pitch-axis). In other words, as B ~ is along the pitch axis, the pitch torque
is null. However, if a satellite is in a polar orbit (i.e., im =90 deg), Bz =0
3.6. SUMMARY 145

using Eq. (3.66). The Bx and By terms become zero if the satellite is at the
equator (ascending or descending node) and at the poles, respectively. That
is to say, at the equator, the roll torque is not available while at the poles, yaw
torque is not available. In fact, as the satellite moves from the equator to the
poles, the roll torque increases while the yaw torque decreases and reaches
a minimum at the poles. For satellite motion in a circular orbit, it can
be proven that the components of B ~ along the pitch axis remain constant.
Furthermore, it can be shown that Earth’s magnetic field strength for a
satellite in equatorial orbit (orbital radius R) is approximately B = µf /R3
while in polar orbit, the Earth’s magnetic field strength increases by a factor
of two (i.e., B = 2µf /R3 ).

3.6 Summary
In this chapter, environmental forces and torques (due to gravity, aerody-
namic, magnetic fields, and solar radiation) on space systems are discussed
and mathematical models for these forces and torques are presented.

References

1. Gombosi, T. I., Gaskinetic Theory, Cambridge Univ. Press, Cambridge,


England, U.K., 1994, pp. 227255.

2. United States Committee on Extension to the Standard Atmosphere,U.S.


Standard Atmosphere, 1976, National Oceanic and Atmospheric Ad-
ministration, Washington, DC, 1976, pp. 5097.

3. Wertz, J. R., editor, Spacecraft Attitude Determination and Control, D.


Reidel, Dordrecht, Holland, 1978.

4. Wiley J. Larson and James R. Wertz, editors, Space Mission Analysis


and De- sign, Microcosm, Inc., Torrance, CA, second edition, 1995.

5. Wright, J. L., Space Sailing, Gordon and Breach, New York, 1992.

6. McInnes, C. R., Solar Sailing: Technology, Dynamics and Mission Ap-


plications, Springer Praxis, 1999.
146 CHAPTER 3. FORCES AND TORQUES

7. Friedman, L., Star Sailing: Solar Sails and Interstellar Travel, Wiley,
New York, 1988.
8. Modi, V. J.,“On the Semi-passive Attitude Control and Propulsion of
Space Vehicles using Solar Radiation Pressure,” Acta Astronautica, Vol.
35, No. 2/3, 1995, pp. 231-246.
9. Geological Survey of Canada, National Geomagnetism Program, http://
www.geolab.nrcan.gc.ca/geomag.
10. Division V, Working Group 8. International Geomagnetic Reference
Field - 2000. International Association of Geomagnetism and Aeron-
omy (IAGA), 2000.
11. Price, H., Ayon, J., Buehler, M., Garner, C., Klose, G., Mettler, E.,
Nakazono, B., and Sprague, G., “Design for a Solar Sail Demonstra-
tion Mission,” Space Technology and Applications International Forum
(STAIF 2001), Albuquerque, NM, Feb. 2001.
3.7. APPENDIX 147

3.7 Appendix
3.7.1 Atmospheric Data
148 CHAPTER 3. FORCES AND TORQUES

Table 3.2: Physical Properties of Standard Atmosphere in SI Units

Altitude Temperature Pressure Density Viscosity


(meters) (degrees K) (Pa) (kg/m3 ) (N-s/m2 )
-5,000 320.7 1.778E+5 1.931 1.942E-5
-4,000 314.2 1.596E+5 1.770 1.912E-5
-3,000 307.7 1.430E+5 1.619 1.882E-5
-2,000 301.2 1.278E+5 1.478 1.852E-5
-1,000 294.7 1.139E+5 1.347 1.821E-5
0 288.2 1.013E+5 1.225 1.789E-5
1,000 281.7 8.988E+4 1.112 1.758E-5
2,000 275.2 7.950E+4 1.007 1.726E-5
3,000 268.7 7.012E+4 9.093E-1 1.694E-5
4,000 262.2 6.166E+4 8.194E-1 1.661E-5
5,000 255.7 5.405E+4 7.364E-1 1.628E-5
6,000 249.2 4.722E+4 6.601E-1 1.595E-5
7,000 242.7 4.111E+4 5.900E-1 1.561E-5
8,000 236.2 3.565E+4 5.258E-1 1.527E-5
9,000 229.7 3.080E+4 4.671E-1 1.493E-5
10,000 223.3 2.650E+4 4.135E-1 1.458E-5
15,000 216.7 1.211E+4 1.948E-1 1.422E-5
20,000 216.7 5.529E+3 8.891E-2 1.422E-5
30,000 226.5 1.197E+3 1.841E-2 1.475E-5
40,000 250.4 2.871E+2 3.996E-3 1.601E-5
50,000 270.7 7.978E+1 1.027E-3 1.704E-5
60,000 255.8 2.246E+1 3.059E-4 1.629E-5
70,000 219.7 5.520 8.754E-5 1.438E-5
80,000 180.7 1.037 1.999E-5 1.216E-5
90,000 180.7 1.644E-1 3.170E-6 1.216E-5
3.7. APPENDIX 149

Table 3.3: Physical Properties of Standard Atmosphere in U.S. Units

Altitude Temperature Pressure Density Viscosity


(feet) (degrees R) (psia) (slug/ft3 ) (lb-s/ft2 )
-15,000 572.2 24.626 3.610E-3 4.031E-7
-10,000 554.3 20.847 3.155E-3 3.935E-7
-5,000 536.5 17.554 2.745E-3 3.835E-7
0 518.7 14.696 2.377E-3 3.736E-7
5,000 500.8 12.054 2.048E-3 3.636E-7
10,000 483.0 10.108 1.756E-3 3.534E-7
15,000 465.2 8.297 1.496E-3 3.431E-7
20,000 447.4 6.759 1.267E-3 3.326E-7
25,000 429.6 5.461 1.066E-3 3.217E-7
30,000 411.8 4.373 8.907E-4 3.107E-7
35,000 394.1 3.468 7.382E-4 2.995E-7
40,000 390.0 2.730 5.873E-4 2.969E-7
45,000 390.0 2.149 4.623E-4 2.969E-7
50,000 390.0 1.692 3.639E-4 2.969E-7
55,000 390.0 1.332 2.865E-4 2.969E-7
60,000 390.0 1.049 2.256E-4 2.969E-7
65,000 390.0 0.826 1.777E-4 2.969E-7
70,000 392.2 0.651 1.392E-4 2.983E-7
75,000 395.0 0.514 1.091E-4 3.001E-7
80,000 397.7 0.406 8.571E-5 3.018E-7
85,000 400.4 0.322 6.743E-5 3.035E-7
90,000 403.1 0.255 5.315E-5 3.052E-7
95,000 405.8 0.203 4.196E-5 3.070E-7
100,000 408.6 0.162 3.318E-5 3.087E-7
150,000 479.1 0.020 3.456E-6 3.512E-7
200,000 457.0 0.003 5.270E-7 3.382E-7
250,000 351.8 0.000 7.034E-8 2.721E-7
300,000 332.9 0.000 4.625E-9 2.593E-7
150 CHAPTER 3. FORCES AND TORQUES

Table 3.4: Atmospheric Scale Height and Density to 2,000 km

Altitude Scale Height Atmospheric Density


(km) (km) Mean Maximum
3
(kg/m ) (kg/m3 )
0 8.4 1.225 1.225
100 5.9 5.25E-7 5.75E-7
150 25.5 1.73E-9 1.99E-9
200 37.5 2.41E-10 3.65E-10
250 44.8 5.97E-11 1.20E-10
300 50.3 1.87E-11 4.84E-11
350 54.8 6.66E-12 2.18E-11
400 58.2 2.62E-12 1.05E-11
450 61.3 1.09E-12 5.35E-12
500 64.5 4.76E-13 2.82E-12
550 68.7 2.14E-13 1.53E-12
600 74.8 9.89E-14 8.46E-13
650 84.4 4.73E-14 4.77E-13
700 99.3 2.36E-14 2.73E-13
750 121 1.24E-14 1.59E-13
800 151 6.95E-15 9.41E-14
850 188 4.22E-15 5.67E-14
900 226 2.78E-15 3.49E-14
950 263 1.98E-15 2.21E-14
1,000 296 1.49E-15 1.43E-14
1,250 408 5.70E-16 2.82E-15
1,500 516 2.79E-16 1.16E-15
2,000 829 9.09E-17 3.80E-16
3.7. APPENDIX 151

Table 3.5: Atmospheric Scale Height and Density from 2,500 km to


35,786 km

Altitude Scale Height Atmospheric Density


(km) (km) Mean Maximum
(kg/m3 ) (kg/m3 )
2,500 1220 4.23E-17 1.54E-16
3,000 1590 2.54E-17 7.09E-17
3,500 1900 1.77E-17 3.67E-17
4,000 2180 1.34E-17 2.11E-17
4,500 2430 1.06E-17 1.34E-17
5,000 2690 8.62E-18 9.30E-18
6,000 3200 6.09E-18 5.41E-18
7,000 3750 4.56E-18 3.74E-18
8,000 4340 3.56E-18 2.87E-18
9,000 4970 2.87E-18 2.34E-18
10,000 5630 2.37E-18 1.98E-18
15,000 9600 1.21E-18 1.16E-18
20,000 14600 7.92E-19 8.42E-19
25,000 20700 5.95E-19 6.81E-19
30,000 27800 4.83E-19 5.84E-19
35,000 36000 4.13E-19 5.21E-19
35,786 37300 4.04E-19 5.12E-19

3.7.2 International Geomagnetic Reference Field


The coefficients from the International Geomagnetic Reference Field (IGRF)
and Definitive Geomagnetic Reference Field (DGRF) from 1900 to 2000 are
listed in Table 3.6, along with the secular variation terms valid for the years
2000 to 2005.
152 CHAPTER 3. FORCES AND TORQUES

Table 3.6 (a): Spherical Harmonic Coefficients of the IGRF and


DGRF models (Ref. [10])
3.7. APPENDIX 153

Table 3.6 (b): Spherical Harmonic Coefficients of the IGRF and


DGRF models (Ref. [10])
154 CHAPTER 3. FORCES AND TORQUES

Table 3.6 (c): Spherical Harmonic Coefficients of the IGRF and


DGRF models (Ref. [10])
3.7. APPENDIX 155

Figure 3.13: Variation in the IGRF coefficient g0


156 CHAPTER 3. FORCES AND TORQUES

Problem Set 3

3.1 Show that the gravitational perturbation acceleration due to a planet


on the satellite-Earth two body system (Fig. 3.14) with respect to the
orbital reference frame îĵ k̂ is
f~d = fx î + fy ĵ + fz k̂ (3.69)
where
µp r h i
fx = 3(î · î p )(î p · î) − 1 (3.70)
rp3
3µp r
fy = 3 (î · îp )(îp · ĵ) (3.71)
rp
3µp r
fz = 3 (î · îp )(îp · k̂) (3.72)
rp

Planet
Inertial Fps
Reference
Frame rp rps Satellite
O Fp
Fe
R Satellite
−Fe
Re r
Earth
Earth Free−Body Diagrams

Figure 3.14: Planetary Gravitational Perturbation.

3.2 Show that the aerodynamic drag acting on a satellite orbiting the Earth
with respect to orbital reference frame îo ĵo k̂o is
f~ = f~x îo + fy ĵo + fz k̂o (3.73)
where
1 esinθ
fx = − CD ρAv 2 (3.74)
2m (1 + e + 2ecosθ)1/2
2

1 1 + ecosθ
fy = − CD ρAv 2 (3.75)
2m (1 + e2 + 2ecosθ)1/2
fz = 0 (3.76)
Assume the atmosphere is stationary and not rotating with the Earth.
3.7. APPENDIX 157

3.3 Determine eclipse period for a satellite in a medium earth orbit (alti-
tude=5000 km).
3.4 Derive the SRP force on a satellite surface (assumed as flat plate) with
respect to orbital reference frame. Assume the satellite surface is highly
reflective (i.e., ρa ≈0 and ρd ≈0) and the front surface is only illumi-
nated.
3.5 In Example 3.5, plot the SRP force profile for the case Ω=ω=0, ρs =0.7,
and =23 deg 27 min. Take the orbit semimajor axis a=6378.14+600
km and orbit inclination i any value between 0o to 90o .
For the preliminary simulation, we can assume Ω=ω=0, ρs =0.7, and
=23 deg 27 min. Take the orbit semimajor axis a=6378.14+600 km
and orbit inclination i any value between 0o to 90o .
3.6 Show that the Earth’s magnetic field strength for a satellite in equato-
rial orbit (orbital radius R) is approximately B = µf /R3 while in polar
orbit, the Earth’s magnetic field strength increases by a factor of two
(i.e., B = 2µf /R3 ).

3.7 Prove that the components of B ~ along the pitch axis remain nearly
constant if the satellite motion is in a circular orbit.
3.8 Show that the magnetic force between two satellites; each equipped
with magnetic coil of coil radius a, number of coil turns N, and the
current I passing through it is

3 µ0 πN 2 I 2 a4 d~
F~ = (3.77)
2 d4 d
where d is the distance vector between the satellites. Assume a  d.
158 CHAPTER 3. FORCES AND TORQUES
Chapter 4

Dynamics I

This chapter is focused on Orbital dynamics, also called “orbital mechanics”


or “spaceflight dynamics”, the study of the motions of spacecraft under the
influence of gravity, atmospheric drag, solar radiation pressure, thrust, and
other external forces. This chapter starts with the basic laws of motion at-
tributed to Kepler and Newton. Then the simplest case of the motion of two
bodies under mutual gravitational force of attraction is examined using carte-
sian or polar coordinates and orbital elements. The chapter concludes with
a discussion of the effects of external forces including gravity, atmospheric
drag, and other forces on the motion of a two-body system.

4.1 Introduction
The foundation of this subject was laid down by Johannes Kepler (1571-1630)
and Issac Newton (1642-1727). Kepler (1571-1630) stated three empirical
laws of the orbital motion of planets based on astronomical observations.
The first two laws were published in 1609, while the third law was published
in 1619. These laws were later mathematically proved by Newton and he
presented the law of gravitation along with three laws of motion in Principia
in 1687. These laws can be used to derive the equations of motion of a point
mass system or in other words, orbital motion of a system. We start with
Newton’s laws of motion.
160 CHAPTER 4. DYNAMICS I

4.2 Newton’s Laws of Motion


Newton stated three laws of motion as follows.
First Law
A particle remains at rest or constant velocity if there is no external force
acting on it, i.e.,

V~ = 0 or constant if F~ = 0 (4.1)

where V~ is the velocity of a particle and F~ is the external force. This law
defines reference frames called inertial reference frames in which Newton’s
laws of motion are valid. Thus, inertial reference frames are those reference
frames that are non-accelerating and non-rotating in space; but they may be
fixed or moving with constant velocity.
Second Law
The time rate of change of linear momentum of a particle is equal to the
external force acting on it and is in the direction of this force, i.e.,

d~
p
F~ = (4.2)
dt

where F~ is the external force acting on the particle of constant mass m. For
a point mass m described by the position vector R ~ and velocity vector V
~,
measured with reference to an inertial reference frame (Fig. 4.1), we have

d(mV~) ~
d2 R
F~ = =m 2 (4.3)
dt dt

m F

O
Y

Figure 4.1: Newton’s second law of motion.


4.2. NEWTON’S LAWS OF MOTION 161

Third Law
When two bodies interact, the force exerted by one body on another body
is equal in magnitude but opposite in direction to the force experienced by
the other body due to the first body (Fig. 4.2). This interaction can occur
even if the bodies are at a distance. For example, the gravitational force of
attraction between the Earth and a spacecraft. Note that these action and
reaction forces act on different bodies, not on the same body, and they are
collinear. We can write this law mathematically referring to (Fig. 4.2) as

F~12 = −F~21 (4.4)

where F~12 is the force experienced by a body of mass m1 due to a body of


mass m2 while F~21 is the force experienced by a body of mass m2 due to the
body of mass m1 .

r
m2
F21
F12
m1
F12 = −F21

Figure 4.2: Mutual forces of interaction between two bodies.

Remarks. 1. Newton’s laws of motion are equally valid for a system


of particles or a rigid body if the position, velocity, and acceleration vectors
correspond to the position, velocity, and acceleration vectors of its center of
mass and the force is acting on the center of mass of the body.
2. In general, most of the reference frames happens to be non-inertial
reference frames and therefore, in order to apply Newton’s laws of motion, we
need to obtain inertial velocities and accelerations from non-inertial reference
frames. The procedures to find these terms are explained in the previous
chapter on Kinematics.
3. Newton’s second law of motion as stated above is valid for deriving the
orbital motion of a system. In the case of angular motion or attitude motion
of a system, this law is modified to

dH~
T~ = ~˙
=H (4.5)
dt

where T~ is an external torque and H


~ is the system angular momentum.
162 CHAPTER 4. DYNAMICS I

4.3 Kepler’s Laws of Orbit Motion


Kepler described the orbital motion of planetary bodies based on three laws
which are as follows.
First Law
The orbit of each planet is an ellipse with the Sun at one focus (Fig. 4.3).

Ellipse
Planet

Apoapsis Periapsis
Sun
Focus

Figure 4.3: Kepler’s first law.

Second Law
The radius vector from the Sun to a planet sweeps equal areas in equal
time intervals (Fig. 4.4). In other words, the rate of area swept by the radius
vector is constant.

Ellipse Time (t)

Planet Area 1

Apoapsis Periapsis
Sun
Focus

Area 2
Time (t)
Area 1 = Area 2

Figure 4.4: Kepler’s second law.

Third Law
The square of the period of planetary motion is proportional to the cube
4.4. TWO-BODY MOTION 163

of the semimajor axis of the orbit,i.e.,


T 2 ∝ a3 (4.6)
where T is the orbital period and a is the orbit semimajor axis.
In fact Kepler stated these laws based on astronomical observations and
the proofs of these laws were provided by Newton more than fifty years later.

4.4 Two-Body Motion


Two-body motion involves two bodies, say m1 and m2 at a relative distance
of r from each other (Fig. 4.5) under mutual gravitational force of attraction.
This two-body motion is also referred to as a two-body problem. Our interest
is to derive the equations of motion of these bodies. We consider Newton’s
second laws of motion to derive the equations of motion. To do so, we need
to select an inertial reference frame. Let O − XY Z be an inertial reference
frame. The vectors R ~ 1 and R
~ 2 describe the positions of masses m1 and m2
from the center of the inertial reference frame O.
Z

m2 F
m2 −F
r
R2 m1
Rc
m1
R1
O
Y

Figure 4.5: Two-body motion under gravitational force of attraction.

The equations of motion of the two bodies can be written as


~¨ 1
F~1 = m1 R
X
(4.7)
~¨ 2
F~2 = m2 R
X
(4.8)
164 CHAPTER 4. DYNAMICS I

~¨ j
F~j is the sum of the external forces acting on the jth body and R
P
where
is the inertial acceleration of the jth body.
Using Newton’s law of gravitation to derive the forces acting on the bodies
~2 −R
m1 and m2 with the relative position vector ~r = R ~ 1 , we can rewrite the
equations of motion as

~¨ 1 = Gm1 m2 ~r
m1 R (4.9)
r3
~¨ 2 = − Gm1 m2 ~r
m2 R (4.10)
r3

~ j expressed by its cartesian components (Xj ,Yj ,Zj ,


or in a scaler form with R
j = 1, 2) as

Gm1 m2
m1 k̈1 = (k2 − k1 ), k = X, Y, Z (4.11)
r3
Gm1 m2
m2 k̈2 = − (k2 − k1 ), k = X, Y, Z (4.12)
r3
qP
where r = k=X,Y,Z (k2 − k1 )2 . These equations are nonlinear coupled
differential equations with twelve state variables (k1 , k̇1 , k2 , k̇2 , k = X, Y, Z)
and thus, they can only be solved numerically using twelve initial conditions.
However, if we reformulate the two-body problem, we can obtain a closed-
form solution. In order to do so, we simplify the problem. The motion of the
system center of mass is first analyzed and then the relative motion between
the two bodies is studied.
The center of mass of the system Rc can be expressed as

~ ~
~ c = m1 R1 + m2 R2
R (4.13)
m1 + m2

Differentiating twice with respect to time leads to

~¨ ~¨
~¨ c = m1 R1 + m2 R2
R (4.14)
m1 + m2

Adding Eqs. (4.9) and (4.10), we have

~¨ 1 + m2 R
m1 R ~¨ 2 = 0 (4.15)

Using this relation into Eq. (4.14), we obtain

~¨ c = 0
R (4.16)
4.4. TWO-BODY MOTION 165

Thus, the inertial acceleration of the system center of mass is null. Integrating
this equation twice with respect to time leads to

R ~˙ c (t0 ) [t − t0 ] + R
~c = R ~ c (t0 ) (4.17)

where R ~˙ c (t0 ) are initial position and velocity vectors of the sys-
~ c (t0 ) and R
tem center of mass at t = t0 . The velocity vector R ~˙ c remains constant as
~˙ c (t0 ). Expressing R
R ~ c (t0 ) and R~˙ c (t0 ) in a scaler form, there are six initial
conditions. Thus, the solution in Eq. (4.17) is one-half of the solution of
the two-body problem. The other-half of the solution involving another six
initial conditions is derived next.
Now, we examine the relative motion of the body, say m2 with respect to
the body m1 specified by the position vector ~r. Dividing Eq. (4.10) by m2
and Eq. (4.9) by m1 and then subtracting them leads to

R ~¨ 1 = − G (m1 + m2 )~r
~¨ 2 − R (4.18)
r3
~¨ 2 − R
Using ~r¨ = R ~¨ 1 , we obtain the relation equation of motion as
µ
~r¨ + 3 ~r = 0 (4.19)
r
where µ = G(m1 + m2 ) is called the gravitational parameter. The parameter
µ is more accurately known than either G or masses of the bodies m1 or
m2 , as µ can be derived from satellite tracking data with a high degree of
accuracy while the determination of G is limited by the challenges involved
in conducting laboratory experiments with known masses.
For most cases of practical interest, one of the masses in two-body motion
is much greater than the other. If, say, m1  m2 , then the gravitational pa-
rameter µ can be approximately expressed as µ ≈ Gm1 . Here the parameter
µ = Gm1 is called the heliocentric gravitational constant when m1 is the
mass of the Sun, while it is termed as the geocentric gravitational constant
in the case m1 is the mass of the Earth. If the body of mass m2 orbits around
an inertially fixed body of mass m1 , we call this problem a restricted two-body
problem and the motion of m2 is known as central force motion.
Now, in order to write the relative equation of motion Eq. (4.19) in a
scaler form, we express the relative vector ~r by its cartesian components
(x,y,z) as
~r = xIˆ + y Jˆ + z K̂
Furthermore, assuming cartesian coordinate axes I, J and K as inertial axes
(i.e., Iˆ = Jˆ = K̂ = 0), we can write acceleration r̈ by differentiating the
above equation twice
~r¨ = ẍIˆ + ÿ Jˆ + z̈ K̂
166 CHAPTER 4. DYNAMICS I

Substituting this relation in Eq. (4.19), we obtain the relative equation of


motion in the scaler form as
µ
ẍ + x=0 (4.20)
r3
µ
ÿ + 3 y = 0 (4.21)
r
µ
z̈ + 3 z = 0 (4.22)
r
p
where r = x2 + y 2 + z 2 . These equations are nonlinear coupled differential
equations with six state variables (i.e., x, ẋ, y, ẏ, z, ż). However, we can solve
the vector equation Eq. (4.19) analytically. Before proceeding to solve it, we
first determine all the constants of the motion or “constants of integration”
called integrals or fundamental integrals of the motion. In fact, even without
knowing the solution of Eq. (4.19), we can know its properties referring to
these integrals. Indeed using these integrals, we next derive the closed-form
solution of the relative equation of motion.
We have already found that the velocity vector of the system center of
mass R ~˙ c remains constant. The R~˙ c has three components in any chosen coor-
dinate frame (i.e., cartesian, polar or cylindrical) constituting three constants
of integrals of the given two-body system. The remaining integrals are found
in the following sections.

4.4.1 Angular Momentum Vector


In the given two-body system, the two bodies are only acted upon by mutual
gravitational force of attraction. As we know that the gravitational is a
conservative force, we expect the angular momentum of the two-body system
to remain constant or, in other words, the angular momentum is conserved.
We shall prove this fact in the following derivation.
Taking the cross product of Eq. (4.19) with ~r, we have
µ
~r × ~r¨ + 3 (~r × ~r) = 0 (4.23)
r

Knowing that ~r × ~r=0 and

d  
~r × ~r˙ = ~r˙ × ~r˙ + ~r × ~r¨ = ~r × ~r¨
dt
we can rewrite Eq. (4.23) as

d  
~r × ~r˙ = 0 (4.24)
dt
4.4. TWO-BODY MOTION 167

Integrating, we get

~r × ~r˙ = constant vector = ~h (4.25)

where the constant vector ~h is the angular momentum per unit mass defined
by ~h = ~r × ~r˙ . Thus, the angular momentum is conserved in the two-body
system.

4.4.2 Energy
In the two-body system, we expect the mechanical energy to be conserved as
the gravitation force of attraction acting on the two bodies is a conservative
force. We shall prove this fact in the following derivation.
Taking the dot product of Eq. (4.19) with ~r˙ , we have
µ  
~r˙ · ~r¨ + 3 ~r˙ · ~r = 0 (4.26)
r
or
1 d  ˙ ˙ d µ
~r · ~r − =0 (4.27)
2 dt dt r

Integrating, we get
1 ˙2 µ
~r − = constant = E (4.28)
2 r
where the constant E is the mechanical energy of the system per unit mass
or specific mechanical energy and it is the sum of the system kinetic energy
per unit mass, ~r˙ 2 /2 plus the system potential energy per unit mass, −µ/r.
As per the derivation, the mechanical energy ε remains constant.

4.4.3 Eccentricity Vector


In this section, we introduce an eccentricity vector using the following deriva-
tions.
Taking the post-cross product of Eq. (4.19) with ~h, we have
µ  
~r¨ × ~h + 3 ~r × ~h = 0 (4.29)
r

Using the relations


d ˙ ~ ¨ ~ ˙
~r × h = ~r × h + ~r˙ × ~h = ~r¨ × ~h
dt
168 CHAPTER 4. DYNAMICS I

and ~h = ~r × ~r˙ in Eq. (4.29) lead to


d ˙ ~ µ h i
~r × h + 3 ~r × (~r × ~r˙ ) = 0 (4.30)
dt r
Applying the triple vector product relation: ~a × (~b × ~c) = (~a · ~c)~b − (~a · ~b)~c
yields
d  ˙ ~  µ h ˙  i
~r × h + 3 ~r · ~r ~r − (~r · ~r)~r˙ = 0 (4.31)
dt r
Further, integrating
µ
~r˙ × ~h − ~r = constant vector = µ~e (4.32)
r
where µ~e is a constant vector and ~e is called the eccentricity vector.
We have thus found three components each of the constant vectors: the
velocity vector of the system center of mass R~˙ c , angular momentum vector
~h, and eccentricity vector ~e plus the mechanical energy E. These constant
components are the integrals of the motion and so, in total, only ten integrals
of the motion exist for the two-body problem as well as for the n-body problem.

4.4.4 Solution of Relative Motion


There are various ways to obtain the solution of the relative equation of
motion of the two-body problem, expressed by Eq. (4.19). We consider here
a vector method.
Taking the dot product of Eq. (4.32) with ~r, we have

µ
~r · (~r˙ × ~h) − (~r · ~r) = µ~r · ~e (4.33)
r

Applying the relation for triple scaler product ~a · (~b × ~c) = (~a × ~b) · ~c leads
to

 
~r × ~r˙ · ~h − µr = µ~r · ~e (4.34)

We assume the angle between ~r and ~e be θ, called true anomaly. Con-


sidering this fact and knowing ~h = ~r × ~r˙ , the preceding equation simplifies
to

h2 = µr(1 + ecosθ)
4.5. CONIC SECTIONS 169

or
p
r= (4.35)
1 + ecosθ
where p = h2 /µ is a constant parameter. This equation is the solution of
the relative equation of motion of the two-body problem and it specifies the
motion of m2 with respect to m1 . However, we are unable to determine
whether the path (i.e., orbit) of m2 will be closed or open. To understand
this, we study Conic Sections in the next section.

4.5 Conic sections


A conic section is the locus of a point, say S, that moves in the plane of a
fixed point, called the focus F , and a fixed line, called the directrix (with F
not on the directrix) such that the ratio of the distances from the focus and
the directrix is a constant, called eccentricity, denoted by e (Fig. 4.6). It can
be mathematically expressed as
r
e= (4.36)
d − rcosθ
where d is the distance between the focus F and the directrix. This relation
can be rewritten as

r = e(d − rcosθ) (4.37)

which mathematically states that the locus of the point S is a conic section
that include all its possible paths such that its distance from the focus F is
a constant fraction, e, of its distance from the directrix.
Depending upon orbital eccentricity and energy, we have four possible
conic sections: circle, ellipse, parabola, and hyperbola (Fig. 4.7). These
names were given by Apollonius, a famous geometer and he was the first to
study them systematically around 200 B.C. These sections are so called conic
sections as they can be obtained by slicing through a right circular cone at
various different angles, as illustrated in Fig. 4.8. In the following sections,
we describe the details of the conic sections.

4.5.1 Ellipse
An ellipse is formed by the locus of a point, S, in the plane the sum of whose
distances r1 and r2 from two foci F1 and F2 has a constant value 2a (Fig.
4.9), i.e.,

r1 + r2 = 2a (4.38)
170 CHAPTER 4. DYNAMICS I

Orbit Directrix

S r/e

Focus (F )
d

Figure 4.6: Geometry of a conic section.

Hyperbola
e >1 , ε >0
Parabola
e =1 , ε =0
v S

Ellipse
0< e <1 , ε <0 r
Circle
e =0 , ε <0 θ
Apoapsis Periapsis
F

Figure 4.7: Comparisons between conic sections.

where 2a is the major axis of the ellipse. The points in the orbit nearest to
and farthest from the focus, say F1 , are called apsides. The point of the orbit
closest to F1 is called the periapsis and the point farthest from F1 is called
the apoapsis. The line joining them is known as line of apsides. They have
specific names for different planetary orbits, as mentioned in Table 4.1.
Using Eqs. (4.36) and (4.38), we can derive parameters of ellipse as shown
4.5. CONIC SECTIONS 171

Cutting plane

Circle
Hyperbola

Ellipse
Parabola

Figure 4.8: Conic sections by slicing through a right circular cone.

Local Horizontal v Directrix


γ S r/e
Ellipse r2
b r1=r
Minor Apoapsis θ Periapsis
axis F2 O F1
b c c=ae d=p/e
p

a a
Major axis

Figure 4.9: Ellipse

in Fig. 4.9 as
c
e= (4.39)
a p
b 2 + c2 = a 2 ⇒ b = a 1 − e 2 (4.40)
2
p = a(1 − e ) (4.41)
172 CHAPTER 4. DYNAMICS I

Table 4.1: Periapsis and apoapsis for planetary orbits

Celestial body Periapsis Apoapsis


Earth Perigee Apogee
Moon Perilune or Apolune or
Periselenium Aposelenium
Jupitor Perijove Apojove
Sun Perihelion Apohelion

where the eccentricity e varies as 0 < e < 1. The parameter c is the distance
between the focus F1 or F2 from the center of the ellipse, O. The parameter
2a as stated earlier is the major axis of the ellipse while the parameter 2b
is the minor axis of the ellipse (assuming 2b < 2a). So, the halves of these
parameters, a and b are called the semimajor and semiminor axes, respec-
tively. The parameter 2p is the distance between two intersecting points, on
the conic section, of a line that is drawn perpendicular to the major axis at
one of its foci F1 or F2 and it is called the latus rectum and p is called the
semilatus rectum or simply parameter. The parameter p determines the size
of the conic section while the eccentricity e states its shape.
The distances of periapsis and apoapsis, denoted by rp and ra , from the
focus F1 can be expressed in terms of a and e as

rp = a(1 − e) (4.42)
ra = a(1 + e) (4.43)

Alternatively, if rp and ra are known, a and e can be obtained as

rp + ra
a= (4.44)
2
ra − rp
e= (4.45)
ra + rp

The equation of the ellipse can be obtained by rewriting the equation of


the conic section, Eq. (4.36) as
ed p
r= = (4.46)
1 + ecosθ 1 + ecosθ
4.5. CONIC SECTIONS 173

where the distance from the directrix, d can be found using geometrical re-
lations as d = p/e. Note that when θ = ±90o , the radial distance r equals to
the semilatus rectum p. A summary of parameters for all conic sections are
presented in Table 4.2.

Table 4.2: Parameters of conic sections

Conic Eccentricty Semimajor Semiminor Radius


Section (e) axis (a) axis (b) (r)
Circle e=0 a>0 b=a r=a
√ a(1 − e2 )
Ellipse 0<e<1 a>0 b = a 1 − e2
1 + ecosθ
e = c/a
2rp
Parabola e=1 a=∞ b=∞
1 + cosθ

√ a(e2 − 1)
Hyperbola e>1 a<0 b = a e2 − 1
1 + ecosθ
e = c/a

We can also describe the ellipse in a cartesian coordinate frame assuming


the origin of the coordinate system at center of the ellipse, O:
X2 Y 2
+ 2 =1 (4.47)
a2 b
where the X-axis is along the major axis towards periapsis and the Y-axis is
along the minor axis. From this equation we can obtain the polar equation
of the ellipse given by Eq. (4.46).
Now let us go back to the solution of the relative equation of motion, Eq.
(4.35). This solution is in fact the same as the equation of an ellipse, Eq.
(4.46). Thus, we can say that an orbit of a body of mass m2 is an ellipse with
a fixed body of mass m1 at one focus, say F1 . This is Kepler’s first law of
planetary motion that states that the orbit of each planet is an ellipse with
the Sun at one focus.

4.5.2 Parabola
A parabola is the locus of points whose distance from a focus is equal to the
distance from the directrix (Fig. 4.10), i.e.,
r = d − rcosθ (4.48)
174 CHAPTER 4. DYNAMICS I

Directrix

S Parabola

r1=r

θ
F1
p

d=p

Figure 4.10: Parabola

The parameters of the parabola and its equation are as follows


e=1 (4.49)
a=c=∞ (4.50)
p
r= (4.51)
1 + ecosθ
The parameter p of a parabola is related with the periapsis distance as

p = 2rp (4.52)

Table 4.2 presents these parameters. There are several examples of the
parabola including the motion of projectiles under uniform gravity and the
path of parallel rays of light to a focus. However, parabolic orbits are not
useful spacecraft trajectories.

4.5.3 Hyperbola
A hyperbola is a conic section defined as the locus of all points P in the
plane the difference of whose distances r1 and r2 from two foci F1 and F2 is
a constant 2a (Fig. 4.11), i.e.,
r2 − r1 = 2a (4.53)

Using this relation and the equation of a conic section, Eq. (4.36), we can
derive parameters of a hyperbola as
4.5. CONIC SECTIONS 175

Directrix Asymptote
S
Hyperbola
r2
r1=r
δ b
β θ
F2 O β F1
a a p
c c=ae
d

Figure 4.11: Hyperbola

c
e= (4.54)
a p
b2 + c2 = −a2 ⇒ b = a e2 − 1 (4.55)
2
p = a(e − 1) (4.56)

where the eccentricity e > 1 and a < 0. The equation of the hyperbola is

a(e2 − 1)
r= (4.57)
1 + ecosθ

A hyperbola can be described in a cartesian reference frame considering


the origin of the coordinate system at center of the hyperbola, O: as

X2 Y 2
− 2 =1 (4.58)
a2 b
where the X-axis is along the major axis towards periapsis and the Y-axis is
along the minor axis. Unlike the ellipse, no points of the hyperbola actually
lie on the semiminor axis, but rather the ratio determines the vertical scaling
of the hyperbola.
Examples of hyperbolic orbits include Earth departure on planetary flights,
planetary arrival and targeting, and comet flyby about the Sun. Hyperbolic
planetary flyby orbits are considered for energetic gravity-assist maneuvers.
The circle is a special case of a conic section when eccentricity e = 0. The
conic sections: circle, ellipse, parabola, and hyperbola just described not only
176 CHAPTER 4. DYNAMICS I

differ by eccentricity but they also have different energies. Let us find the
energies associated with these conic sections.
The mechanical energy of an orbit, as discussed in earlier section, is given
by Eq. (4.28):

v2 µ
E= − (4.59)
2 r
which is the sum of the kinetic energy, T = v 2 /2 and potential energy, U =
µ/r per unit mass. Here r and v denote the position and velocity of the body
of mass m2 . We need to determine these to obtain the orbit energy ε. Note
that in fact as v does not say direction and so it denotes speed, but we will
call it velocity everywhere in the text.
The position r of the orbiting body, as we obtained earlier, is
p
r= (4.60)
1 + ecosθ
where p(semilatus rectum) = h2 /µ.
To determine velocity v, we need to find velocity vector ~v . Writing ~r and
~r˙ in terms of a cartesian coordinate frame îĵ k̂ with î aligned with ~r as

~r = rî (4.61)
~r˙ = ~v = ṙ î + rî˙ = ṙ î + r(~
ω × î) (4.62)

Knowing ω
~ = θ̇k, we have

~v = ṙ î + rθ̇ k̂ × î = ṙî + rθ̇ ĵ (4.63)

or

~v · ~v = v 2 = ṙ2 + r2 θ̇2 (4.64)

Differentiating Eq. (4.60) with respect to time leads to


˙
pθesinθ ˙
r2 θesinθ
ṙ = 2
= (4.65)
(1 + ecosθ) p

We write the angular momentum ~h, as explained earlier, as


~h = ~r × ~r˙ (4.66)
4.5. CONIC SECTIONS 177

Using Eq. (4.63) for ~r˙ , we have

~h = ~r × (ṙ î + rθ̇ ĵ) = r2 θ̇k̂ (4.67)

or

h = r2 θ̇ (4.68)

Applying this relation in Eq. (4.65) and knowing h = µp lead to

hesinθ µ
r
ṙ = = esinθ (4.69)
p p

Using this equation along with Eq. (4.60) into Eq. (4.64), we have

µe2 sin2 θ µ(1 + ecosθ)2


v2 = +
p p
µ
= (1 + e2 + 2ecosθ) (4.70)
p

The kinetic energy per unit mass or specific kinetic energy, T can be
expressed as

v2 µ
T = = (1 + e2 + 2ecosθ) (4.71)
2 2p

The potential energy per unit mass or specific potential energy, U is

µ µ(1 + ecosθ)
U =− =− (4.72)
r p

Thus, the specific energy, using Eq. (4.59), can be expressed as

v2 µ µ(1 − e2 )
E =T +U = − =− (4.73)
2 r 2p

In case the orbit is an ellipse, substituting p = a(1 − e2 ) into the above


equation leads to

v2 µ µ
E= − =− (4.74)
2 r 2a
This equation states the energy of an elliptic orbit. We can find the energies
of other conic sections by using the same equation and taking a > 0 and
178 CHAPTER 4. DYNAMICS I

a = r for a circular orbit, a = ∞ for a parabolic orbit, and a < 0 for a


hyperbolic orbit, i.e.,
µ
E=− for circular orbit (4.75)
2r
E = 0 for parabolic orbit (4.76)
µ
E= for hyperbolic orbit (4.77)
2a
Thus, the specific energy of an orbit depends only on the semimajor axis of
the orbit. The energies for elliptic and circular orbits are negative while the
energy for a parabolic orbit is zero. The hyperbolic orbit has positive energy.
Note for all closed orbits specific energy is negative.
We can rewrite Eq. (4.74) in terms of velocity v for an elliptic orbit as
r
2µ µ
v= − (4.78)
r a
This equation is known as the vis-viva equation or energy equation. Here the
name vis viva means living forces in classical mechanics.
We can derive the velocities in circular (a = r), parabolic (a = ∞), and
hyperbolic (a < 0) orbits using the above equation as
r
µ
v= for circular orbit (4.79)
r
r

v= for parabolic orbit (4.80)
r
r
2µ µ
v= + for hyperbolic orbit (4.81)
r a
Thus, for a given position r the velocities and energies increase in the order:
circular, elliptical, parabolic, hyperbolic.

4.6 Orbit Motion in Relation with Time


Let us go back to the solution of the relative motion, given by Eq. (4.35).
This solution describes the position r of the body of mass m2 with respect
to the body of mass m1 at the focus, F1 as the true anomaly θ varies. How-
ever, this does not provide the time taken to move from one point to another
in the orbit. In fact, for carrying out various mission operations including
deployment of solar arrays, ignition of thrusters, and establishment of com-
munication links, we need to know the time with respect to the position of a
spacecraft in orbit.
4.6. ORBIT MOTION IN RELATION WITH TIME 179

In the case of a spacecraft in a circular orbit, the velocity of the spacecraft


remains constant and therefore, the angle θ (a reference angle instead of a
true anomaly as the true anomaly is undefined for circular orbits) is directly
proportional to time t. But for a spacecraft in an elliptic orbit, the velocity
is not uniform and varies as per Eq. (4.78) derived earlier. The velocity is
maximum at periapsis with rp = a(1 − e) and minimum at apoapsis with
ra = a(1 + e). Therefore, the time travelled differs throughout the orbit
and it is not directly related to true anomaly. In order to find the relation
between the time t and the true anomaly θ for the spacecraft in an elliptic
orbit, we introduce two angles called eccentric anomaly and mean anomaly.
The eccentric anomaly, denoted by E, is the angle between the direction
of periapsis and the current position of a spacecraft on its orbit, projected
onto the ellipse’s circumscribing auxiliary circle as shown in Fig. 4.12. It can
be expressed mathematically as
Auxiliary Circle

Orbit a
b r
E θ Periapsis
O F
ae

Figure 4.12: Eccentric anomaly for an elliptic orbit.

acosE = ae + rcosθ (4.82)

Substituting r = p/(1 + ecosθ) = a(1 − e2 )/(1 + ecosθ) and simplifying,


we have
e + cosθ
cosE = (4.83)
1 + ecosθ
We can also rewrite as
r
E 1−e θ
tan = tan (4.84)
2 1+e 2
180 CHAPTER 4. DYNAMICS I

Alternatively, we can write in terms of the true anomaly θ as

e − cosE
cosθ = (4.85)
ecosE − 1
It is to be noted here that θ and E are always in the same quadrant of the
orbit.
Next we define mean anomaly. The mean anomaly, denoted by M , is an
angle given by

M = n(t − tp ) (4.86)

where tp is the starting time which can be taken at the periapsis. The n is
the mean motion or average angular velocity given as
r
µ 2π
n= = (4.87)
a3 T

where T is orbital period. We can rewrite Eq. (4.86) using n = 2π/T as

2π(t − tp )
M= (4.88)
T
Thus, the the mean anomaly, M , is the fraction of an orbit period which has
elapsed since periapsis passage.
Next our task is to relate the mean anomaly M to the eccentric anomaly
E. We write the angular momentum ~h as derived in Eq. (4.68) considering
a cartesian coordinate frame îĵ k̂ with î aligned with ~r as

h = r2 θ̇ (4.89)

Using h2 /µ = p = a(1 − e2 ) leads to


p
µa(1 − e2 ) = r2 θ̇ (4.90)

We derive θ̇ using Eq. (4.84). Differentiating Eq. (4.84) with respect to


time and rearranging and simplifying yield
  "r #
2 θ dθ 1 + e 2 E dE
sec = sec (4.91)
2 dt 1−e 2 dt

or
  "r #
θ dθ 1 + e E dE
1 + tan2 = sec2 (4.92)
2 dt 1−e 2 dt
4.6. ORBIT MOTION IN RELATION WITH TIME 181

Replacing θ in terms of E using Eq. (4.84) leads to


    "r #
1+e 2 E dθ 1 + e 2 E dE
1+ tan = sec (4.93)
1−e 2 dt 1−e 2 dt

Simplifying,

dθ 1 − e2 dE
= (4.94)
dt 1 − ecosE dt
From Fig. 4.12, we can write
rsinθ b p
= = 1 − e2 (4.95)
asinE a
Rewriting Eq. (4.82) as

rcosθ = ae − acosE

and squaring both sides of this equation as well as Eq. (4.95), and adding
them yield

r = a(1 − ecosE) (4.96)

Substituting θ̇ = dθ/dt and r from Eqs. (4.94) and (4.96) into Eq. (4.90)
leads to

p 2 2 1 − e2 dE
µa(1 − e2 ) = a (1 − ecosE) (4.97)
1 − ecosE dt
or
r
µ
dt = (1 − ecosE)dE (4.98)
a3

Integrating,
r t E
µ
Z Z
dt = (1 − ecosE)dE (4.99)
a3 t=tp E=0

where tp is the time at periapsis; at periapsis E = 0.

n(t − tp ) = E − esinE (4.100)

Using the definition of the mean anomaly given by Eq. (4.86), we can write

M = n(t − tp ) = E − esinE (4.101)


182 CHAPTER 4. DYNAMICS I

This equation was in fact derived by Kepler and it is known as Kepler’s time
equation. This equation always gives the smallest value of the elapsed time
since perigee passage (t − tp ) (i.e., less than one half of the orbital period),
which is only correct for θ ≤ π. For the case θ ≥ π, we must subtract (t − tp )
from the orbital period. Note that M is equal to the eccentric anomaly at
periapsis and apoapsis.
We can obtain the time period of the orbit, denoted by T , using the above
equation (4.101). When t − tp equals to T , thepeccentric anomaly E = 2π.
Using these into Eq. (4.101) and knowing n = µ/a3 , we have

nT = E − esinE = 2π − esin(2π) = 2π (4.102)

or
2πa3/2
T = √ (4.103)
µ

Further, we can write

T 2 ∝ a3 (4.104)

This equation states that the square of the period of planetary motion is
proportional to the cube of the semimajor axis of the orbit. This is known
as Kepler’s third law.
The equation (4.101) relates the eccentric anomaly E with the time after
periapsis passage t − tp . We already derived the equation (4.83) relating
eccentric anomaly E with the true anomaly θ. The relation between the true
anomaly θ and the position r is obtained in the solution of the two-body
problem for an elliptic orbit given by Eq. (4.46). Thus, for a given position
r or the true anomaly θ of a spacecraft with a known semimajor axis a and
orbital eccentricity e, we can determine the time t if the time at the perigee
passage tp is given.
However, for a given time t, it is impossible to obtain the true anomaly θ
or the position r of the spacecraft exactly. This is due the the fact that Eq.
(4.101) is an equation involving trigonometric sine function (known as a tran-
scendental equation) and thus, there exists no closed-form solution. There-
fore, we can only solve this equation numerically using Newton-Raphson and
other methods. However, we can derive approximate expressions for small
eccentricities
 
1 3 1 3
E = M + e − e sinM + e2 sin2M + e3 sin3M + O(e4 ) (4.105)
8 2 8
θ = M + 2esinM + 1.25e2sin2M + O(e3 ) (4.106)

applying the Lagrange reversion theorem.


4.7. DESCRIBING ORBIT MOTION IN SPACE 183

4.7 Describing Orbit Motion in Space


In previous sections, we derived the relative equation of motion of two bodies
in a vector form, given by Eq. (4.19) and then considered two approaches
to study this motion. In the first approach, by defining the relative position
vector in a three-dimensional space using a cartesian coordinate frame we
find the solution of the relative equations of motion numerically. However,
this solution is not intuitive as it does not directly provide the shape, size,
and orientation of the orbit in space.
In the second approach, we derived constants of motion leading to an
analytical solution of the relative motion of two bodies. The eccentricity and
semimajor axis of an orbit define the shape and size of an orbit, respectively,
while the orientation of the orbit remains constant under the conservative
force of gravitational attraction. Thus, the analytical solution is very intu-
itive, but it describes only the planar motion of the two-body system. In
the presence of nonconservative forces or perturbations, the motion of the
two bodies will not be planar as the orbit orientation does not remain con-
stant. Therefore, we need to consider the orbit orientation in space besides
the orbit’s shape and size.
In order to define the orbit orientation in space, we select a reference di-
rection and a reference plane. The reference direction is taken as the vernal
equinox direction or the first point of Aries that points at the zodiac constel-
lation Aries, also called the Ram, and is denoted by the sign of the Ram,
Υ (Fig. 4.13). This direction is drawn by a vector from the center of the
Earth to the center of the Sun on the spring equinox day, March 21. The
spring equinox is the day when the length of daylight equals the length of
night; this lies in the northern hemisphere when the Sun appears to cross
the equator from south to north. However, we use the same direction of the
vernal equinox that was found more than 500 years ago, even though it has
now moved into Pisces (the Fish)1 . In fact, there is a small precession of
the equinoxes, about 0.014 deg per year, but we still can consider the vernal
equinox vector as fixed for most purposes.
Next we select a reference plane. The equatorial plane (i.e., the plane of
the Earth’s equator) is taken as the reference plane (see Fig. 4.13). Now we
fix a cartesian coordinate frame O − XY Z with the origin 0 at the center of
the Earth such that the X-axis lies in the equatorial plane and points to the
vernal equinox or the first point of Aries. The Z-axis, perpendicular to the
equatorial plane, is along the spin axis of the Earth towards the North Pole.
Finally, the Y-axis completes the right-hand set in the equatorial plane. A
spacecraft in orbit crosses the equatorial plane at two points: the ascending
node where the spacecraft goes from below the equator (Southern Hemi-
1 Larson, W. J. and Wertz, J. R. (Editors), Space Mission Analysis and Design, Second

edition, Microcosm, Inc., Torrance, CA, 1995.


184 CHAPTER 4. DYNAMICS I

North Pole
K Z
h
Orbit Plane
Descending Node S
Periapsis
r
i θ
e
Equatorial Plane Focus O ω J
i Y

n
Ascending Node

Line of Nodes
Apoapsis
X I
Vernal Equinox

Figure 4.13: Defining position of a spacecraft S in space.

sphere) to above the equator (Northern Hemisphere), and the descending


node where the spacecraft moves from above the equator (Northern Hemi-
sphere) to below the equator (Southern Hemisphere). The line joining these
descending and ascending nodes is called line of nodes.
We can define the position of the spacecraft, S with reference to the
O − XY Z frame using six elements called orbital elements as

(a) Longitude of the ascending node or Argument of aries or Right


ascension of the ascending node (RAAN) (Ω): the angle from
vernal equinox to the ascending node, measured in the reference plane
in a counterclockwise direction as viewed from the northern hemisphere.
Its range is 0 ≤ Ω ≤ 360o.
(b) Inclination of the orbit plane (i): the angle between the orbit plane
and the reference plane with its range of 0 ≤ i ≤ 180o. The orbit
inclination i along with the longitude of the ascending nodeΩ specifies
orientation of the orbit plane with respect to the equatorial plane.
(c) Argument of the periapsis (ω): the angle from the ascending node to
the periapsis, measured in the orbit plane in the direction of spacecraft
motion. It specifies the orientation of the orbit (i.e., periapsis position)
within the orbit plane with its range 0 ≤ ω ≤ 360o.
(d) True anomaly (θ): the angle from periapsis to the spacecraft S, mea-
4.7. DESCRIBING ORBIT MOTION IN SPACE 185

sured in the orbit plane in the direction of spacecraft motion. Thus, it


specifies the position of the spacecraft in the orbit plane with respect to
the periapsis with its range 0 ≤ θ ≤ 360o. Instead of the true anomaly
θ, we can specify the position of a spacecraft from periapsis using other
parameters: Time of periapsis passage, denoted by tp (i.e., time since
periapsis passage) or Mean anomaly M .

(e) Semimajor axis (a): defines the size of the orbit and is related to an
orbit’s energy.

(f ) Eccentricity (e): defines the shape of the orbit.

The orbital elements a, e, i, ω, Ω, and θ are called classical orbital elements.


Taking the time derivatives of the above parameters leads to

dΩ
= 0 (since nodal line and vernal equinox vector are constant) (4.107)
dt
di
= 0 (since angular momentum vector ~h = constant) (4.108)
dt

= 0 (since nodal line and eccentricity vectors ~n, ~e = constant)
dt
(4.109)
dθ h
= 2 (not a constant as r varies) (4.110)
dt r
dM
= n (a constant) (4.111)
dt
da µ
= 0 (since energyε = constant = − ) (4.112)
dt 2a
de
= 0 (since eccentricity vector ~e = constant) (4.113)
dt
(4.114)

Thus, the parameters a, e, i, ω, Ω, and M or tp remain constant in the Ke-


plerian orbit(i.e., relative orbit of two bodies as described by Kepler when
they are under the conservative gravitational force of attraction) and there-
fore, they are also called constants of motion. As a consequence, the orbit
plane remains fixed in inertial space. These orbital elements are not always
defined. There are some exceptions as follows:

(a) Circular orbit (eccentricity e=0): The perigee does not exist and so,
the argument of perigee ω and the true anomaly θ are undefined. We
introduce an alternate orbital element called argument of latitude u, the
angle measured from the ascending node to the spacecraft’s position in
the direction of spacecraft motion.
186 CHAPTER 4. DYNAMICS I

(b) Equatorial orbit (orbit inclination i=0o or 180o): The line of nodes
does not exist and so, the longitude of the ascending node, Ω, and
the argument of perigee, ω are undefined. We introduce an alternate
orbital element called longitude of perigee, $, the angle measured from
the vernal equinox to perigee in the direction of spacecraft motion.
(c) Circular equatorial orbit (eccentricity e=0 and inclination i=0o or
180o): The perigee does not exist and the line of nodes is missing. So,
the longitude of the ascending node, Ω, the argument of perigee ω, and
the true anomaly θ are undefined. We introduce an alternate orbital
element defined as the true longitude l, the angle measured from the
vernal equinox to the spacecraft position in the direction of spacecraft
motion.

4.8 Relating Orbital Elements with Position


and Velocity Vectors
We can define the motion of a spacecraft using position and velocity vectors
or using orbital elements. In the former approach, the spacecraft position
and velocity vectors can be determined using ground-tracking stations or
the global positioning satellite system (GPS). However, in the case of latter
approach, we cannot measure orbital elements directly. Therefore, to un-
derstand the orbit we need to determine these elements for given position
and velocity vectors at a known time. This is called an orbit determination
problem. In addition, we require the reverse, i.e., for given orbital elements,
determine the position and velocity vectors. In this section we will discuss
these two problems. Let us begin with an orbit determination problem.

4.8.1 Determination of Orbital Elements from Position


and Velocity Vectors
The determination of all six orbital elements (a, e, i, ω, Ω, θ or M or t) for a
given spacecraft position ~r and velocity ~v at a known time is explained below.
Semimajor Axis (a)
The semimajor axis a can be found using energy Eq. (4.74) as

v2 µ µ
E= − =−
2 r 2a
or
r
a= (4.115)
2 − rv 2 /µ
4.8. RELATING ORBITAL ELEMENTS WITH POSITION AND VELOCITY VECTORS187

For a given position, r and velocity, v, the semimajor axis, a is determined.


For parabolic trajectories, the energy E = 0 and as a result the semimajor
axis a = ∞.
Eccentricity (e)
Referring to Eq. (4.32) the eccentricity vector ~e can be written as

1 h˙ ~ µ i
~e = ~r × h − ~r
µ r

Knowing the relation ~h = ~r × ~v , we have

1 h 2 µ  i
~e = v − ~r − (~r · ~v )~v (4.116)
µ r

Thus, the eccentricity e can be determined by finding the magnitude of the


eccentricity vector |~e| for a given ~r and ~v .
Inclination (i)
The orbit inclination, i is the angle between the angular momentum vector
~h and the K̂-axis, i.e.,

~h · K̂ hZ
cosi = = (4.117)
h h
If we find the inclination, i, using a calculator or computer and taking the
inverse of this cosine function, we will get only one of the two possible correct
angles and that too the lowest one lying between 0o and 180o. We must
subtract this result from 360o to get the second possible angle. In fact, this
is true for finding the inverse of any trigonometric functions. Incidently, as
the inclination lies in the range 0 ≤ i ≤ 180o, the calculator will give the
correct answer and so, we do not require a quadrant check.
Longitude of the Ascending Node (Ω)
ˆ
The longitude of the ascending Node Ω is the angle from the I-axis and
the line of nodes or the ascending node vector ~n and so, we can write

Iˆ · n̂
cosΩ = (4.118)
|~n|

where ~n can be determined from the cross product of the K̂-axis and angular
momentum vector ~h, i.e.,

~n = K̂ × ~h = K̂ × (~r × ~v ) (4.119)
188 CHAPTER 4. DYNAMICS I

As Ω lies in the range 0 ≤ Ω ≤ 360o , we must check the quadrant as


follows:

0o ≤ Ω ≤ 180o if ~n · Jˆ = nY ≥ 0 (4.120)
180o < Ω ≤ 360o if ~n · Jˆ = nY < 0 (4.121)

Note in the case Ω lies in the range 180o < Ω ≤ 360o, we must subtract Ω
from 360o to get the correct answer.
Argument of the Periapsis (ω)
The argument of periapsis ω is the angle between the ascending node
vector ~n and the eccentricity vector ~e passing through periapsis, measured in
the direction of spacecraft motion, i.e.,

~n · ~e
cosω = (4.122)
|~n|e

Since ω lies in the range 0 ≤ ω ≤ 360o , we check the quadrant as follows:

0o ≤ ω ≤ 180o if ~e · K̂ = eZ ≥ 0 (4.123)
180o < ω ≤ 360o if ~e · K̂ = eZ < 0 (4.124)

Note that if 0 ≤ ω ≤ 180o the perigee lies in the northern hemisphere or


above the equator, whereas if 180o ≤ ω ≤ 360o the perigee is in the southern
hemisphere or below the equator.
True Anomaly (θ)
The true anomaly θ is the angle from the eccentricity vector ~e to the
spacecraft position vector ~r, measured in the direction of spacecraft motion
and it can be mathematically expressed as

~e · ~r
cosθ = (4.125)
er
As θ lies in the range 0 ≤ θ ≤ 360o, we should check the quadrant as stated
below.

0o ≤ θ ≤ 180o if ~r · ~v ≥ 0 (γ ≥ 0) (4.126)
180o < θ ≤ 360o if ~r · ~v < 0 (γ < 0) (4.127)

where γ is flight path angle (i.e., the angle between the local horizontal and
the velocity vector ~v ; see Fig. 3.3). We can determine γ as follows:

h = |~h| = |~r × ~v |
= rvsin(90 − γ) = rvcosγ (4.128)
4.8. RELATING ORBITAL ELEMENTS WITH POSITION AND VELOCITY VECTORS189

or
rv
cosγ = (4.129)
h
The γ lies in the range −90o ≤ γ ≤ 90o and so we do not need a quadrant
check. For given position and velocity vectors, we can find flight path angle
γ as well as true anomaly θ.
Alternate Orbital Elements
In some cases, we cannot determine the orbital elements. For example,
refereing to Eqs. (4.118), (4.122), and (4.125), we have the following obser-
vations:
(i) If n = 0, Ω does not exist.
(ii) If n = 0 and/or e = 0, ω cannot exist.
(iii) If e = 0, θ does not exist.
These cases correspond to particular orbits and we consider alternate
orbital elements as discussed in Section 4.7.
Circular orbit. The case n 6= 0, e = 0 corresponds to a circular orbit
and we define argument of latitude u, as the angle measured from the as-
cending node vector ~n to the spacecraft’s position vector ~r in the direction
of spacecraft motion, i.e.,
~n · ~r
cosu = (4.130)
nr
The angle u lies in the range 0 ≤ θ ≤ 360o , and we check the quadrant as
stated below.

0o ≤ u ≤ 180o if ~r · K̂ = Z ≥ 0 (4.131)
180o < u ≤ 360o if ~r · K̂ = Z < 0 (4.132)

Equatorial orbit. The case n = 0, e 6= 0 refers to an equatorial orbit


and we take an alternate orbital element longitude of perigee, $ the angle
measured from the vernal equinox direction, Iˆ to the eccentricity vector ~e in
the direction of spacecraft motion, i.e.,

Iˆ · ~e
cos$ = (4.133)
e
The angle $ takes the value in the range 0 ≤ θ ≤ 360o, and we check the
quadrant as follows.

0o ≤ $ ≤ 180o if ~e · Jˆ = eY ≥ 0 (4.134)
180o < $ ≤ 360o if ~e · Jˆ = eY < 0 (4.135)
190 CHAPTER 4. DYNAMICS I

Circular equatorial orbit. The case n = 0, e = 0 corresponds to


a circular equatorial orbit and we have an alternate orbital element, true
longitude l, the angle measured from the vernal equinox direction Iˆ to the
spacecraft position vector ~r in the direction of spacecraft motion, i.e.,

~r · Iˆ
cosl = (4.136)
r
As the angle l lies in the range 0 ≤ θ ≤ 360o, we must check the quadrant as
follows.

0o ≤ l ≤ 180o if ~r · Jˆ = Y ≥ 0 (4.137)
180o < l ≤ 360o if ~r · Jˆ = Y < 0 (4.138)

4.9 Orbital Perturbations


We have so far in this chapter discussed the relative motion between two bod-
ies only under mutual gravitational force of attraction. Furthermore, the two
bodies are assumed to be point masses or spherically-symmetric mass bodies.
As a consequence of of these assumptions, the relative motion of two bodies
is found to be an ideal or Keplerian orbit with constant angular momentum
~h, energy E, and eccentricity vector ~e resulting in the orbit plane remaining
fixed in inertial space. However, in practical situations, the interacting bod-
ies may have aspherical and asymmetric mass distributions and they may
be subjected to gravitational attraction from other bodies, and other forces.
For example, in the Low-Earth orbit system, the satellite can be assumed
as point mass body but considering the Earth as spherically-symmetric mass
body may not be correct; the Earth is actually an ellipsoidal shape. In ad-
dition to these, the satellite experiences several forces including atmospheric
force, Earth’s magnetic force, and solar radiation pressure. Results of the
aspherical and asymmetric mass distributions of the interacting bodies and
the presence of nonconservative forces lead to deviations from the ideal Ke-
plerian orbital motion of the two bodies, often called orbital perturbations.
Furthermore, the angular momentum ~h, energy E, and eccentricity vector ~e
no longer remain constant and thereby the orbit plane cannot remain fixed
in inertial space. We shall prove these facts in the following sections. Let us
first begin with the derivation of the equations of motion of the two bodies.

4.9.1 Non-Keplerian Orbit Motion


Applying Newton’s second law of motion as per the procedure outlined in the
Section 4.4, the relative equation of motion of the two bodies in the presence
4.9. ORBITAL PERTURBATIONS 191

of perturbation force f~ can be derived as


µ
~r¨ + 3 ~r = f~ (4.139)
r
where ~r is the position vector of the center of mass of the body m2 with
respect to the center of of mass of the body m1 , ~r¨ is the second derivative of
~r with respect to time, and µ = G(m1 + m2 ).
Now, in order to write the relative equation of motion Eq. (4.139) in
a scaler form, we express the relative vector ~r by its cartesian components
(X,Y ,Z) as

~r = X Iˆ + Y Jˆ + Z K̂

Furthermore, assuming cartesian coordinate axes I, J and K be inertial axes


˙ ˙ ˙
(i.e., Iˆ = Jˆ = K̂ = 0), we can write acceleration ~r¨ by differentiating the
above equation twice

~r¨ = Ẍ Iˆ + Ÿ Jˆ + Z̈ K̂

Substituting this relation in Eq. (4.139), we obtain the relative equation of


motion in the scaler form as
µ
Ẍ = − 3 X + fX (4.140)
r
µ
Ÿ = − 3 Y + fY (4.141)
r
µ
Z̈ = − 3 Z + fZ (4.142)
r
where r = [X 2 + Y 2 + Z 2 ]1/2 . The forces fX , fY , and fZ are the components
of force f~ along I,
ˆ Jˆ, and K̂ axes, respectively.

4.9.2 Angular Momentum Vector


Taking the cross product of Eq. (4.139) with ~r, we have
µ
~r × ~r¨ + 3 (~r × ~r) = ~r × f~ (4.143)
r

Knowing that ~r × ~r=0 and


d  
~r × ~r˙ = ~r˙ × ~r˙ + ~r × ~r¨ = ~r × ~r¨
dt
we can rewrite Eq. (4.143) as
d  
~r × ~r˙ = ~r × f~ (4.144)
dt
192 CHAPTER 4. DYNAMICS I

Knowing ~h = ~r × ~r˙ leads to


d~h
= ~r × f~ (4.145)
dt
Thus, the angular momentum ~h does not remain constant if f~ 6= 0 or ~r×f~ 6= 0.
In other words, the angular momentum is not conserved in the two-body
system under nonconservative forces.

4.9.3 Energy
Taking the dot product of Eq. (4.139) with ~r˙ , we have
µ
~r¨ · ~r˙ + 3 ~r · ~r˙ = f~ · ~r˙ (4.146)
r
or
1 d  ˙ ˙ d µ ~ ˙
~r · ~r − = f · ~r (4.147)
2 dt dt r
Knowing energy E = ~r˙ 2 /2 − µ/r leads to
dE
= f~ · ~r˙ (4.148)
dt
As per the above equation, the energy E does not remain constant if f~ 6= 0
or f~ · ~r˙ 6= 0. Thus, the orbit energy E varies under nonconservative forces.

4.9.4 Eccentricity Vector


Taking the post-cross product of Eq. (4.139) with ~h, we have
µ  
~r¨ × ~h + 3 ~r × ~h = f~ × ~h (4.149)
r
Using the relations
d ˙ ~ ¨ ~ ˙ ˙ × ~h˙
~r × h = ~r × h + ~r˙ × ~h = ~r¨ × ~h +~r
dt
and ~h = ~r × ~r˙ in Eq. (4.149) leads to
d ˙ ~ µ ˙ × ~h˙ + f~ × ~h
~r × h + 3 [~r × (~r × ~r)] =~r (4.150)
dt r
Applying the triple vector product relation: ~a × (~b × ~c) = (~a · ~c)~b − (~a · ~b)~c
yields
d ˙ ~ µ h i
~r × h + 3 (~r · ~r) ~r − (~r · ~r)~r˙ = f~ × ~h (4.151)
dt r
4.9. ORBITAL PERTURBATIONS 193

or
d h ˙ ~ µ i ˙ ~˙
~r × h − ~r =~r × h + f~ × ~h (4.152)
dt r

Using Eq. (4.145) and knowing µ~e = ~r˙ × ~h − µ~r/r as per Eq. (4.32) lead to

d~e ˙
µ = ~r˙ × ~h + f~ × ~h = ~r˙ × (~r × f~) + f~ × ~h (4.153)
dt
The eccentricity ~e does not remain constant if the right-hand size of the above
equation is nonzero.
Thus, we can conclude that in the presence of nonconservative forces the
angular momentum ~h, the energy E, and eccentricity vector ~e of an orbit are
not constant. As a result, it is expected that the orbital elements (i.e., a, e,
i, ω, Ω, and M ) may vary. The time derivatives of the orbital elements or
the equations of motion in terms of orbital elements are derived in the next
section.

4.9.5 Equations of Motion (Orbital Elements)


The specific energy E which is only related to the semimajor axis a as per
the energy Eq. (4.74) can be rewritten as

v2
 
µ µ
E= − =− (4.154)
2 r 2a

Substituting E into Eq. (4.148), we have


 
dE d −µ
= = f~ · ~r˙ (4.155)
dt dt 2a
or
da 2a2  ~ ˙ 
= f · ~r (4.156)
dt µ

In order to obtain (f~ · ~r˙ ), we define ~r and f~ in an cartesian coordinate


frame O − xyz with î, ĵ, k̂ be unit vectors along ~r, the transverse orbit
direction, and the direction normal to the orbit plane, respectively, such that
î × ĵ = k̂ (refer to Fig. (4.13)). We can thus express ~r and f~ as

~r = rî (4.157)
f~ = fx î + fy ĵ + fz k̂ (4.158)
194 CHAPTER 4. DYNAMICS I

Upon substitution of these in Eq. (4.145), we can determine the effect of the
perturbation force on the angular momentum vector ~h as

d~h
= ~r × f~ = rî × (fx î + fy ĵ + fz k̂)
dt
= rfy k̂ − rfz ĵ (4.159)
If the component of the perturbation force normal to the orbit plane is zero
(i.e., fz = 0), then

d~h
= rfy k̂ (4.160)
dt
Thus, if fz =0 the angular momentum vector ~h initial direction along k̂
(i.e., ~h = r2 θ̇ k̂) remains unchanged but its magnitude changes. In other
words, the orbit plane remains constant if fz =0. In case fy = fz = 0 then ~h
remains unchanged. So, the force along î (i.e., fx ) does not affect the angular
momentum.
Differentiating Eq.(4.157) and noting that î, ĵ, k̂ are non-inertial axes
˙
(i.e., î˙ 6= 0, ĵ˙ 6= 0, and k̂ 6= 0), we have the velocity as

~v = ~r˙ = ṙ î + rî˙ = ṙ î + r(~


ω × î) (4.161)
where ω
~ can be derived considering the transformation from the inertial co-
ordinate frame O − XY Z to the rotating coordinate frame O − xyz by Ω̇
about the K̂ axis, followed by i̇ about the nodal în axis, and finally (ω̇ + θ̇)
about the k̂ axis. Thus, ω
~ can be written as
ω
~ = Ω̇K̂ + i̇în + (ω̇ + θ̇)k̂ (4.162)
Applying the coordinate transformations, we can express ω
~ in terms of the
rotating coordinate components xyz
ω
~ = ωx î + ωy ĵ + ωz k̂ (4.163)
where
   
 ω   Ω̇sinisin(ω + θ) + i̇cos(ω + θ) 
 x

 
 
 

  
ωy = Ω̇sinicos(ω + θ) − i̇sin(ω + θ) (4.164)

 
 
 

 ω
 
  
Ω̇cosi + θ̇ + ω̇


z

Using the above relations in Eq. (4.165), we write ~v as


h i h i
~v = ṙî + r (ωx î + ωy ĵ + ωz k̂) × î = ṙ î + r −ωy k̂ + ωz ĵ (4.165)
(4.166)
4.9. ORBITAL PERTURBATIONS 195

Alternatively, we rewrite ~v as

~v = vx î + vy ĵ + vz k̂ (4.167)

where
   
 v   ṙ 
 x

 
 
 

  
vy = r[Ω̇cosi + θ̇ + ω̇] (4.168)

 
 
 

 vz 
    −r[Ω̇sinicos(ω + θ) − i̇sin(ω + θ)] 

We now assume that that ~r and ~r˙ lie in the orbit plane. This orbit plane is
called the osculating orbital plane. With this assumption, the components of
~v simplify to

µ
r
vx ≡ ṙ = esinθ (4.169)
p
µ
r
vy ≡ rθ̇ = (1 + ecosθ) (4.170)
p
vz = 0 (4.171)

Note that the assumption of osculating orbital plane will lead to approximate
relative equations of the two-body motion in terms of orbital elements.
Using Eqs. (4.167) and (4.158) along with the osculating orbit assump-
tion, the term f~ · ~r˙ can be written as

f~ · ~r˙ = fx vx + fy vy (4.172)

Substituting the above relation into Eq. (4.156) leads to

da 2a2
= √ [fx esinθ + fy (1 + ecosθ)] (4.173)
dt µp

The above equation states the time derivative of the semimajor axis. In the
case perturbation force components fx and fy are zero, the semimajor axis
remains constant. Thus, the semimajor axis is only affected by the force
components in the orbit plane. Next we derive the variational equations of
other orbital elements.
The angular momentum ~h = ~r × ~r˙ implies that ~h is perpendicular to ~r
and ~r˙ . Considering the osculating orbit, we can write ~h as
~h = hk̂ (4.174)

which is normal to the orbit plane containing ~r and ~r˙ .


196 CHAPTER 4. DYNAMICS I

Knowing h2 /µ = p, we have
~h = √µpk̂ (4.175)

Differentiating this equation leads to

~h˙ = 1 µ ṗk̂ + √µpk̂˙


r
(4.176)
2 p
˙
Knowing k̂ = ω
~ × k̂ and using Eq. (4.163) for ω
~ , we have

~h˙ = 1 µ ṗk̂ + √µp(−ωx ĵ + ωy î)


r
(4.177)
2 p
˙ ˙
Since ~h = ~r × f~, we can write ~h using Eqs. (4.157) and (4.158) as

~h˙ = ~r × f~ = rî × fx î + fy ĵ + fz k̂ = rfy k̂ − rfz ĵ


 
(4.178)

Equating the coefficients of Eqs. (4.177) and (4.178) gives

1 µ
r
ṗ = rfy (4.179)
2 p

− µp[Ω̇sinisin(ω + θ) + i̇cos(ω + θ)] = −rfz (4.180)

µp[Ω̇sinicos(ω + θ) − i̇sin(ω + θ)] = 0 (4.181)

or we simplify and rewrite as


dp p
r
=2 rfy (4.182)
dt µ
dΩ rfz
= √ sin(ω + θ) (4.183)
dt sini µp
di rfz
= √ cos(ω + θ) (4.184)
dt µp

The above Eqs. (4.183) and (4.184) show the time derivatives of longitude
of ascending node Ω and orbit inclination i, respectively.
Differentiating the relation, p = a(1 − e2 ), yields
dp da de
= (1 − e2 ) − 2ae (4.185)
dt dt dt
We write in term of (de/dt) as
 
de 1 da dp
= (1 − e2 ) − (4.186)
dt 2ea dt dt
4.9. ORBITAL PERTURBATIONS 197

Substituting (da/dt) and (dp/dt) from Eqs. (4.173) and (4.182), respectively
into the above equation and using the following relationships

p = a(1 − e2 )
r = a(1 − ecosE)
e + cosθ
cosE =
1 + ecosθ
lead to
  
de p e + cosθ
r
= fx sinθ + fy cosθ + (4.187)
dt µ 1 + ecosθ

Next we derive the time derivative of ω. Considering eccentricity vector


~e = eîe and differentiating, yields

d~e de dîe de
= îe + e = îe + e(~
ωe × îe ) (4.188)
dt dt dt dt
where ω~ e can be obtained considering the transformation as followed for
ω
~ in Eq. (4.162). Considering Eq. (4.153) of ~e in conjunction with Eqs.
(4.157),(4.157), and (4.187), we have
r    
1 p r
ω̇ + Ω̇cosi = −fx cosθ + fy 1 + sinθ (4.189)
e µ p
In term of dω/dt as
r    
dω 1 p r dΩ
= −fx cosθ + fy 1 + sinθ − cosi (4.190)
dt e µ p dt

In order to obtain the time derivative of the mean anomaly M , we consider


Kepler’s time equation

M = E − esinE (4.191)

Differentiating it yields
dM dE de dE
= − sinE − ecosE (4.192)
dt dt dt dt
The (dE/dt) can be found by differentiating r = a(1 − ecosE), as
dr da de dE
= (1 − ecosE) − acosE + aesinE (4.193)
dt dt dt dt
or
 
dE 1 dr da de
= − (1 − ecosE) + acosE (4.194)
dt aesinE dt dt dt
198 CHAPTER 4. DYNAMICS I

Substituting the above relation for (dE/dt) into Eq. (4.192) and using ex-
pressions for (da/dt) and (de/dt) from Eqs. (4.173) and (4.187), respectively,
as well as applying the following relationships
dr µ
r
= esinθ
dt p
e + cosθ
cosE =
1 + ecosθ

1 − e2 sinθ
sinE =
1 + ecosθ
we obtain (dM/dt) as

1 − e2
   
dM 2rfx r
=n− + f x cosθ − f y 1 + sinθ (4.195)
dt na2 nae p

In summary, we can write the equations of motion in terms of orbital


elements as
da 2
= [fx esinθ + fy (1 + ecosθ)]
dt n(1 − e2 )1/2
de (1 − e2 )1/2
  
e + cosθ
= fx sinθ + fy cosθ +
dt na 1 + ecosθ
2 1/2
   
dω −(1 − e ) r
= fx cosθ − fy 1 + sinθ
dt nae a(1 − e2 )
fz rcosisin(ω + θ)
+ 2 (4.196)
na (1 − e2 )1/2 sini
dΩ fz rsin(ω + θ)
= 2
dt na (1 − e2 )1/2 sini
di fz rcos(ω + θ)
=
dt na2 (1 − e2 )1/2
1 − e2
   
dM r 2rfx
=n + fx cosθ − fy 1 + 2
sinθ −
dt nae a(1 − e ) na2

where
a(1 − e2 )
r
µ
n= , a = semi-major axis, r= = a(1 − ecosE)
a3 1 + ecosθ
The above Eqs. (4.196) are called Gauss form of Lagrange’s Planetary Equa-
tions. Note that when e=0 and / or sini=0, this set of equations has a
singularity problem. Therefore, we consider another set of equations that are
free of singularities using a set of the so-called equinoctial elements attributed
to Professor Roger A. Broucke[1]. Again, this set comprises of six elements:
4.9. ORBITAL PERTURBATIONS 199

semimajor axis a and five other elements (P1 , P2 , Q1 , Q2 , l) that are defined
in terms of the classical orbital elements as

P1 = esin(ω + Ω)
P2 = ecos(ω + Ω)
Q1 = tan(i/2)sinΩ (4.197)
Q2 = tan(i/2)cosΩ
l =Ω+ω+M

where the element l is called the mean longitude. We further define the true
longitude L and eccentric longitude K as

L=Ω+ω+θ (4.198)
K =Ω+ω+E (4.199)

Using Kepler’s orbit equation, r = a(1 − cosE) and Kepler’s time equation,
M = E − esinE, we can express the position r and mean longitude l in terms
of P1 , P2 , and K as

r = a(1 − P1 sinK − P2 cosK) (4.200)


l = K + P1 cosK − P2 sinK (4.201)

Now we write Gauss’ variational equations in terms of the equinoctial


elements as (Refer to Battin [2] for a complete derivation)

da 2a2
= [(P2 sinL − P1 cosL)fx + gfy ]
dt nab
dP1
=q {(−gcosL)fx + [P1 + (1 + g)sinL] fy − P2 (Q1 cosL − Q2 sinL)fz }
dt
dP2 n o
=q (gsinL)fx + [P2 + (1 + g)cosL] fy + P1 (Q1 cosL − Q2 sinL)fz
dt
dQ1 q
= (1 + Q21 + Q22 )fz sinL
dt 2
dQ2 q
= (1 + Q21 + Q22 )fz cosL
dt 2 ( 
dl ag 2b
=n − q (P1 sinL + P2 cosL) + fx
dt a+b a
)
a
+ (1 + g)(P1 cosL − P2 sinL)fy + (Q1 cosL − Q2 sinL)fz
a+b
(4.202)
200 CHAPTER 4. DYNAMICS I

where
p q
b = semi-minor axis = a 1 − e2 = a 1 − P12 − P22
p
g = = 1 + P1 sinL + P2 cosL
r
r r nab
q= = =
h nab µ(1 + P1 sinL + P2 cosL)

Here in order to find L for a given equinoctial elements, we first determine


K by solving

l = K + P1 cosK − P2 sinK

and then find L using the relations (that can be derived)


  
a a 2 a
sinL = 1− P sinK + P1 P2 cosK − P1
r a+b 2 a+b
  
a a 2 a
cosL = 1− P cosK + P1 P2 sinK − P2 (4.203)
r a+b 1 a+b

where
a 1 1
= p = q
a+b 1+ 1−e 2
1 + 1 − P12 − P22
r = a(1 − P1 sinK − P2 cosK)

The Earth orbiting satellite system is under the influence of perturbation


forces due to the Earth’s oblateness and triaxiality, solar radiation pressure,
gravitational forces from the Sun and Moon, and aerodynamic drag. In the
next section, we study Earth’s oblateness, gravitational forces from the Sun
and Moon, and aerodynamic drag.
While the orbital elements discussed at the beginning of this section pro-
vide an excellent reference for describing orbits, there are other forces acting
on a satellite that perturb it away from the nominal orbit. These perturba-
tions, or variations in the orbital elements, can be classified based on how
they affect the Keplerian elements. Secular variations represent a linear vari-
ation in the element, short-period variations are periodic in the element with
a period less than the orbital period, and long-period variations are those
with a period greater than the orbital period. Because secular variations
have long-term effects on orbit prediction (the orbital elements affected con-
tinue to increase or decrease), they will be discussed here for Earth-orbiting
satellites. Precise orbit determination requires that the periodic variations
be included as well.
4.9. ORBITAL PERTURBATIONS 201

4.9.6 Earth’s Oblateness


The potential energy U of a satellite (point mass, m) is given by

GmdMe
Z Z
U= dU = − (4.204)
Me Me ~
|R|

~ = ~r − ρ
where R ~. ~r is the position vector of the satellite with respect to the
Earth center of mass while ρ is the position vector of the element of mass
dMe .
~ given by
Substituting |R|

~ = |~r − ρ
|R| ~| = [r2 − 2~r.~ ~2 ]1/2
ρ+ρ (4.205)

into Eq.(4.204) yields

Gm
Z
U = − dMe
Me (r2
− 2~r · ρ
~+ρ ~2 )1/2
−1/2
~ ρ2
Z 
Gm 2~r · ρ
= − 1− 2 + 2 dMe (4.206)
r Me r r

Applying Binomial series expansion,

∞  
X n n
(1 + x)n = x if |x| < 1 (4.207)
k
k=0

where,
 
n n n(n − 1) · · · (n − k + 1)
= = (4.208)
k k!(n − k)! 1 · 2···k

for the term inside the bracket, we get

~ 1 ρ2 ~)2 3 ρ4
 
Gm ~r · ρ 3 (~r · ρ 3 (~r.~
ρ) 2
Z
U =− 1+ 2 − + + − ρ + · · · dMe
r Me r 2 r2 2 r4 8 r4 2 r4
(4.209)

Now, considering the fact ρ  r, and the carrying out expansion upto
202 CHAPTER 4. DYNAMICS I

O(1/r3 ), we get

~ 1 ρ2 3 (~r · ρ~)2
 
Gm ~r · ρ 1
Z
U = − 1+ 2 − + + O( 3 ) dMe
r Me r 2 r2 2 r4 r
"
Gm 1 1
Z Z
= − Me + 2 (~r · ρ
~)dMe − 2 ρ2 dMe
r r Me 2r Me
#
3
Z
2
+ 4 (~r · ρ
~) dMe (4.210)
2r Me

Here ~r and ρ
~ are defined with respect to the reference frame S − xo yo zo
and the Earth-body fixed reference frame S − xyz, respectively as

~r = rîo = r(Ci î + Cj ĵ + Ck k̂) (4.211)

ρ
~ = xî + y ĵ + z k̂ (4.212)

Substituting ~r and ρ
~ into Eq.(4.210), we get

(
µ 1
Z
U =− Me + [x(îo · î) + y(îo · ĵ) + z(îo · k̂)]dMe
r r Me
3
Z
+ 2 [x2 (îo · î)2 + y 2 (îo · ĵ)2 + z 2 (îo · k̂)2
2r Me
)
+ 2xy(îo · î)(îo · ĵ) + 2xz(îo · î)(îo · k̂) + 2yz(îo · ĵ)(îo · k̂)]dMe

(4.213)

As the center of mass lies at E, we get

Z
ρ~dMe = 0 (4.214)
Me

i.e.,
Z Z Z
xdMe = ydMe = zdMe = 0 (4.215)
Me Me Me
4.9. ORBITAL PERTURBATIONS 203

We have the following relations for the moment of inertia of the the Earth
as

Z Z
Ixx = (y 2 + z 2 )dMe ; Iyy = (z 2 + x2 )dMe ;
Me Me
Z Z
Izz = (x2 + y 2 )dMe ; Ixy = xydMe ;
Me Me
Z Z
Iyz = yzdMe ; Ixz = xzdMe (4.216)
Me Me

From Eq. (4.216), we have

Iyy + Izz − Ixx Izz + Ixx − Iyy


Z Z
x2 dMe = ; y 2 dMe = ;
Me 2 Me 2
Ixx + Iyy − Izz
Z
z 2 dMe = (4.217)
Me 2

Substituting Eqs. (4.214-4.217) into Eq. (4.213) yields

(
GmMe Gm
U =− − 3 (Iyy + Izz − Ixx )[3(îo · î)2 − 1]
r 4r
+ (Izz + Ixx − Iyy )[3(îo · ĵ)2 − 1]
+ (Ixx + Iyy − Izz )[3(îo · k̂)2 − 1]
+ 12Ixy (îo · î)(îo · ĵ) + 12Ixz (îo · î)(îo · k̂)
)
+ 12Iyz (îo · ĵ)(îo · k̂) (4.218)

In the case of the axes x, y, and z to be principal axes, then the product
of inertia terms vanish, i.e., (Ixy = Ixz = Iyz = 0). Denoting (îo · î), (îo · ĵ),
and (îo · k̂) as direction cosines Ci , Cj , and Ck , respectively, the potential
energy of the spacecraft U simplifies to

(
GmMe Gm
U =− − 3 (Iyy + Izz − Ixx )(3Ci2 − 1)
r 4r
+ (Izz + Ixx − Iyy )(3Cj2 − 1)
)
+ (Ixx + Iyy − Izz )(3Ck2 − 1) (4.219)
204 CHAPTER 4. DYNAMICS I

We consider only Earth’s oblateness effect and so assuming the Earth


is symmetric about Y and Z-axes leads to Ixx = Iyy = Iy and Izz = Iz .
Knowing Ci2 + Cj2 + Ck2 = 1 and Ck = Z/r, we have

( )
GMe m Gm(Iy − Iz ) Z2
U =− − 3 2 −1 (4.220)
r 2r3 r

Iy − Iz
Writing = J2 , called J2 perturbation , the preceding equation sim-
Me R 2
plifies to

1 J2 Re2
  2 
GMe m Z
U =− 1− 3 − 1 (4.221)
r 2 r2 r2

The preceding equation is the particular case of the general expression of


the potential energy of the Earth-satellite system (Refer Eq. 3.7):
" ∞  n #
GMe m X Re
U =− 1− Jn Pn (sinλ) (4.222)
r n=2
r

where Pn (sinλ) denote Legendre polynomials and Jn represent dimensionless


geopotential coefficients or zonal coefficients and they are infact constant re-
lated to the inertia integrals. The first few coefficients are J2 =1082.63×10−6,
J3 =-2.54×10−6, J4 =-1.61×10−6. The nomenclature λ denotes latitude of the
satellite. For n = 1, the term in the series correspond to the assumption that
the origin coincides with the Earth’s center of mass. For n = 2,

1 J2 Re2
  2 
µm Z
U =− 1− 3 2 −1 (4.223)
r 2 r2 r

where µ = GME , Z = rsinλ, P2 (sinλ) = (3sin2 λ − 1)/2, and sinλ =


sini sin(ω + θ).
The force per unit mass due to the Earth’s gravitational field is given by
 
~ 1 1 ∂U ∂U
fg = − ∇U = − î + K̂
m m ∂r ∂Z
3 J2 Re2 5Z 2
  
µ 2Z
= − 2 î + − 2 + 1 î + K̂ (4.224)
r 2 r4 r r

The equations of motion of a satellite in the geocentric inertial coordinate


4.9. ORBITAL PERTURBATIONS 205

frame O − XY Z can be written as

3µJ2 Re2 5Z 2
 
µ
Ẍ = − 3 X − 1− 2 X (4.225)
r 2r5 r
2
5Z 2
 
µ 3µJ2 Re
Ÿ = − 3 Y − 1− 2 Y (4.226)
r 2r5 r
2
5Z 2
 
µ 3µJ2 Re
Z̈ = − 3 Z − 3− 2 Z (4.227)
r 2r5 r

where r = [X 2 + Y 2 + Z 2 ]1/2 , and µ = GME . J2 and Re are second zonal


harmonic of the Earth’s gravitational field and the Earth radius.
We next study the effect of J2 perturbation on the orbital elements.
The Lagrange planetary equations of motion derived in earlier section are
used here and for this we should have the force components in î, ĵ, and
k̂ directions. Applying the transformation from the Earth-centered inertial
frame O − IJK to the orbital fixed frame O − ijk using rotation sequence
R3 (ω + θ)R1 (i)R3 (Ω), we get

   


 î 




 Iˆ 


   
ĵ = R313 (Ω, i, ω + θ) Jˆ (4.228)

 
 
 

 k̂ 
  
 K̂ 

where

 
cΩc(ω + θ) − sΩcis(ω + θ) sΩc(ω + θ) + cΩcis(ω + θ) sis(ω + θ)
 
R313 (Ω, i, ω + θ) = −cΩs(ω + θ) − sΩcic(ω + θ) −sΩs(ω + θ) + cΩcic(ω + θ) sic(ω + θ)
 
 
sΩsi −cΩsi ci
(4.229)

where cΩ stands for cosω, sΩ stands for sinω, and similarly for other param-
eters.
Thus, using

   


 ˆ
I 
 

 î 


   
Jˆ −1
= R313 ĵ (4.230)

 
 
 

 K̂ 
   k̂ 
 
206 CHAPTER 4. DYNAMICS I

we can write

K̂ = sinisin(ω + θ)î + sinicos(ω + θ)ĵ + cosik̂ (4.231)

Let η be the angle between K̂ and î. Then, we have

K̂ · î = cosη = sinisin(ω + θ) (4.232)

and
Z
sinφ = = cosη (4.233)
r
where φ is the latitude of the satellite.
Using Eqs. (4.231-4.233) in Eq. (4.224), the components of perturbation
force due to the Earth’s oblateness (J2 ) along î, ĵ, and k̂ directions can be
written as

3 µJ2 Re2
1 − 3sin2 isin2 (ω + θ)
 
fx = − (4.234)
2 r4
3 µJ2 Re2  2 
fy = − sin isin2(ω + θ) (4.235)
2 r4
3 µJ2 Re2
fz = − [sin2isin(ω + θ)] (4.236)
2 r4
Upon substitution of the perturbation force due to J2 in the so called
Gauss form of the Lagrange planetary equations of motion, we get the re-
sulting equations of motion as

(
da 3µJ2 Re2
1 − 3sin2 isin2 (ω + θ) esinθ
 
=− 4
dt nr (1 − e2 )1/2
)
+ (1 + ecosθ)sin2 isin2(ω + θ)
(
de 3µ(1 − e2 )1/2 J2 Re2 
1 − 3sin2 isin2 (ω + θ) sinθ

= 4
dt 2nar
  )
e + cosθ 2
+ cosθ + sin isin2(ω + θ) (4.237)
1 + ecosθ
di 3µJ2 Re2 sin2isin(ω + θ)cos(ω + θ)
=−
dt 2na2 (1 − e2 )1/2 r3
4.9. ORBITAL PERTURBATIONS 207

(
dω 3µ(1 − e2 )1/2 J2 Re2 
1 − 3sin2 isin2 (ω + θ) cosθ

=− 4
dt 2naer
  )
r 2
− 1+ sin isin2(ω + θ)sinθ
a(1 − e2 )
3µJ2 Re2 cos2 i
− sin2 (ω + θ) (4.238)
na2 (1 − e2 )1/2 r3
dΩ 3µJ2 Re2 cosisin2 (ω + θ)
=−
dt na2 (1 − e2 )1/2 r3
(
dM 3µ(1 − e2 )J2 Re2 
1 − 3sin2 isin2 (ω + θ) cosθ

=n −
dt 2naer4
  )
r 2
− 1+ sin isin2(ω + θ)sinθ
a(1 − e2 )
3µJ2 Re2 
1 − 3sin2 isin2 (ω + θ)

+ 2 3
na r

To obtain the average values of the orbital elements in one orbit of time
period T = 2π/n = 2πa3/2 /µ1/2 , we consider the following relation

T
1
Z
Ẋavg = Ẋdt, X = a, e, i, ω, Ω, M (4.239)
T t=0

As Ẋ depends upon θ or ω + θ, the variable time t is required to be


changed to θ or ω + θ and the corresponding limits of integration will get
changed to 0 to 2π. The orbital angular momentum h is given by

h = r2 ωz = r2 [Ω̇cosi + θ̇ + ω̇] (4.240)

Ignoring the second order term Ω̇cosi, Eq. (4.240) simplifies to

h = r2 (θ̇ + ω̇) (4.241)



Substituting h = µp, we can rewrite Eq. (4.241) as

r2
dt = √ dθ1 (4.242)
µp
208 CHAPTER 4. DYNAMICS I

Here θ1 = ω + θ, p=semi-latus rectum=a(1 − e2 ). Using the relation


(4.242) in Eq. (4.239) and replacing r by the relation r = p/(1 + ecosθ), we
express the orbit-average values of orbital elements as


3µ1/2 J2 Re2
 
da
Z n
1 − 3sin2 isin2 θ1 [1 + ecos(θ1 − ω)]2

=−
dt avg np a1/2 (1 − e2 )T
2
0
o
× esin(θ1 − ω) + [1 + ecos(θ1 − ω)]3 sin2 isin2θ1 dθ1
3J2 Re2 2π n 
 
de
Z
1 − 3sin2 isin2 θ1 [1 + ecos(θ1 − ω)]2

= 2
dt avg 2p T 0
 
e + cos(θ1 − ω)
× sin(θ1 − ω) + cos(θ1 − ω) +
1 + ecos(θ1 − ω)
o
2 2
× [1 + ecos(θ1 − ω)] sin isin2θ1 dθ1
3J2 Re2 sin2i 2π
 
di
Z
=− [1 + ecos(θ1 − ω)]sinθ1 cosθ1 dθ1
dt avg 2p2 T 0

3J2 Re2 2π n 
 

Z
1 − 3sin2 isin2 θ1 [1 + ecos(θ1 − ω)]2 cos(θ1 − ω)

=−
dt avg 2ep2 T 0
 
r
− 1+ [1 + ecos(θ1 − ω)]2 sin2 isin2θ1 sin(θ1 − ω)
a(1 − e2 )
recos2 i 2 2
o
+ [1 + ecos(θ 1 − ω)] sin θ 1 dθ1 (4.243)
a(1 − e2 )

3J2 Re2 cosi 2π


 
dΩ
Z
=− [1 + ecos(θ1 − ω)]sin2 θ1 dθ1
dt avg p2 T 0

1 T 3(1 − e2 )1/2 J2 Re2


 
dM
Z
= ndt −
dt avg T 0 2ep2 T
Z 2π (
1 − 3sin2 isin2 θ1 [1 + ecos(θ1 − ω)]2 cos(θ1 − ω)
 
×
0
 
r
− 1+ [1 + ecos(θ1 − ω)]2 sin2 isin2θ1 sin(θ1 − ω)
a(1 − e2 )
)
2pe  2 2

+ 1 − 3sin isin θ1 [1 + ecos(θ1 − ω)] dθ1
a(1 − e2 )

In order to integrate the preceding equations, we assume that there are


no changes in orbital elements in one orbit to a first-order approximation.
We apply the following relations for the integrals:
4.9. ORBITAL PERTURBATIONS 209

Z 2π
[1 + ecos(θ1 − ω)]sin2θ1 dθ1 = 0
0
Z 2π
[1 + ecos(θ1 − ω)]cos2θ1 dθ1 = 0
0
Z 2π
[1 + ecos(θ1 − ω)]sin2 θ1 dθ1 = π
0
Z 2π
sin(θ1 − ω)sin2θ1 dθ1 = 0 (4.244)
0
Z 2π
sin(θ1 − ω)sin2 θ1 dθ1 = 0
0
Z 2π
cos(θ1 − ω)sin2θ1 dθ1 = 0
0
Z 2π
cos(θ1 − ω)sin2 θ1 dθ1 = 0
0

As an example, the derivation for two integrals of the above equations are
shown:
Z 2π Z 2π
[1 + ecos(θ1 − ω)]sinθ1 cosθ1 dθ1 = cosθ1 dθ1 + ecosω
0 0
Z 2π Z 2π
× sinθ1 cos2 θ1 dθ1 + esinω sin2 θ1 cosθ1 dθ1
0 0
=0+0+0=0 (4.245)

2π 2π
1 − cos2θ1
Z Z
2
[1 + ecos(θ1 − ω)]sin θ1 dθ1 = dθ1 + ecosω
0 0 2
Z 2π Z 2π
2
× sin θ1 cosθ1 dθ1 + esinω sin3 θ1 dθ1
0 0
=π+0+0=π (4.246)

Furthermore, we also considered the following integral:


Z 2π
[1 − 3sin2 isin2 θ1 ][1 + ecos(θ1 − ω)]2
0
Z 2π
cos(θ1 − ω) − 3sin2 isin2 θ1 cos(θ − ω)
 
× cos(θ1 − ω)dθ1 =
0
× [1 + 2ecos(θ1 − ω) + e2 cos2 (θ1 − ω)]dθ1
= eπ(2 − 3sin2 i) (4.247)
210 CHAPTER 4. DYNAMICS I

Thus, the resulting orbit-averaged values of orbital elements are

 
da
=0
dt avg
 
de
=0
dt avg
 
di
=0
dt avg
3nJ2 Re2
 

= (5cos2 i − 1) (4.248)
dt avg 4p2
3nJ2 Re2 cosi
 
dΩ
=−
dt avg 2p2
3nJ2 (1 − e2 )1/2 Re2
 
dM
=n+ (3cos2 i − 1)
dt avg 4p2
As per the preceding equations, the semimajor axis a, orbit eccentricity e,
and orbit inclination i remain constant; but, the argument of perigee ω,
the longitude of the ascending node Ω, and the mean anomaly M vary in
the presence of Earth oblateness (i.e., , J2 perturbation). It is interesting to
determine the situation when the mean argument of perigee remains constant
(i.e., (dω/dt)avg =0) or the mean longitude of the ascending node does not
vary (i.e., (dΩ/dt)avg = 0). For the case of (dω/dt)avg = 0,

3nJ2 Re2
 

= (5cos2 i − 1) = 0 (4.249)
dt avg 4p2
or
5cos2 i − 1 = 0
r
1
⇒ cosi = ±
5
i = 63.4o or 116o (4.250)
o o
Thus, at the orbit inclination of i=63.4 or 116 , the eccentricity vector ~e
or perigee or major axis remains fixed under the J2 perturbation. In fact,
many satellites are deployed with this particular orbit inclination. These
include satellites in Molniya orbits (typical perigee altitude = 200∼ 1000
km; eccentricity ≈ 0.75; period ≈ 12 hrs); the satellites remain above the
Northern Hemisphere near apogee for ∼ 11 hrs.
In the case (dΩ/dt)avg = 0, we have
cosi = 0 ⇒ i = 90o (4.251)
4.9. ORBITAL PERTURBATIONS 211

Thus, for the satellite in a polar orbit (i.e., i = 90o ) the nodal line ~n remains
constant under J2 perturbation.
Some qualitative results include the following.
1. The line of nodes regresses unless i = 90o , so the orbit rotates back-
wards in inertial space.

2. For all i < sin−1 ( 5/2) = 63.4o , the line of apsides advances, so the
periapsis rotates forward around the orbit. Together with (3), this makes
possible the orbits of Sunsynchronous satellites, such as Landsat, around
Earth; the satellite orbit plane remains approximately fixed with respect to
the Sun i.e., Ω̇avg is equal to the Earth rotation rate around the Sun (0.9856
deg/day). Note if Ω̇avg > 0 the sun synchronous orbit must be retrograde
and its inclination can be found for a given orbital semi-major axis and
eccentricity (i.e., a and e).
p
3. For all i < sin−1 ( 2/3) = 54.7o, the mean anomaly increases at a rate
greater than n, the Keplerian mean motion. Therefore, the orbital period is
less than that of a Keplerian orbit. This implies that one must be careful in
calculating planetary masses from observed orbital elements of satellites.

4.9.7 Planetary Gravitational Perturbations


We examine here the effect of the gravitational forces of the planetary bodies,
specifically the Sun and the Moon, on the orbital motion of the satellite with
respect to the Earth. To begin, we consider a planetary body in the Earth-
satellite system (Fig. 4.14). The orbital equations of motion of the satellite
and the Earth can be written as

mR ~¨ = F~e + F~ps = − GmMe ~r + Gmmp ~rps (4.252)


r3 3
rps
~¨ e = −F~e + F~p = GMe m ~r + GMe mp ~rp
Me R (4.253)
r3 rp3
where ~rps = ~rp − ~r. The constant G is the universal gravitational constant
and Me , m, and mp are the masses of the Earth, satellite, and the planetary
body, respectively.
Using the preceding equations, the relative equation of motion of the
satellite with respect to the Earth is
 
¨

¨ ~¨
~
 Gm(Me + m) ~rp − ~r ~rp
m~r = m R − Re = − ~r + Gmmp − 3 (4.254)
r3 |~rp − ~r|3 rp
or
!
G(Me + m) ~rp − ~r ~rp
~r¨ + ~r = µp − 3 (4.255)
r3 [rp2 − 2(~rp · ~r) + r2 ]3/2 rp
212 CHAPTER 4. DYNAMICS I

Planet
Inertial Fps
Reference
Frame rp rps Satellite
O Fp
Fe
R Satellite
−Fe
Re r
Earth
Earth Free−Body Diagrams

Figure 4.14: Planetary Gravitational Perturbation.

where µp = Gmp is the gravitational parameter of the planetary body. The


right hand side of the preceding equation represents the perturbing force,
denoted by f~p , to the satellite-Earth two body system. Considering the fact
r  rp , we can approximate f~p by carrying out Binomial series expansion
upto O(1/rp2 ) as
( )
2 −3/2
 
~
rp − ~
r (~
rp · ~
r ) r ~
rp
f~p = µp 1−2 + 2 − 3
rp3 rp2 rp rp
   
µp (~rp · ~r) 1
= 3 (~rp − ~r) 1 + 3 + O( ) − ~
rp (4.256)
rp rp2 rp2
or
 
µp (~rp · ~r)
f~p = 3 −~r + 3 ~rp (4.257)
rp rp2

Writing ~r and ~rp in terms of unit vectors as ~r = rî and ~rp = rp îp leads to
µp h i
f~p = 3 r 3(î · îp )îp − î (4.258)
rp

Now we express the perturbation force f~p along x, y, and z-coordinates as


µp r h i
fx = f~p · î = 3 3(î · îp )(îp · î) − 1 (4.259)
rp
3µp r
fy = f~p · ĵ = 3 (î · îp )(îp · ĵ) (4.260)
rp
3µp r
fz = f~p · k̂ = 3 (î · îp )(îp · k̂) (4.261)
rp
4.9. ORBITAL PERTURBATIONS 213

To write unit vector îp in terms of the satellite orbit fixed frame O − îĵ k̂, we
first apply the transformation from the Earth-centered inertial frame O−IˆJˆK̂
to the perturbing orbit fixed frame O − îp ĵp k̂p using the rotation sequence
R3 (ωp + θp )R1 (ip )R3 (Ωp ) as

   
 î   Iˆ 
 p

 
 
 

  
ĵp = R313 (Ωp , ip , ωp + θp ) Jˆ (4.262)

 
 
 

k̂p   K̂ 

   

and next we consider the transformation from the satellite orbit fixed frame
O − îĵ k̂ to the Earth-centered inertial frame O − IˆJˆK̂ using rotation sequence
R3 (Ω)R1 (i)R3 (ω + θ) as

   


 Iˆ 




 î 


   
Jˆ = R313 (ω + θ, i, Ω) = R −1
313 (Ω, i, ω + θ) ĵ (4.263)

 
 
 

 K̂ 
   k̂ 
 

where

 
cΩc(ω + θ) − sΩcis(ω + θ) sΩc(ω + θ) + cΩcis(ω + θ) sis(ω + θ)
 
R313 (Ω, i, ω + θ) = −cΩs(ω + θ) − sΩcic(ω + θ) −sΩs(ω + θ) + cΩcic(ω + θ) sic(ω + θ)
 
 
sΩsi −cΩsi ci
(4.264)

where cΩ stands for cosω, sΩ stands for sinω, and similarly for other param-
eters.
214 CHAPTER 4. DYNAMICS I

The vector îp can be expressed as


 T  
cΩp c(ωp + θp ) − sΩp cip s(ωp + θp ) 

 î 


   
−1
îp =  sΩp c(ωp + θp ) + cΩp cip s(ωp + θp )  R313 (Ω, i, ω + θ) ĵ
 
  
 

sip s(ωp + θp )  k̂ 
 
  
[cΩp c(ωp + θp ) − sΩp cip s(ωp + θp )][cΩc(ω + θ) − sΩcis(ω + θ)] 





   

 +[sΩp c(ωp + θp ) + cΩp cip s(ωp + θp )][sΩc(ω + θ) + cΩcis(ω + θ)]   î
   

 

   

+sip s(ωp + θp )sis(ω + θ)
   

   

   

 [cΩp c(ωp + θp ) − sΩp cip s(ωp + θp )][−cΩs(ω + θ) − sΩcic(ω + θ)]  
   

 

  
=  +[sΩp c(ωp + θp ) + cΩp cip s(ωp + θp )][−sΩs(ω + θ) + cΩcic(ω + θ)]  ĵ
 
  
 
+sip s(ωp + θp )sic(ω + θ)
  
  




   

[cΩp c(ωp + θp ) − sΩp cip s(ωp + θp )]sΩsi
   

   

   

−[sΩp c(ωp + θp ) + cΩp cip s(ωp + θp )]cΩsi k̂
   

   

   

+sip s(ωp + θp )ci

 

(4.265)

Perturbation due to the Sun.


Let us consider the Sun as a planet and study the effect of its gravitational
force on satellite orbital elements. Replacing the subscript “p” by “s” (for
Sun) in Eq. (4.265) and taking Ωs =0 and assuming θts = ωs + θs , we write
îs as
  
cθts [cΩc(ω + θ) − sΩcis(ω + θ)]  î 

 

  
 
 +cis cθts sΩc(ω + θ) + cΩcis(ω + θ)] + sis sθts sis(ω + θ)  
 
 

 

  
îs =  cθts c[−cΩs(ω + θ) − sΩcic(ω + θ)] ĵ
 

  
 
 +cis cθts [−sΩs(ω + θ) + cΩcic(ω + θ)] + sis sθts sic(ω + θ)  
  


 

  

cθts sΩsi − cis cθts cΩsi + sis sθts ci k̂ 

 

(4.266)

Substituting the preceding expression for îs in Eq. (4.261), we obtain the
perturbing force fz
3µs r n
fz = 3 cθts [cΩc(ω + θ) − sΩcis(ω + θ)] + cis cθts [sΩc(ω + θ)
rs
o n o
+ cΩcis(ω + θ)] + sis sθts sis(ω + θ) × cθts sΩsi − cis cθts cΩsi + sis sθts ci
(4.267)
4.9. ORBITAL PERTURBATIONS 215

Examining the effect of this force fz on the orbit inclination as per the
relation,

di fz rcos(ω + θ)
= (4.268)
dt na2 (1 − e2 )1/2

we have

di r2 3µs n
= cθts [cΩc2 (ω + θ) − sΩcis(ω + θ)c(ω + θ)]
dt na2 (1 − e2 )1/2 rs3
+ cis cosθts [sΩc2 (ω + θ) + cΩcis(ω + θ)c(ω + θ)]
o
+ sis sθts sis(ω + θ)c(ω + θ)
n o
× cθts sΩsi − cis cθts cΩsi + sis sθts ci
(4.269)

In order to obtain the average value of the orbital element in one orbit
of time period T = 2π/n = 2πa3/2 /µ1/2 , we consider the relations dt =
(r2 /h)dθt , r = p/(1 + ecosθ), p=semi-latus rectum=a(1 − e2 ) and θt = ω + θ,
and we obtain
  T
di 1 di
Z
= dt (4.270)
dt avg T t=0 dt

n r4 3µs n
Z
= cθts [cΩc2 θt − sΩcisθt cθt ]
2π 0 h2 rs3
o
+ cis cosθts [sΩc2 θt + cΩcisθt cθt ] + sis sθts sisθt cθt
n o
× cθts sΩsi − cis cθts cΩsi + sis sθts ci dθt (4.271)

or


np4
 
di 1 3µs n
Z
= cθts [cΩc2 θt − sΩcisθt cθt ]
dt avg 2πh2 0 [1 + e cos(θt − ω)]4 rs3
o
+ cis cθts [sΩc2 θt + cΩcisθt cθt ] + sis sθts sisθt cθt
n o
× cθts sΩsi − cis cθts cΩsi + sis sθts ci dθt
(4.272)
216 CHAPTER 4. DYNAMICS I

or
(
np4 3µs 2
 
di
= 2 s is cΩsΩsi + sis cis cisΩ
dt avg h 4rs3
+ c2θts cΩsΩsi + c2 is cΩsΩsi − sΩcisis cis
 
)
+ s2θts [−cip sic2Ω + sis cicΩ] (4.273)

Neglecting the periodic terms associated with sin2θts and cos2θts and
considering only the secular terms yield
np4 3µs  2
 
di 
= 2 3
s is cΩsΩsi + sis cis cisΩ (4.274)
dt avg h 4rs
p √
Knowing n = µ/a3 , p = a2 (1 − e2 ), and h = µp, we get the relation

np4 /h = a2 (1 − e2 )

and substituting it in Eq. (4.274) leads to

a2 3a2 (1 − e2 )7/2 µs  2
 
di 
= s is cΩsΩsi + sis cis cisΩ (4.275)
dt avg h 4hrs3

For a particular case of a satellite in geostationary orbit (i.e., i, e = 0 ), the


preceding relation reduces to
3µs a2
 
di
= sinΩsinis cosis (4.276)
dt avg 4hrs3

Taking geostationary orbit parameters: a=42,164 km, is =23.45o, and h=129,640


km2 /s, and the Sun’s gravitational parameter µs =1,32686×1011 km3 /s2 , we
have
 
di
= 0.269 deg/year for Ω = 90o , -0.269 deg/year for Ω = 270o
dt avg
(4.277)

Perturbation due to the Moon.


In this case of the Moon, we proceed similarly to the previous case of the
Sun. However, Ω for the Moon is not zero and thus, considering this fact
and changing the subscript “p” by “m”, we can derive the average change in
orbit inclination for geostationary satellites taking only secular terms as
3µm a2
 
di
= 3
sin(Ω − Ωm )sinim cosim (4.278)
dt avg 4hrm
4.9. ORBITAL PERTURBATIONS 217

Substituting for geostationary orbit parameters of a=42,164 km is =23.45o,


and h=129,640 km2 /s, and the moon’s gravitational parameter µm =4,902.8
km3 /s2 as well as knowing the variation of the inclination angle of the moon
im from 18.3 deg to 28.6 deg, we have

 
di  0.478 deg/year for Ω = 90o , i = 18.3o, Ω = 0;
m m
= (4.279)
dt avg  0.674 deg/year for Ω = 90o , i = 28.6o, Ω = 0
m m

Luni-Solar Gravitational Perturbations.


Next we study the combined effect of the Moon and the Sun on the orbital
motion of the satellite. The orbital equations of motion of the satellite and
the Earth can be written as
µe µM µS
R̈ = − 3 ~r − 3 ~rms − 3 ~rss (4.280)
r rms rss
µe µM µS
R̈E = 3 ~r + 3 ~rm + 3 ~rs (4.281)
r rm rs
where the nomenclature R ~ and R~ E denote the position vectors of the satellite
and the Earth from the inertial reference frame, respectively. ~r, ~rm and ~rs
represent the position vectors of the satellite, the Moon, and the Sun with
respect to the Earth, respectively. ~rms and ~rss denote the position vectors of
the satellite with respect to the Moon and the Sun, respectively.
Using the preceding equations and knowing ~rm − ~r + ~rms = 0 and ~rss −
~r + ~rs = 0, the relative equation of motion of the satellite with respect to the
Earth can be written as
~r¨ = R ~¨ E = − µe ~r + f~
~¨ − R (4.282)
r3
or
µe
~r¨ + 3 ~r = f~ (4.283)
r
where f~, the perturbing acceleration caused by the luni-solar gravitational
effects on the satellite and Earth, is
   
~ ~rms ~rm ~rss ~rs
f = −µM 3
+ 3 − µS 3
+ 3 (4.284)
rms rm rss rs

For geostationary satellites, solar gravitational perturbation is less than


4×10−6 while lunar gravitational perturbation is less than 9×10−6. The
effect on orbit inclination change can be linearly combined from Eqs. (4.277)
and (4.279) as
3µm r2 3µS r2
 
di
= 3
sin(Ω − Ω m )sini m cosi m + sinΩsinis cosis
dt avg 4hrm 4hrs3
(4.285)
218 CHAPTER 4. DYNAMICS I

and
 
di
= 0.478 to 0.674 deg/year + 0.269 deg/year (4.286)
dt avg

The gravitational forces of the Sun and the Moon also cause secular vari-
ations in the longitude of the ascending node, argument of perigee, and mean
anomaly. For nearly circular orbits, neglecting the variation caused by the
changing orientation of the orbital plane with respect to both the Moon’s
orbital plane and the ecliptic plane, we obtain the secular rates of changes as
Longitude of the ascending node:

dΩm 0.00338cosi
=− (4.287)
dt n

dΩs 0.00154cosi
=− (4.288)
dt n

Argument of perigee:

dωm 0.00169(4 − 5sin2 i)


=− (4.289)
dt n

dωs 0.0007(4 − 5sin2 i)


=− (4.290)
dt n

where i is the orbit inclination, n is the number of orbit revolutions per


day, and Ω and ω are in degrees per day.

4.9.8 Aerodynamic Drag


Aerodynamic drag on a satellite assuming it is a point mass m, is given by

1 ~
2 Vrel
F~d = − CD ρAVrel (4.291)
2 Vrel
where CD is the drag coefficient, A is the area of the satellite surface perpen-
dicular to V~rel , Vrel is the velocity of the satellite relative to the atmosphere,
and ρ is the density of the atmosphere.
Assuming the atmosphere is stationary, V ~rel equals to the satellite veloc-
ity ~v . In order to find the effect of the aerodynamic drag on the satellite
4.9. ORBITAL PERTURBATIONS 219

orbital parameters, we will use Lagrange planetary equations of motion. We


are required to find the components of the aerodynamic drag along î, ĵ,
and k̂ (orthogonal right-handed unit vectors along r (orbital radius), θ (true
anomaly), and z directions). The velocity of the satellite ~v can be expressed
as
~v = ṙî + rθ̇ ĵ (4.292)

Using the orbital motion relations h = µp and r = p/(1 + ecosθ), we get
hesinθ
ṙ = (4.293)
p
h
θ̇ = 2 (4.294)
r
where h is the orbital angular momentum per unit mass of the satellite, e is
the orbital eccentricity, and p is the semi-latus rectum.
Substituting the values of ṙ and θ̇ from Eqs. (4.293-4.294) into Eq.(4.292)
and applying the orbital motion relations, we get
µ
v 2 = ṙ2 + r2 θ̇2 = (1 + e2 + 2ecosθ) (4.295)
p

Using Eqs. (4.292-4.295), the aerodynamic drag as per Eq. 4.291 can be
written along î, ĵ, and k̂ as follows:
1 esinθ
fx = − CD ρAv 2 (4.296)
2m (1 + e2 + 2ecosθ)1/2
1 1 + ecosθ
fy = − CD ρAv 2 (4.297)
2m (1 + e2 + 2ecosθ)1/2
fz = 0 (4.298)

Putting the expressions for fx , fy and fz into Lagrange’s planetary equa-


tions of motion lead to
da A CD ρv 2 (1 + e2 + 2ecosθ)1/2
=− (4.299)
dt m n(1 − e2 )1/2
de A CD ρv (1 − e2 )(e + cosθ)
2
=− (4.300)
dt m na(1 + e2 + 2ecosθ)1/2
di
=0 (4.301)
dt
2 2
dω A CD ρv (1 − e )sinθ
=− (4.302)
dt m nae(1 + e2 + 2ecosθ)1/2
dΩ
=0 (4.303)
dt
dM A CD ρv 2 (1 − e2 )(1 + e2 + 2ecosθ)1/2 sinθ
=n + (4.304)
dt m nae(1 + ecosθ)
220 CHAPTER 4. DYNAMICS I

As per Eqs.(4.301) and (4.303), the orbital inclination i and the right
ascension of the ascending node Ω remain constant under the aerodynamic
drag. In other words, as there is no component of the atmospheric drag
normal to the orbit, the orientation of the orbit plane remain unaffected.
With regard to Eqs. (4.302) and (4.304), they contain the term sinθ and
therefore the solutions ω and M are periodic. As the drag forces are small,
the amplitude of oscillation must be relatively small. However, Eqs. (4.299-
4.300) indicate secular changes in a and e.
If we assume the initial orbit is circular (i.e., e=0), semi-major axis a
only gets effected as per Eqs. (4.299-4.300). Taking e=0, Eq. (4.299) can be
rewritten as
da CD ρAna2 ρCD Aµ1/2 a1/2
=− =− (4.305)
dt m m
or
da ρCD Aµ1/2
= − dt (4.306)
a1/2 m

Assuming the atmospheric density ρ remains constant (note: however,


this may only be true for small changes in a), we integrate Eq. (4.306) from
initial time ti to final time tf as

1/2 1/2 ρCD Aµ1/2 (tf − ti )


af − ai =− (4.307)
2m
where ai and af are the corresponding semi-major axes at ti and tf , re-
spectively. Note that Eq. (4.307) for ∆a=af − ai is only valid when ∆a is
small.

4.10 Summary
In this chapter, we have described motion of a satellite in Keplerian and non-
Keplerian orbits. Orbital perturbations due to Earth’s Oblateness, planetary
gravitational forces, and aerodynamic drag are explained. The important
results discussed in this chapter are summarized as follows.
4.10. SUMMARY 221

Angular Momentum and Energy


2 µ µ
~h = ~r × ~r˙ , E = v2 − r = − 2a

Eccentricity vector
1 ~r˙ × ~h − µ ~r
h i
~e = µ r

Line of Nodes
~n = K̂ × ~h

Orbit Equation
(h2 /µ) p
r= =
1 + ecosθ 1 + ecosθ

Eccentric Anomaly
q
tan E
2 = 1 − e tan θ ,
1+e 2 if E < 0 then E = E + 2π.

Mean Anomaly
M = n(t − tp ) = E − esinE, where E > 0 and in right quadrant.

Mean Angular Velocity


µ
q
n=
a3

Inclination
~
cosi = h · K̂
h

Argument of the Periapsis



 0o ≤ ω ≤ 180o if ~e · K̂ ≥ 0
cosω = ~nne
· ~e
 180o < ω ≤ 360o if ~e · K̂ < 0

Longitude of the Ascending Node



 0o ≤ Ω ≤ 180o if ~n · Jˆ ≥ 0
ˆ
cosΩ = ~nn· I
 180o < Ω ≤ 360o if ~n · Jˆ < 0

True Anomaly

 0o ≤ θ ≤ 180o if ~r · ~v ≥ 0
cosθ = ~rre
· ~e
 180o < θ ≤ 360o if ~r · ~v < 0
222 CHAPTER 4. DYNAMICS I

References

1. Broucke, R. A., and Cefda, P. J., Celestial Mechanics, Vol. 5, 1972, pp.
303-310.
2. R. Battin, An Introduction to the Mathematics and Methods of Astro-
dynamics, American Institute of Aeronautics and Astronautics, Inc.,
New York, 1987.
3. John E. Prussing, and Bruce A. Conway, Orbital Mechanics, Oxford
University Press, New York, 1993.
4. A.E. Roy, Orbital Motion, Third Edition, Adam Hilger/IDP Publishing,
1991.

5. P.R. Escobal, Methods of Orbit Determination, Krieger, 1965.


6. F.R. Moulton, An Introduction to Celestial Mechanics, Second Revised
Edition, Dover, 1970.
7. J.M.A. Danby, Fundamentals of Celestial Mechanics, Second Edition,
Willmann-Bell, 1988.
8. R.R. Bate, D.D. Mueller, and J.E. White, ”Fundamentals of Astrody-
namics,” Dover, 1971.
9. V.R. Bond and M.C. Allman, ”Modern Astrodynamics,” Princeton,
1996.
10. D. Brouwer and G.M. Clemence, ”Methods of Celestial Mechanics,”
Academic Press, 1961.
11. T. Logsdon, ”Orbital Mechanics,” Wiley, 1998.
12. S.W. McCuskey, ”Introduction to Celestial Mechanics,” Addison-Wesley,
1963.
13. W.M. Smart, Celestial Mechanics, Longmans, Green and Co., 1953.

14. V. Szebehely, Adventures in Celestial Mechanics, University of Texas


Press, 1989.

15. V. Szebehely, Theory of Orbits, Academic Press, 1982 (1967).


16. V. A. Chobotov, Orbital Mechanics, second edition, AIAA Education
series.

17. David A. Vallado, Fundamentals of Astrodynamics and Applications,


McGraw- Hill, New York, 1997.
4.10. SUMMARY 223

18. Seller J. J. (Editor), Undertsanding Space: An Introduction to Astro-


nautics, McGraw-Hill Inc., New York, 1994.
19. Brij N. Agrawal, Design of Geosynchronous Spacecraft, Prentice-Hall,
Englewood Cli(r)s, NJ, 1986.

Problem Set 4

4.1 At time t=0, a satellite on an Earth orbit with a=46800 km and e=0.85
is presently located at θ=52 deg. How much time has elapsed since the
satellite passed through periapsis? (µ=3.986× 105 km3 /s2 )
4.2 For the same orbit as in Problem 4.1, calculate the time at which the
satellite will reach θ=297 deg.
4.3 Radar measurements can determine the components of a satellite ve-
locity: vr (the radial component along the radius vector) and vθ (the
transverse component perpendicular to the radius vector). If a satellite
is observed at time t0 with vr =-3.475 km/s and a transverse component
vθ =5.940 km/s, at a radial distance of r=12595.9 km, then calculate
a) Flight-path angle
b) orbital period
c) eccentricity (magnitude only)
d) true anomaly
4.4 Derive an expression for the eccentricity e in terms of the initial radius
r0 , speed v0 , and flight-path angle β0 .
4.5 Starting with the expression for orbital position ~r = rcosθîe + rsinθîp ,
show that the orbital velocity is given in the perifocal system by

µ µ
~v = − sinθîe + (e + cosθ)îp
h h

Here îe and ip are inertial coordinates (i.e., dîe /dt = dîp /dt = 0).
4.6 How does J2 -perturbation affect the orbital elements? Which of the
elements on average will remain constant, and which on average will
change? The term on average means ”averaged over one revolution”.
If a spacecraft is in orbit around Earth with a=7000 km, e=0.08, and
i=28.5 deg, at what rate would the right ascension of the ascending
node change?
224 CHAPTER 4. DYNAMICS I
Chapter 5

Dynamics II

This chapter deals with the dynamics of a system under external forces. The
system may comprise of point masses, rigid bodies, flexible rigid bodies or a
combination of these. The chapter starts with Euler equations of motion for
a rigid body system. Next, Lagrange equations of motion are explained. The
chapter concludes with the derivation of equations of motion of rigid-bodies
with appendages, multi-body systems, and flexible body systems.

5.1 Introduction
The equations of motion for a rigid body system was stated by Leonard
Euler (1707-1783). In 1788, Joseph Louis Lagrange described the equations of
motion based on kinetic and potential energies of the system. These equations
are known as Lagrange’s equations of motion or Euler-Lagrange equations of
motion. We start with Euler Equations of Motion.

5.2 Euler Equations of Motion


Euler’s equation of motion for a rigid body is


T~ = H (5.1)

where T~ is torque acting on the body and H~ is the angular momentum of


the body. The rate of change of angular momentum H ~˙ is obtained
~ (i.e., , H)
by using the relation
226 CHAPTER 5. DYNAMICS II

~
~˙ = ( dH )xyz + ω
H ~
~ ×H (5.2)
dt

Taking T~ = Tx î + Ty ĵ + Tz k̂ and ω
~ = ωx î + ωy ĵ + ωz k̂, we get Euler’s
equation of motion as

T~ = (Ḣx − Hy ωz + Hz ωy )î + (Ḣy − Hz ωx + Hx ωz )ĵ


+ (Ḣz − Hx ωy + Hy ωx )k̂ (5.3)

or

Tx =Ḣx − Hy ωz + Hz ωy
Ty =Ḣy − Hz ωx + Hx ωz (5.4)
Tz =Ḣz − Hx ωy + Hy ωx

~ of a rigid body as
We have the angular momentum H

Hx =Ixx ωx − Ixy ωy − Ixz ωz


Hy =Iyy ωy − Iyx ωx − Iyz ωz (5.5)
Hz =Izz ωz − Izx ωx − Izy ωy

For the case where the body-fixed x, y, and z axes are the principal axes
(i.e., Iij = 0, i = x, y, z; j = x, y, z; i 6= j), Euler’s equations of motion
simplify to

Tx =Ixx ω̇x − (Iyy − Izz )ωy ωz


Ty =Iyy ω̇y − (Izz − Ixx )ωz ωx (5.6)
Tz =Izz ω̇z − (Ixx − Iyy )ωx ωy

5.3 Lagrange’s Equations of Motion


Lagrange’s equations of motion for a system with n generalized coordinates
are described by

  m
d ∂L ∂L X ∂fj
− = Qk + λj , k = 1, 2, . . . , n (5.7)
dt ∂ q˙k ∂qk j=1
∂qk
5.3. LAGRANGE’S EQUATIONS OF MOTION 227

Here L = T − U . The symbol L is called the Lagrangian function. T and


U are kinetic and potential energies of the system, respectively. Writing Eq.
(5.7) in terms of T and U and considering U to be independent of velocity,
we get
  m
d ∂T ∂T ∂U X ∂fj
− + = Qk + λj , j = 1, 2, . . . , n (5.8)
dt ∂ q˙k ∂qk ∂qk j=1
∂qk

qk (k = 1, 2, . . . , n) are generalized coordinates i.e., a set of coordinates,


that may be Cartesian/Spherical/Cylindrical coordinates, used to describe
uniquely the geometric position and/or the orientation of a system even
though they are not independent. If these coordinates are the minimum
number of coordinates necessary to completely describe the motion of the
system, we call them degrees of freedom and they are, in fact, independent
coordinates of the system. Qk is a generalized force corresponding to a gen-
eralized coordinate qk and is given by

p
∂~rj
F~j ·
X
Qk = (5.9)
j=1
∂q k

where F~j , j = 1, 2, . . . , p are the forces acting at ~rj . For a conservative force
F~j , such as gravitational force, that is derivable from a potential energy func-
tion, Qk =0. λj is a Lagrange multiplier corresponding to the fj constraint
given as

fj (q1 , q2 , . . . , qn ) = 0, j = 1, 2, . . . , m (5.10)

For a system of N particles subjected to m kinematical constraints, its


motion can be described uniquely by n independent coordinates qk (k =
1, 2, . . . , n) given as

n = 3N − m (5.11)

Example 5.1
Derive the orbital equations of motion of a dumbbell system described in
Example 2.4. Apply the Newton’s method and the Lagrange’s method.
Solution.
Newton’s Method
The orbital equation of motion of the dumbbell system is given by


F~ = M R (5.12)
228 CHAPTER 5. DYNAMICS II

where M = m1 + m2
The gravitational force exerted on the center of mass of the system at a
distance of R from the center of Earth is given by
~
µM R
F~ = − 3 (5.13)
R
where µ = GME is the gravitational constant. G is the Universal gravita-
tional constant and ME is the mass of the Earth. The (-) sign signifies the
attractive force between mass M and ME .
From the Chapter 2: Kinematics, Momentum and Energy,

~¨ = R̈ − Rθ̇2 îo + Rθ̈ + 2Ṙθ̇ ĵo


   
R (5.14)

the orbital equations of motion of the dumbbell system are obtained as


µM
M (R̈ − Rθ̇2 ) = − (5.15)
R2
M (Rθ̈ + 2Ṙθ̇) = 0 (5.16)
or
µ
R̈ − Rθ̇2 + =0 (5.17)
R2
Rθ̈ + 2Ṙθ̇ = 0 (5.18)

Lagrange’s Method
The equations of motion of the system with generalized coordinates q1 =
R and q2 = θ are given by
 
d ∂T ∂T ∂U
− + = 0, k = 1, 2 (5.19)
dt ∂ q̇k ∂qk ∂qk

The potential and kinetic energies of the system referring to Eqs. (2.153)
and (2.154) in Chapter 2 are
µM
U =− (5.20)
R
1
T = M (Ṙ2 + θ̇2 R2 ) (5.21)
2
Using the preceding expressions, the equations of motion of the system are
derived as follows:
R-equation
 
d ∂T ∂T ∂U
− + =0 (5.22)
dt ∂ Ṙ ∂R ∂R
5.3. LAGRANGE’S EQUATIONS OF MOTION 229

Here
 
∂T d ∂T
=M Ṙ, ⇒ = M R̈ (5.23)
∂ Ṙ dt ∂ Ṙ
∂T
=M θ̇2 R (5.24)
∂R
∂U µM
= 2 (5.25)
∂R R
Thus, R-equation of motion is
µM
M R̈ − M θ̇2 R + =0 (5.26)
R2

θ-equation
 
d ∂T ∂T ∂U
− + =0 (5.27)
dt ∂ θ̇ ∂θ ∂θ
Here
 
∂T 2 d ∂T
=M θ̇R ⇒ = M θ̈R2 + 2M θ̇RṘ (5.28)
∂ θ̇ dt ∂ θ̇
∂T ∂U
=0, =0 (5.29)
∂θ ∂θ

Thus, θ-equation of motion is

M θ̈R2 + 2M θ̇RṘ = 0 (5.30)

Example 5.2
Derive the inplane attitude equation of motion of a dumbbell system as
described in Example 2.4. Assume the system is moving in a circular orbit.
Apply the Newton’s method as well as the Lagrange’s method to derive the
equation of motion.
Solution.
Newton’s Method
The inplane attitude equation of motion of the dumbbell system is given
by


T~ = H (5.31)

where the angular momentum of the system is


~ = M R2 ω
H ~ o + Me L2 ω
~ (5.32)
230 CHAPTER 5. DYNAMICS II

Here M = m1 + m2 , Me = (m1 m2 )/(m1 + m2 ), ω


~ o = θ̇k̂o , and ω
~ = (θ̇ + β̇)k̂.
Assuming the system is in a circular orbit (i.e., Ṙ=θ̈=0), the rate of
change of the system angular momentum is
~˙ = Me L2 ω
H ~˙ = Me L2 β̈ k̂ (5.33)

Next we determine the external toque T~ due to gravitational force. The


gravitational force exerted on mass m1 at a distance of R1 from the center
of Earth is given by

~1
µm1 R
F~1 = − (5.34)
R13
Thus, the torque exerted on a mass of m1 is
~1
R
T~1 = ~r1 × F~1 = −µm1~r1 × 3 (5.35)
R1
~1 = R
Here R ~ + ~r1 . Substituting |R
~ 1 | in Eq.(5.35), we get
~ + ~r1 )
(R
T~1 = − µm1~r1 ×
~ · ~r1 + ~r2 )3/2
(R2 + 2R 1
" #−3/2
µm1 2 R~ · ~r1 r 2
~ 1+
= − 3 (~r1 × R) + 2 1
(5.36)
R R2 R

Applying Binomial series expansion for the term inside the bracket, we
get
" #
µm1 ~ · ~r1 ) 3 r2
3(R ~ · ~r1 )2
15 (R
~ ~
T1 = − 3 (~r1 × R) 1 − − 1
+ + ··· (5.37)
R R2 2 R2 2 R4

Now, considering the fact r1  R and carrying out expansion upto


O(1/R3 ), we get
" #
µm 1 3(R ~ · ~r1 ) 1
T~1 = − 3 (~r1 × R)
~ 1− + O( 3 )
R R2 R
" #
~
µR ~ · ~r1 )~r1
3(R
= 3 × m1~r1 − m1 (5.38)
R R2

Similarly, the torque exerted on mass m2 can be obtained as


" #
µR~ 3(R~ · ~r2 )~r2
T~2 = 3 × m2~r2 − m2 (5.39)
R R2
5.3. LAGRANGE’S EQUATIONS OF MOTION 231

Thus, the total torque exerted on the system is the sum of the torques
exerted on mass m1 and m2 . Adding T1 and T2 , we obtain the total torque
~ n
µR 3 o
T~ = 3 × m1~r1 + m2~r2 − 2 [m1 (R
~ · ~r1 )~r1 + m2 (R
~ · ~r2 )~r2 ] (5.40)
R R
Knowing
m2 ~
~r1 = − L
m1 + m2
m1 ~
~r2 = L
m1 + m2
and
m1~r1 + m2~r2 = 0
the preceding equation simplifies to

T~ = − 3 Me L2 (îo · î)(îo × î) (5.41)
R
or

T~ = − 3 Me L2 cosβsinβ k̂ (5.42)
R

Referring to Eqs. (5.31) and (5.33), we can write the equation of motion
of the system as

− Me L2 cosβsinβ = Me L2 β̈ (5.43)
R3
or
β̈ + 3ωo2 sinβcosβ = 0 (5.44)
p
where ωo = µ/R3 .
Lagrange’s Method
The equation of motion of the system is given by
 
d ∂T ∂T ∂U
− + =0 (5.45)
dt ∂ β̇ ∂β ∂β
Here qk = β and Qk = 0.
The potential and kinetic energies of the system referring to Eqs. (2.177)
and (2.188) in Chapter 2 are
µM µ
U =− + Me (1 − 3cos2 β)L2 (5.46)
R 2R3
1 1
T = M (θ̇2 R2 ) + Me (θ̇ + β̇)2 L2 . (5.47)
2 2
232 CHAPTER 5. DYNAMICS II

The equation of motion of the system is

β̈ + 3ωo2 sinβcosβ = 0 (5.48)

Example 5.3
Derive the in-plane translational (orbital) and rotational (attitude) mo-
tion of a dumbbell system using Newton’s method and Lagrange’s method.
Newton’s Method
The translational and rotational equations of the motion are given by

d~
p
F~ = (5.49)
dt
dH~
T~ = (5.50)
dt

where F~ and T~ denote total external force and torque acting at the center of
mass of the system. p
~ and H ~ specify the system linear (orbital) momentum
and attitude angular momentum, given by


p~ = M R, ~ = Me L2 ω
H ~ (5.51)

where M = m1 + m2 , Me = m1 m2 /(m1 + m2 ), ω ~ o = θ̇ k̂o , and ω


~ = (θ̇ + β̇)k̂o .
Taking derivatives with respect to time, we have

d~
p ~¨
=M R (5.52)
dt
dH~ d~
ω
=Me L2 (5.53)
dt dt
Knowing

~¨ = R̈ − Rθ̇2 îo + Rθ̈ + 2Ṙθ̇ ĵo


   
R (5.54)
d~
ωo
=θ̈ k̂o (5.55)
dt
d~
ω
=(θ̈ + β̈)k̂o (5.56)
dt
we get

d~
p    
=M R̈ − Rθ̇2 îo + M Rθ̈ + 2Ṙθ̇ ĵo (5.57)
dt
dH~
=Me L2 (θ̈ + β̈)k̂ (5.58)
dt
5.3. LAGRANGE’S EQUATIONS OF MOTION 233

Next we derive the external force and torque acting on the system. The
external force vector is the sum of the gravitational forces acting on m1 (F~1 )
and m2 (F~2 ), given as

F~ = F~1 + F~2 (5.59)

where F~1 is
~ + ~r1 )
(R
F~1 = − µm1
(R2~ · ~r1 + ~r2 )3/2
+ 2R 1
" #−3/2
µm1 2 ~
R · ~
r1 r 2
~ 1+
= − 3 (~r1 × R) + 12 (5.60)
R R2 R

Applying Binomial series expansion for the term inside the bracket, we
get
" #
µm1 ~ ~ · ~r1 ) 3 r2
3(R ~ · ~r1 )2
15 (R
~
F1 = − 3 (R + ~r1 ) 1 − − 1
+ + ··· (5.61)
R R2 2 R2 2 R4
or
(
µm ~ ~ ~ 2
1
F~1 = − 3 R ~ + ~r1 − 3(R + ~r1 )(R · ~r1 ) − 3 (R + ~r1 )r1
R R 2 2 R 2
)
~ + ~r1 )(R
15 (R ~ · ~r1 )2
+ + ··· (5.62)
2 R4

Now, considering the fact r1  R and carrying out expansion upto O(1/R4 )1
yields
(
µm ~ ~ ~ ~ 2
F~1 = − 3 R
1 ~ + ~r1 − 3R(R · ~r1 ) − 3~r1 (R · ~r1 ) − 3 Rr1
R R 2 R 2 2 R2
)
~ R
15 R( ~ · ~r1 )2
+ (5.63)
2 R4

Similarly, the external force acting on mass m2 is derived as


(
µm2 ~ ~ R
3R( ~ · ~r2 ) 3~r2 (R
~ · ~r2 ) 3 Rr
~ 2
~
F2 = − 3 R + ~r2 − − − 2
R R2 R2 2 R2
)
~ R
15 R( ~ · ~r2 )2
+ (5.64)
2 R4
1 Here, we consider the terms upto O(1/R4 ) as compared to the case of potential energy

where we considered upto O(1/R3 ) since F = ∂U/∂R.


234 CHAPTER 5. DYNAMICS II

Thus, the total external force acting on the system is


(
µ ~ ~
F~ = − 3 (m1 + m2 )R ~ + (m1~r1 + m2~r2 ) − 3R[R · (m1~r1 + m2~r2 )]
R R2
~ · ~r1 ) + m2~r2 (R
3[m1~r1 (R ~ · ~r2 )] 3 R(m~ 1 r 2 + m2 r 2 )
1 2
− −
R2 )2 R 2

~ 1 (R
15 R[m ~ · ~r1 )2 + m2 (R~ · ~r2 )2 ]
+ (5.65)
2 R4

Knowing

m1~r1 + m2~r2 = 0 (5.66)

and
~
~r1 = −γ L, ~
~r2 = (1 − γ)L, ~ = Lî,
L ~ = Rîo
R (5.67)

we write

m1 r12 + m2 r22 = Me L2 (5.68)

where Me = m1 m2 /(m1 + m2 ). Using the preceding relation with Eq. (5.65)


results in
( )
µ 3
F~ = − 3 M R~− 2 2
Me L [2(îo · î)î − 5(îo · î) îo + îo ] (5.69)
R 2R

where M = m1 + m2 .
For this problem of in-plane motion of the system, the unit vector î in the
body-fixed rotating frame Sxyz is related to the unit vectors in the orbital
reference frame Sxo yo zo as follows

î = cosβ îo + sinβ ĵo (5.70)

Applying the above relation, we have the resultant force vector on the system
as
( )
~ µ 3 2 2
F = − 3 M Rîo − Me L [−3cos β îo + sin2β ĵo + îo ] (5.71)
R 2R

Alternatively, the force F~ can derived from the system potential energy.
It is the gradient of potential energy U , given by

F~ = −∇U (5.72)
5.3. LAGRANGE’S EQUATIONS OF MOTION 235

where ∇ denotes the gradient of the potential energy U and is written as


∂U ∂U ∂U
∇U = î + ĵ + k̂ (5.73)
∂x ∂y ∂z

Here î, ĵ, and k̂ are the unit vectors of the body-fixed coordinate frame
S − xyz. For this planar problem we have
∂U ∂U
∇U = î + ĵ (5.74)
∂x ∂y
~ as
We write R
~ = Rîo
R (5.75)

Expressing îo in terms of unit vectors in the body-fixed coordinate frame


S − xyz

îo = cosβ î − sinβ ĵ (5.76)

we have
~ = Rcosβ î − Rsinβ ĵ
R (5.77)

Thus, we write

x = Rcosβ, y = −Rsinβ (5.78)

As we are required to derive the force F~ along unit vectors îo , ĵo , and k̂o
in the local-vertical coordinate frame S − xo yo zo , we express Eq. (5.74) in
terms of unit vectors îo , ĵo , and k̂o

   
∂U ∂U ∂U ∂U
∇U = cosβ − sinβ îo + sinβ + cosβ ĵo (5.79)
∂x ∂y ∂x ∂y

From the potential energy expression,

U = f (R, β) (5.80)

So, we have to express ∇U with respect to R and β.

 
∂U ∂U ∂x ∂U ∂y ∂U ∂U
= + = −R cosβ − sinβ (5.81)
∂R ∂x ∂R ∂y ∂R ∂x ∂y
 
∂U ∂U ∂x ∂U ∂y ∂U ∂U
= + = −R sinβ + (5.82)
∂β ∂x ∂β ∂y ∂β ∂x ∂y
236 CHAPTER 5. DYNAMICS II

Thus, we obtain
∂U 1 ∂U
∇U = îo − ĵo (5.83)
∂R R ∂β
Alternatively, we can obtain the same equation for ∇U by first deriving
∂U ∂U ∂R ∂U ∂β ∂U ∂η
= + + , X = x, y (5.84)
∂X ∂R ∂X ∂β ∂X ∂η ∂X
and substitute these derivatives in Eq. (5.79).
Knowing
µM µ
U =− + Me (1 − 3cos2 β)L2
R 2R3

the force F~ is obtained as


 
µ 3 µ
F~ = −∇U = − M− 2 2
Me (1 − 3cos β)L îo
R2 2 R4
 
3 µ 2
− Me L sin2β ĵo (5.85)
2 R4

Referring to the previous Example 5.2 the external torque due to gravi-
tational force acting on the system is

T~ = − 3 Me L2 cosβsinβ k̂ (5.86)
R

The equations of motion of the system are obtained as


µM 3µ
M (R̈ − Rθ̇2 ) + − Me L2 [1 − 3cos2 β] = 0 (5.87)
R2 2R4

M (Rθ̈ + 2Ṙθ̇) − Me L2 sin2β = 0 (5.88)
2R4

Me L2 (θ̈ + β̈) + 3 Me L2 cosβsinβ = 0 (5.89)
R
or

µM 3µ
M (R̈ − Rθ̇2 ) + 2
− Me L2 [1 − 3cos2 β] = 0 (5.90)
R 2R4

M (Rθ̈ + 2Ṙθ̇) − 4
Me L2 sin2β = 0 (5.91)
2R
Me L2
 
3 µ
β̈ + sin2β + sin2β =0 (5.92)
2 R3 M R2
5.3. LAGRANGE’S EQUATIONS OF MOTION 237

Lagrange’s Method
The equations of motion of the system with generalized coordinates q1 =
R, q2 = θ, and q3 = β are given by
 
d ∂T ∂T ∂U
− + = 0, k = 1, 2, 3 (5.93)
dt ∂ q̇k ∂qk ∂qk

The potential and kinetic energies of the system are


µM µ
U =− + Me (1 − 3cos2 β)L2 (5.94)
R 2R3
1 1
T = M (Ṙ2 + θ̇2 R2 ) + Me (θ̇ + β̇)2 L2 (5.95)
2 2

Using the preceding expressions, the equations of motion of the system


are derived as follows:
R-equation
 
d ∂T ∂T ∂U
− + =0 (5.96)
dt ∂ Ṙ ∂R ∂R
Here
 
∂T d ∂T
=M Ṙ, ⇒ = M R̈ (5.97)
∂ Ṙ dt ∂ Ṙ
∂T
=M θ̇2 R (5.98)
∂R
∂U µM 3µ
= 2 − Me (1 − 3cos2 β)L2 (5.99)
∂R R 2R4
Thus, R-equation of motion is
µM 3µ
M R̈ − M θ̇2 R + 2
− Me (1 − 3cos2 β)L2 = 0 (5.100)
R 2R4

θ-equation
 
d ∂T ∂T ∂U
− + =0 (5.101)
dt ∂ θ̇ ∂θ ∂θ
Here
∂T
=M θ̇R2 + Me (θ̇ + β̇)L2 (5.102)
 ∂ θ̇
d ∂T
⇒ =M θ̈R2 + 2M θ̇RṘ + Me (θ̈ + β̈)L2 (5.103)
dt ∂ θ̇
∂T ∂U
=0, =0 (5.104)
∂θ ∂θ
238 CHAPTER 5. DYNAMICS II

Note L̇ = 0 since the cable is rigid.


Thus, θ-equation of motion is

M θ̈R2 + 2M θ̇RṘ + Me (θ̈ + β̈)L2 = 0 (5.105)

β-equation
 
d ∂T ∂T ∂U
− + =0 (5.106)
dt ∂ β̇ ∂β ∂β

Here
∂T
=Me (θ̇ + β̇)L2 (5.107)
∂ β̇
 
d ∂T
⇒ =Me (θ̈ + β̈)L2 (5.108)
dt ∂ β̇
∂T
=0 (5.109)
∂β
∂U 3µ
= Me L2 cosβsinβ (5.110)
∂β R3

Thus, β-equation of motion is



Me (θ̈ + β̈)L2 + Me L2 cosβsinβ = 0 (5.111)
R3

Thus, the equations of motion of the system with simplifications can be


written as
µM 3µ
M (R̈ − Rθ̇2 ) + 2
− Me L2 [1 − 3cos2 β] = 0 (5.112)
R 2R4

M (Rθ̈ + 2Ṙθ̇) − Me L2 sin2β = 0 (5.113)
2R4
Me L2
 
Ṙ 3 µ
β̈ − θ̇ + sin2β + sin2β = 0 (5.114)
R 2 R3 M R2

Example 5.4
A system is comprised of a rigid-body spacecraft m1 and an auxiliary
mass m2 (a point mass) orbiting about the Earth. The auxiliary mass m2 is
attached to the spacecraft m1 by a flexible cable of length L at an offset of
~a = aî from the center of mass of the spacecraft, where î and ĵ are unit vectors
in the spacecraft body-fixed frame Sxyz . The auxiliary mass m2 undergoes an
in-plane libration or oscillation of β about the local vertical. The mass of the
spacecraft is very large in comparison to the auxiliary mass m2 (i.e., m1 
5.3. LAGRANGE’S EQUATIONS OF MOTION 239

m2 ). Assume the cable is massless, but flexible with rigid length Lo and
longitudinal deformation u (i.e., L = Lo +u). Determine the system equations
of motion with degrees of freedom or generalized coordinates as R, θ, α, β,
and u. Take Iz as the mass moment of inertia of m1 about the body-fixed
z-axis. Note. In the figure, ~a makes an angle of α with the local vertical.

m2

Orbit L

β
a
S m1
Y
R

θ
E
X

Solution.
Using Summary Sheet (System: One Rigid Body and One Point Mass)
and taking ~r1 = 0, we derive the kinetic and potential energies of the system
as
1 1 n
T = M [Ṙ2 + θ̇2 R2 ] + m2 ω 2 a2 + u̇2 + ωL 2 2
L + 2ωa[−u̇sin(α − β)
2 2
o 1
+ ωL Lcos(α − β)] + Iz ω 2 (5.115)
2
µM µm2 2
n
2 2
o
U =− + a + L + 2aLcos(α − β) − 3[acosα + Lcosβ]
R 2R3
µ n o 1 EA
+ (Ix + Iy + Iz ) − 3[Iz + (Iy − Ix )cos2α] + u2 (5.116)
4R3 2 L0

where ω = θ̇ + α̇ and ωL = θ̇ + β̇.


Applying Lagrange’s method, the equations of the motion of the system
are derived as follows:
R-equation
 
d ∂T ∂T ∂U
− + =0 (5.117)
dt ∂ Ṙ ∂R ∂R
240 CHAPTER 5. DYNAMICS II

Here
 
∂T d ∂T
=M Ṙ, ⇒ = M R̈ (5.118)
∂ Ṙ dt ∂ Ṙ
∂T
=M θ̇2 R (5.119)
∂R
∂U µM 3µm2 2n
2 2
o
= 2 − a + L + 2aLcos(α − β) − 3[acosα + Lcosβ]
∂R R 2R4
3µ n o
− (Ix + I y + Iz ) − 3[Iz + (Iy − Ix )cos2α]
4R4
(5.120)

Thus, R-equation of motion is

µM 3µm2 n 2
M R̈ − M θ̇2 R + 2
− a + L2 + 2aLcos(α − β)
R 2R4
o 3µ n
− 3[acosα + Lcosβ]2 − 4
(Ix + Iy + Iz )
o4R
− 3[Iz + (Iy − Ix )cos2α] = 0 (5.121)

θ-equation
 
d ∂T ∂T ∂U
− + =0 (5.122)
dt ∂ θ̇ ∂θ ∂θ

Here

∂T n
=M θ̇R2 + m2 ωa2 + ωL L2 + a[−u̇sin(α − β)
∂ θ̇ o
+ ωL Lcos(α − β)] + ωaLcos(α − β)
 
d ∂T n
⇒ =M θ̈R2 + 2M θ̇RṘ + m2 ω̇a2 + ω̇L L2 + 2ωL Lu̇
dt ∂ θ̇
+ a[−üsin(α − β) − u̇(α̇ − β̇)sin(α − β)
+ ω̇L Lcos(α − β) + ωL u̇cos(α − β)
− ωL L(α̇ − β̇)sin(α − β)] + ω̇aLcos(α − β)
o
+ ωau̇cos(α − β) − ω(α̇ − β̇)aLsin(α − β) (5.123)
∂T ∂U
=0, =0 (5.124)
∂θ ∂θ

where ω̇ = θ̈ + α̈, ω̇L = θ̈ + β̈.


5.3. LAGRANGE’S EQUATIONS OF MOTION 241

Thus, θ-equation of motion is


n
M θ̈R2 + 2M θ̇RṘ + m2 ω̇a2 + +ω̇L L2 + 2ωL Lu̇
+ a[−üsin(α − β) − u̇(α̇ − β̇)sin(α − β) + ω̇L Lcos(α − β)
+ ωL u̇cos(α − β) − ωL L(α̇ − β̇)sin(α − β)] + ω̇aLcos(α − β)
o
+ ωau̇cos(α − β) − ω(α̇ − β̇)aLsin(α − β) = 0 (5.125)

α-equation
 
d ∂T ∂T ∂U
− + =0 (5.126)
dt ∂ θ̇ ∂θ ∂θ
Here
∂T n o
=Iz ω + m2 ωa2 + a[−u̇sin(α − β) + ωL Lcos(α − β)]
 ∂ α̇
d ∂T n
⇒ =Iz ω̇ + m2 ω̇a2 + a[−üsin(α − β) − u̇(α̇ − β̇)cos(α − β)
dt ∂ α̇
+ ω̇L Lcos(α − β) + ωL u̇cos(α − β)
o
− ωL (α̇ − β̇)Lsin(α − β)] (5.127)
∂T n o
= − m2 ωa u̇cos(α − β) + ωL Lsin(α − β) (5.128)
∂α
∂U µm2 n o
= 3 − aLsin(α − β) + 3[acosα + Lcosβ]asinα
∂α R

+ (Iy − Ix )sin2α (5.129)
2R3
Thus, α-equation of motion is
n
Iz ω̇ + m2 ω̇a2 + a[−üsin(α − β) − u̇(α̇ − β̇)cos(α − β)
o
+ ω̇L Lcos(α − β) + ωL u̇cos(α − β) − ωL (α̇ − β̇)Lsin(α − β)]
n o
+ m2 ωa u̇cos(α − β) + ωL Lsin(α − β)
µm2 n o
+ 3 − aLsin(α − β) + 3[acosα + Lcosβ]asinα
R

+ (Iy − Ix )sin2α = 0 (5.130)
2R3

β-equation
 
d ∂T ∂T ∂U
− + =0 (5.131)
dt ∂ β̇ ∂β ∂β
242 CHAPTER 5. DYNAMICS II

Here
∂T n o
=m2 ωL L2 + ωa[Lcos(α − β)]
∂ β̇
 
d ∂T n
⇒ =m2 ω̇L L2 + 2ωL Lu̇ + ω̇a[Lcos(α − β)] + ωau̇cos(α − β)
dt ∂ β̇
o
− ω(α̇ − β̇)aLsin(α − β) (5.132)
∂T n o
=m2 ωa[u̇cos(α − β) + ωL Lsin(α − β)] (5.133)
∂β
∂U µm2 n o
= 3 aLsin(α − β) + 3[acosα + Lcosβ]Lsinβ (5.134)
∂β R
Thus, β-equation of motion is
n
m2 ω̇L L2 + 2ωL Lu̇ + ω̇a[Lcos(α − β)] + ωau̇cos(α − β)
o n
− ω(α̇ − β̇)aLsin(α − β) − m2 ωa[u̇cos(α − β)
o µm n
2
+ ωL Lsin(α − β)] + 3 aLsin(α − β)
Ro
+ 3[acosα + Lcosβ]Lsinβ = 0 (5.135)

u-equation
 
d ∂T ∂T ∂U
− + =0 (5.136)
dt ∂ u̇ ∂u ∂u
Here
∂T n o
=m2 u̇ − ωasin(α − β)
 ∂ u̇
d ∂T n o
⇒ =m2 ü − ω̇asin(α − β) − ω(α̇ − β̇)acos(α − β) (5.137)
dt ∂ u̇
∂T n
2
o
=m2 ωL L + ωaωL cos(α − β) (5.138)
∂u
∂U µm2 n o
= 3 L + acos(α − β) − 3[acosα + Lcosβ]cosβ
∂u R
EA
+ u (5.139)
L0
Thus, u-equation of motion is
n o
m2 ü − ω̇asin(α − β) − ω(α̇ − β̇)acos(α − β)
n o µm n
2 2
− m2 ω L L + ωaωL cos(α − β) + 3 L + acos(α − β)
R
o EA
− 3[acosα + Lcosβ]cosβ + u=0 (5.140)
L0
5.3. LAGRANGE’S EQUATIONS OF MOTION 243

Example 5.5
Derive equations of motion of the system given in Example 5.4 using
Lagrange’s Method with multipliers considering the generalized coordinates
as

q = [α, βc , Lc , u]T (5.141)

where Lc is the distance between the centers of mass of m1 and m2 and βc


~ c makes with the local vertical.
is the angle that L
Solution. The generalized coordinates are not independent and they are
~ = −~a + L
related by a constraint equation obtained using the relation L ~ c , as

f = L − [a2 + L2c − 2aL1 cos(α − βc )]1/2 = 0 (5.142)

where L = Lo + u.
Knowing r1 = 0 and r2 = Lc , the kinetic and potential energies of the
system are obtained as
1 1 1
T = M [Ṙ2 + θ̇2 R2 ] + m2 [L̇2c + ωL
2 2
Lc ] + Iz ω 2 (5.143)
2 2 2
µ(m1 + m2 ) µ 2 2
U =− + m2 Lc [1 − 3cos βc ]
R 2R3
µ n o 1 EA
+ 3
(Ix + Iy + Iz ) − 3[Iz + (Iy − Ix )cos2α] + u2 (5.144)
4R 2 L0

where ω = θ̇ + α̇ and ωL = θ̇ + β̇1 .


Applying Lagrange’s method, the equations of the motion of the system
are derived as follows:
R-equation
 
d ∂T ∂T ∂U ∂f
− + =λ (5.145)
dt ∂ Ṙ ∂R ∂R ∂R
Here
 
∂T d ∂T
=M Ṙ, ⇒ = M R̈ (5.146)
∂ Ṙ dt ∂ Ṙ
∂T
=M θ̇2 R (5.147)
∂R
∂U µM 3µm2
= 2 − m2 L2c [1 − −3cos2 βc ]
∂R R 2R4
3µ n o
− (Ix + Iy + Iz ) − 3[Iz + (Iy − Ix )cos2α]
4R4
∂f
=0 (5.148)
∂R
244 CHAPTER 5. DYNAMICS II

Thus, R-equation of motion is


µM 3µm2 2
M R̈ − M θ̇2 R + 2 − L [1 − 3cos2 βc ]
R 2R4 c
3µ n o
− (Ix + Iy + Iz ) − 3[Iz + (Iy − Ix )cos2α] =0 (5.149)
4R4

θ-equation
 
d ∂T ∂T ∂U ∂f
− + =λ (5.150)
dt ∂ θ̇ ∂θ ∂θ ∂θ

Here
∂T
=M θ̇R2 + m2 ωL L2c
 ∂ θ̇

d ∂T
⇒ =M θ̈R2 + 2M θ̇RṘ + m2 (ω̇L L2c + 2ωL L̇c ) (5.151)
dt ∂ θ̇
∂T ∂U ∂f
=0, = 0, =0 (5.152)
∂θ ∂θ ∂θ

where ω̇ = θ̈ + α̈, ω̇L = θ̈ + β̈.


Thus, θ-equation of motion is

M θ̈R2 + 2M θ̇RṘ + m2 (ω̇L L2c + 2ωL L̇c ) = 0 (5.153)

α-equation
 
d ∂T ∂T ∂U ∂f
− + =λ (5.154)
dt ∂ α̇ ∂α ∂α ∂α

Here
 
∂T d ∂T
=Iz ω ⇒ = Iz ω̇ (5.155)
∂ α̇ dt ∂ α̇
∂T
=0 (5.156)
∂α
∂U 3µ
= (Iy − Ix )sin2α
∂α 2R3
∂f 1
= − aLc sin(α − βc ) (5.157)
∂α L
Thus, α-equation of motion is
3µ λ
Iz ω̇ + 3
(Iy − Ix )sin2α + aLc sin(α − βc ) = 0 (5.158)
2R L
5.3. LAGRANGE’S EQUATIONS OF MOTION 245

βc -equation
 
d ∂T ∂T ∂U ∂f
− + =λ (5.159)
dt ∂ β˙c ∂βc ∂βc ∂βc
Here
∂T
=m2 ωL L2c
∂ β˙c
 
d ∂T
⇒ =m2 (ω̇L L2c + 2ωL L̇c ) (5.160)
dt ∂ β̇c
∂T
=0 (5.161)
∂βc
∂U 3µm2 2
= 3 Lc 3cosβc sinβc (5.162)
∂βc R
∂f 1
= aLc sin(α − βc ) (5.163)
∂βc L
Thus, βc -equation of motion is
n o 3µm λ
2 2
m2 ω̇L L2 + 2ωL L̇c + L 3cosβc sinβc − aLc sin(α − βc ) = 0
R3 c L
(5.164)

Lc -equation
 
d ∂T ∂T ∂U ∂f
− + =λ (5.165)
dt ∂ L̇c ∂Lc ∂Lc ∂Lc
Here
∂T
=m2 L̇c
∂ L̇c
 
d ∂T
⇒ =m2 L̈c (5.166)
dt ∂ L̇c
∂T 2
=m2 ωL Lc (5.167)
∂Lc
∂U µm2
= 3 Lc (1 − 3cos2 βc ) (5.168)
∂Lc R
∂f 1
= − [Lc − aLc cos(α − βc )] (5.169)
∂Lc L
Thus, Lc -equation of motion is

2 µm2 λ
m2 L̈c − m2 ωL Lc + Lc (1 − 3cos2 βc ) + [Lc − aLc cos(α − βc )] = 0
R3 L
(5.170)
246 CHAPTER 5. DYNAMICS II

u-equation
 
d ∂T ∂T ∂U ∂f
− + =λ (5.171)
dt ∂ u̇ ∂u ∂u ∂u
Here
 
∂T d ∂T
=0 ⇒ =0 (5.172)
∂ u̇ dt ∂ u̇
∂T
=0 (5.173)
∂u
∂U EA
= u (5.174)
∂u L0
∂f
=1 (5.175)
∂u
Thus, u-equation of motion is
EA
u=λ (5.176)
L0

Substituting λ from the preceding Eq. (5.170) into Eqs. (5.158) to (5.176),
we have three equations of motion corresponding to α, βc , Lc degrees of
freedom of the system.
Example 5.6
Derive the equations of motion of a rigid body satellite undergoing three-
dimensional attitude motion. Consider that the satellite is in a circular orbit
and only gravitational force is acting on it. Apply Euler’s method as well as
Lagrange’s method to derive the equations of motion.
Solution.
We consider dimensionless parameters so that that our analysis and re-
sults are applicable to any type of spacecraft system undergoing attitude
motion. Keeping these aspects into consideration, we take the following pa-
rameters:

(a) Spacecraft Moment of Inertia:


We define new dimensionless parameters k1 and k2 given by
Iz − Ix Iz − Iy
k1 = , k2 = (5.177)
Iy Ix
The other parameters like (Iy /Ix ) and (Iz /Ix ) are obtained from Eqs.
(5.177) as
Ix 1 − k1 Iy 1 − k2
kxz = = , kyz = = (5.178)
Iz 1 − k1 k2 Iz 1 − k1 k2
5.3. LAGRANGE’S EQUATIONS OF MOTION 247

(b) Time (t):


The variable of integration t is changed to θ, an angle measured from
the reference line using the following transformation
∂q ∂q ∂θ
q̇ = = = q 0 θ̇ (5.179)
∂t ∂θ ∂t
∂2q ∂(q 0 θ̇) ∂q 0 ∂ θ̇ ∂q 0 ∂θ
q̈ = 2 = = θ̇ + q 0 = θ̇ + q 0 θ̈
∂t ∂t ∂t ∂t ∂θ ∂t
=q 00 θ̇2 + q 0 θ̈ (5.180)

where q 0 = ∂q/∂θ.

Euler’s Method
The attitude equations of motion of the rigid body about the body-fixed
principal x, y, and z axes are given by

Tx =Ix ω̇x − (Iy − Iz )ωy ωz (5.181)


Ty =Iy ω̇y − (Iz − Ix )ωz ωx (5.182)
Tz =Iz ω̇z − (Ix − Iy )ωx ωy (5.183)

Here the angular velocity vector components considering 3-2-1 sequence of


Euler angles (α, φ, γ) are

ωx = − (θ̇ + α̇)sinφ + γ̇ (5.184)


ωy =(θ̇ + α̇)cosφ sin γ + φ̇cosγ (5.185)
ωz =(θ̇ + α̇)cosφ cos γ − φ̇sinγ (5.186)

Knowing R2 θ̇ = h and R = a(1 − e2 )/(1 + ecosθ), we can derive θ̈ and Ṙ as

2Ṙθ̇ 2µ
θ̈ = − = − 3 esinθ (5.187)
R R
Resinθ
Ṙ = θ̇ (5.188)
1 + esinθ
Using the preceding Eqs.(5.177-5.188), the equations of motion of the rigid
satellite are obtained as
Satellite: Pitch(α)

α00 cosφcosγ − φ00 sinγ − (1 − kxz + kyz )(1 + α0 )φ0 sinφcosγ


− (1 + kxz − kyz )[(1 + α0 )γ 0 cosφsinγ + φ0 γ 0 cosγ]
Tz
+ (kxz − kyz )(1 + α0 )2 sinφcosφsinγ = (5.189)
Iz ωo2
248 CHAPTER 5. DYNAMICS II

Satellite: Roll(φ)
kyz α00 cosφsinγ + kyz φ00 cosγ − (1 − kxz + kyz )(1 + α0 )φ0 sinφsinγ
− (1 − kxz − kyz )[(1 + α0 )γ 0 cosφcosγ − φ0 γ 0 sinγ]
Ty
+ (1 − kxz )(1 + α0 )2 sinφcosφcosγ = (5.190)
Iz ωo2

Satellite: Yaw(γ)
−kxz α00 sinφ + kxz γ 00 + [(1 − kyz )cos2γ − kxz ](1 + α0 )φ0 cosφ
Tx
+ (1 − kyz )[(1 + α0 )2 cos2 φ − φ02 ]sinγcosγ =
Iz ωo2
(5.191)
p
where ωo = θ̇ = µ/R3 , ()0 = d()/dθ, and ()00 = d2 ()/dθ2 . Tx , Ty , and
Tz are the gravitational torques about rotating body-fixed x, y, and z axes.
Here kxz = Ix /Iz and kyz = Iy /Iz .
The gravitational torques are derived next. The gravitational force ex-
erted on mass dm at a distance of Rm from the center of Earth is given
by
µdmR~m
F~ = − 3
(5.192)
Rm
where µ = GME is the gravitational constant, G is the Universal gravitational
constant and ME is the mass of the Earth. The (-) sign signifies the attractive
force between mass dm and ME . Thus, the torque exerted on a mass of dm
is
Z Z
R~m
T~ = ~ × F~ = −µ
ρ ρ
~ × 3 dm (5.193)
m m Rm
~m = R
Here R ~ +ρ ~ m | in Eq.(5.193), we get
~. Substituting |R

Z ~ +ρ
(R ~)dm
T~ = − µ ρ

m 2 ~
(R + 2R · ρ~ + ρ2 )3/2
" #−3/2
µ
Z
2 ~ ·ρ
R ~ ρ 2
=− 3 (~ ~ 1+
ρ × R) + 2 dm (5.194)
R m R2 R

Applying Binomial series expansion for the term inside the bracket, we
get
" #
µ ~ ~ ) 3 ρ2 ~ · ρ~)2
~ 1 − 3(R · ρ 15 (R
Z
T~ = − 3 (~
ρ × R) − + + ··· (5.195)
R m R2 2 R2 2 R4
5.3. LAGRANGE’S EQUATIONS OF MOTION 249

Now, considering the fact ρ  R and carrying out expansion upto O(1/R3 ),
we get
" #
µ
Z
3(R~ · ρ~) 1
T~ = − 3 (~ ~ 1−
ρ × R) + O( 3 ) dm
R m R2 R
"Z #
~
µR
Z ~ ·ρ
3(R ~)~ ρ
= 3 × ρ
~dm − 2
dm (5.196)
R m m R

The R~ and ρ~ are defined with respect to the orbital coordinate frame
S−Xo Yo Zo and the satellite-body fixed coordinate frame S−xyz, respectively
as

~ = Rîo
R (5.197)

ρ
~ = xî + y ĵ + z k̂ (5.198)

~ and ρ
Substituting R ~ in Eq.(5.196), we get

~
µR
Z
3
Z 
T~ = 3 × ρ
~dm − x2 (îo · î)î + y 2 (îo · ĵ)ĵ + z 2 (îo · k̂)k̂
R m R m

+ xy[(îo · î)ĵ + (îo · ĵ)î] + xz[(îo · î)k̂ + (îo · k̂)î]


 !
+ yz[(îo · ĵ)k̂ + (îo · k̂)ĵ] dm (5.199)

As the center of mass lies at S, we get

Z
ρ~dm = 0 (5.200)
m

We have the following mass moment of inertia relations:


Z Z
Ixx = (y 2 + z 2 )dm; Iyy = (z 2 + x2 )dm;
m m
Z Z
2 2
Izz = (x + y )dm; Ixy = xydm; (5.201)
Zm Zm
Iyz = yzdm; Ixz = xzdm
m m
250 CHAPTER 5. DYNAMICS II

From Eqs. (5.201), we can obtain

Iyy + Izz − Ixx Izz + Ixx − Iyy


Z Z
x2 dm = ; y 2 dm = ;
2 2
Zm m
Ixx + Iyy − Izz
z 2 dm = (5.202)
m 2

Using Eqs. (5.200-5.202) into Eq. (5.199), we have


(

T~ = − 3 (Iyy + Izz − Ixx )[(îo · î)(îo × î)]
2R
+ (Izz + Ixx − Iyy )[(îo · ĵ)(îo × ĵ)]
+ (Ixx + Iyy − Izz )[(îo · k̂)(îo × k̂)]
+ 2Ixy [(îo · î)(îo × ĵ) + (îo · ĵ)(îo × î)]
+ 2Ixz [(îo · î)(îo × k̂) + (îo · k̂)(îo × î)]
)
+ 2Iyz [(îo · ĵ)(îo × k̂) + (îo · k̂)(îo × ĵ)] (5.203)

In the case where the axes x, y, and z are the principal axes, the product
of inertia terms vanishes, i.e., (Ixy = Ixz = Iyz = 0). Then, the torque T
exerted on the the satellite is

(

T~ = − 3 (Iyy + Izz − Ixx )[(îo · î)(îo × î)]
2R
+ (Izz + Ixx − Iyy )[(îo · ĵ)(îo × ĵ)]
)
+ (Ixx + Iyy − Izz )[(îo · k̂)(îo × k̂)] (5.204)

Let îo = Cxo x î + Cxo y ĵ + Cxo z k̂ and T~ = Tx î + Ty ĵ + Tz k̂. We can write


Eq.(5.204) as


Tx = − (Iyy − Izz )Cxo y Cxo z
R3

Ty = − 3 (Izz − Ixx )Cxo x Cxo z (5.205)
R

Tz = − 3 (Ixx − Iyy )Cxo x Cxo y
R
5.3. LAGRANGE’S EQUATIONS OF MOTION 251

In addition to the contribution to gravitational force, Earth’s oblate-


ness makes a contribution to gravity-gradient torque with a coefficient of
3µJ2 Re2 /R5 . By comparing this coefficient with the main term, 3µ/R3 , it is
found that at geosynchronous orbit the contribution of J2 to gravity-gradient
torque is approximately 5 orders of magnitude less than the main term. Thus,
the oblateness torque effects can be ignored for studying the attitude motion
of satellites around planetary bodies. However, such effects are significant
while analyzing satellite attitude motion around asteroids.
The transformation from the frame S − io jo ko to the body fixed frame
S − ijk using 3-2-1 Euler angle rotation sequence becomes

Cxo x =cosαcosφ (5.206)


Cxo y =cosαsinφsinγ − sinαcosγ (5.207)
Cxo z =cosαsinφcosγ + sinαsinγ (5.208)

Substituting these into Eq. (5.205), we have


Tx = − (Iyy − Izz )(cosαsinφcosγ + sinαsinγ)
R3
× (cosαsinφsinγ − sinαcosγ)

Ty = − 3 (Izz − Ixx )(cosαsinφcosγ + sinαsinγ)cosαcosφ (5.209)
R

Tz = − 3 (Ixx − Iyy )(cosαsinφsinγ − sinαcosγ)cosαcosφ
R

The equations of motion of the satellite can be written as


Satellite: Pitch(α)

α00 cosφcosγ − φ00 sinγ − (1 − kxz + kyz )(1 + α0 )φ0 sinφcosγ


− (1 + kxz − kyz )[(1 + α0 )γ 0 cosφsinγ + φ0 γ 0 cosγ]
+ (kxz − kyz )(1 + α0 )2 sinφcosφsinγ
+ 3(kxz − kyz )(cosαsinφsinγ − sinαcosγ)cosαcosφ = 0
(5.210)

Satellite: Roll(φ)

kyz α00 cosφsinγ + kyz φ00 cosγ − (1 − kxz + kyz )(1 + α0 )φ0 sinφsinγ
− (1 − kxz − kyz )[(1 + α0 )γ 0 cosφcosγ − φ0 γ 0 sinγ]
+ (1 − kxz )(1 + α0 )2 sinφcosφcosγ
+ 3(1 − kxz )(cosαsinφcosγ + sinαsinγ)cosαcosφ = 0
(5.211)
252 CHAPTER 5. DYNAMICS II

Satellite: Yaw(γ)

−kxz α00 sinφ + kxz γ 00 + [(1 − kyz )cos2γ − kxz ](1 + α0 )φ0 cosφ
+ (1 − kyz )[(1 + α0 )2 cos2 φ − φ02 ]sinγcosγ
− 3(1 − kyz )(cosαsinφcosγ + sinαsinγ)
× (cosαsinφsinγ − sinαcosγ) = 0 (5.212)

Lagrange’s Method
The potential and kinetic energies of the satellite are

(
µM µ
U =− − (Iyy + Izz − Ixx )[3(cosαcosφ)2 − 1]
R 4R3
+ (Izz + Ixx − Iyy )[3(cosαsinφsinγ − sinαcosγ)2 − 1]
)
+ (Ixx + Iyy − Izz )[3(cosαsinφcosγ + sinαsinγ)2 − 1]

(5.213)
1 1
T = M (θ̇2 R2 ) + [Ix ωx2 + Iy ωy2 + Iz ωz2 ] (5.214)
2 2

where M = m1 + m2 and

ωx = − (θ̇ + α̇)sinφ + γ̇
ωy =(θ̇ + α̇)cosφ sin γ + φ̇cosγ (5.215)
ωz =(θ̇ + α̇)cosφ cos γ − φ̇sinγ

Substituting the expressions for U and T in the Lagrangian Equations of


motion, we obtain the equations of motion of the satellite as follows:
Satellite: Pitch(α)

 
d ∂T ∂T ∂U
− + =0 (5.216)
dt ∂ α̇ ∂α ∂α
5.3. LAGRANGE’S EQUATIONS OF MOTION 253

(cos2 φcos2 γ + kxz sin2 φ + kyz cos2 φsin2 γ)α00 + (kyz − 1)φ00 sinγcosγcosφ
− kxz γ 00 sinφ + (1 + α0 )[−φ0 sin2φcos2 γ − γ 0 cos2 φsin2γ
+ kxz φ0 sin2φ + kyz (−φ0 sin2φsin2 γ + γ 0 cos2 φcos2γ)]
− φ0 γ 0 cosφcos2 γ + φ0 sinγ(φ0 sinφcosγ + γ 0 cosφsinγ)
− kxz φ0 γ 0 cosφ − kyz φ0 γ 0 cosφsin2 γ + kyz φ0 cosγ(−φ0 sinφsinγ
3
+ γ 0 cosφcosγ) − [(kxz + kyz − 1)(cosαsinφcosγ
2
+ sinαsinγ)(−sinαsinφcosγ + cosαsinγ)
+ (1 + kxz − kyz )(cosαsinφsinγ − sinαcosγ)(−sinαsinφsinγ
− cosαcosγ) − (1 − kxz + kyz )sinαcosαcos2 φ] = 0 (5.217)

Satellite: Roll(φ)

 
d ∂T ∂T ∂U
− + =0 (5.218)
dt ∂ φ̇ ∂φ ∂φ

(kyz − 1)α00 cosφcosγsinγ + (sin2 γ + kyz cos2 γ)φ00


+ (1 + α0 )(1 − kyz )(φ0 sinφcosγsinγ − γ 0 cosφcos2γ)
+ (1 − kyz )φ0 γ 0 sin2γ + (1 + α0 ){[(1 + α0 )cosφcosγ − φ0 sinγ]sinφcosγ
+ kxz [−(1 + α0 )sinφcosφ + γ 0 ]cosφ + kyz [(1 + α0 )cosφsinγ
+ φ0 cosγ]sinφsinγ} − (3/2)[(kxz + kyz − 1)(cosαsinφcosγ
+ sinαsinγ)cosαcosφcosγ + (1 − kxz + kyz )cosαcosφ(−cosαsinφ)
+ (1 + kxz − kyz )(cosαsinφsinγ − sinαcosγ)(cosαcosφsinγ)] = 0
(5.219)

Satellite: Yaw(γ)

 
d ∂T ∂T ∂U
− + =0 (5.220)
dt ∂ γ̇ ∂γ ∂γ

−kxz α00 sinφ + kxz γ 00 + [(1 − kyz )cos2γ − kxz ](1 + α0 )φ0 cosφ
+ (1 − kyz )[(1 + α0 )2 cos2 φ − φ02 ]sinγcosγ
− 3(1 − kyz )(cosαsinφcosγ + sinαsinγ)
× (cosαsinφsinγ − sinαcosγ) = 0 (5.221)
254 CHAPTER 5. DYNAMICS II

In this example problem we have derived the three-dimensional attitude


equations of motion of a rigid body using Euler’s method as well as the
Lagrange’s method. However, these equations of motion are different - only
satellite yaw(γ) equation of motion in both the cases happens to be the same.
This can be explained by deriving generalized forces Qk , k = α, φ, γ based on
the virtual work method. The virtual angular displacement λ of the satellite
with respect to 3 − 2 − 1 Euler angle sequence is given by
~λ =αk̂o + φĵ2 + γ î (5.222)

The generalized forces Qk , k = α, φ, γ are given by

∂~λ ∂~λ ∂~λ


Qα = T~ · , Qφ = T~ · , Qγ = T~ · (5.223)
∂α ∂φ ∂γ
Using the relation (5.222) yields

Qα = T~ · k̂o , Qφ = T~ · ĵ2 , Qγ = T~ · î (5.224)

Knowing

T~ =Tγ î + Tφ ĵ + Tα k̂

and

k̂o = − sinφî + cosφsinγ ĵ + cosφcosγ k̂


ĵ2 =cosγ ĵ − sinγ k̂

we have

Qα = −Tγ sinφ + Tφ cosφsinγ + Tα cosφcosγ (5.225)


Qφ = Tφ cosγ − Tα sinγ (5.226)
Qγ = T γ (5.227)

Thus, we can write the Lagrange equations of motion (Lk , k = α, φ, γ) in


relation to the Euler’s equations of motion (Ek , k = α, φ, γ) as

Lα = Eα cosφcosγ + Eφ cosφsinγ − Eγ sinφ (5.228)


Lφ = −Eα sinγ + Eφ cosγ (5.229)
Lγ = Eγ (5.230)

Example 5.7
Derive the three-dimensional attitude equations of motion of a rigid body
satellite in an elliptic orbit. Only gravity gradient torque is acting on the
satellite.
5.3. LAGRANGE’S EQUATIONS OF MOTION 255

Solution.
The kinetic and potential energies of the satellite orbiting in an elliptic
orbit are
1 1
T = M (Ṙ2 + θ̇2 R2 ) + [Ix ωx2 + Iy ωy2 + Iz ωz2 ] (5.231)
2 ( 2
µM µ
U =− − (Iyy + Izz − Ixx )[3(cosαcosφ)2 − 1]
R 4R3
+ (Izz + Ixx − Iyy )[3(cosαsinφsinγ − sinαcosγ)2 − 1]
)
+ (Ixx + Iyy − Izz )[3(cosαsinφcosγ + sinαsinγ)2 − 1]

(5.232)

where

ωx = − (θ̇ + α̇)sinφ + γ̇ (5.233)


ωy =(θ̇ + α̇)cosφ sin γ + φ̇cosγ (5.234)
ωz =(θ̇ + α̇)cosφ cos γ − φ̇sinγ (5.235)

The Lagrangian equations of motion corresponding to the generalized


coordinates q1 = α, q2 = φ and q3 = γ are obtained using the general
relation
 
d ∂T ∂T ∂U
− + = 0, k = 1, 2, 3 (5.236)
dt ∂ q˙k ∂qk ∂qk
We substitute the generalized coordinates α, φ, and γ in the preceding equa-
tions and express the derivative with respect to true anomaly θ using the
relations derived next. Referring to Eq.(5.179) and Eq.(5.180), we have

q̇ =q 0 θ̇ (5.237)
00 2 0
q̈ =q θ̇ + q θ̈ (5.238)

Here θ̇ and θ̈ are first derived. The orbital angular momentum h is

h = R2 θ̇ (5.239)

Knowing
h2
= p = a(1 − e2 ) (5.240)
µ
where p is semi-latus rectum, we have
p
h = µa(1 − e2 ) (5.241)
256 CHAPTER 5. DYNAMICS II

Thus, we can write



h µp
θ̇ = 2 = (5.242)
R R2

Differentiating with respect to time, we have

h
θ̈ = −2 Ṙ (5.243)
R3
p
Knowing the orbital equation R = 1 + ecosθ, Ṙ is

p Resinθ
Ṙ = (esinθ)θ̇ = θ̇ (5.244)
(1 + ecosθ)2 1 + ecosθ

Substituting this in the preceding equation, we have

hθ̇esinθ
θ̈ = −2 (5.245)
R2 (1 + ecosθ)

Substituting for θ̇ = h/R2 yields

h2 esinθ µpesinθ µ
θ̈ = −2 = −2 4 = −2 3 esinθ (5.246)
R4 (1+ ecosθ) R (1 + ecosθ) R

Thus, we have
p
µa(1 − e2 ) 0
q̇ = q (5.247)
R2
µp µpesinθ
q̈ =q 00 4 + q 0 4
R R (1 + ecosθ)
µ
= 3 [(1 + ecosθ)q 00 − 2q 0 esinθ] (5.248)
R

Applying the preceding Eqs. (5.247)-(5.248) and replacing the orbital radius
R by semi-major axis a and eccentricity e as per the relation

a(1 − e2 ) µ1/3 (1 − e2 )
R= = 2/3
1 + ecosθ Ω (1 + ecosθ)
p
with Ω = µ/a3 and further carrying out the algebraic manipulations, we
get the following governing nonlinear, coupled ordinary differential equations
of motion of the system:
5.3. LAGRANGE’S EQUATIONS OF MOTION 257

Satellite: Pitch(α)
e1 (cos2 φcos2 γ + kxz sin2 φ + kyz cos2 φsin2 γ)α00
+ e1 (kyz − 1)φ00 sinγcosγcosφ − (1 + ecosθ)kxz γ 00 sinφ
− e2 [(1 + α0 )(cos2 φcos2 γ + kxz sin2 φ + kyz cos2 φsin2 γ)
− φ0 (1 − kyz )sinγcosγcosφ − kxz γ 0 sinφ] + e1 (1 + α0 )
× {−φ0 sin2φcos2 γ − γ 0 cos2 φsin2γ + kxz φ0 sin2φ
+ kyz (−φ0 sin2φsin2 γ + γ 0 cos2 φcos2γ)} + e1 [−φ0 γ 0 cosφcos2 γ
+ φ0 sinγ(φ0 sinφcosγ + γ 0 cosφsinγ) − kxz φ0 γ 0 cosφ
− kyz φ0 γ 0 cosφsin2 γ + kyz φ0 cosγ(−φ0 sinφsinγ + γ 0 cosφcosγ)]
3
− [(kxz + kyz − 1)(cosαsinφcosγ + sinαsinγ)(−sinαsinφcosγ
2
+ cosαsinγ) + (1 + kxz − kyz )(cosαsinφsinγ − sinαcosγ)
× (−sinαsinφsinγ − cosαcosγ)
+ (1 − kxz + kyz )cosαcosφ(−sinαcosφ)] = 0 (5.249)

Satellite: Roll(φ)
e1 (kyz − 1)α00 cosφcosγsinγ + e1 (sin2 γ + kyz cos2 γ)φ00
− e2 [−(kyz − 1)(1 + α0 )cosφcosγsinγ + φ0 (sin2 γ + kyz cos2 γ)]
+ e1 [(1 + α0 )(1 − kyz )(sinφcosγsinγφ0 − cosφcos2γγ 0 )
+ (1 − kyz )φ0 γ 0 sin2γ] + e1 (1 + α0 ){[(1 + α0 )cosφcosγ
− φ0 sinγ]sinφcosγ + kxz [−(1 + α0 )sinφcosφ + γ 0 ]cosφ
+ kyz [(1 + α0 )cosφsinγ + φ0 cosγ]sinφsinγ}
3
− [(kxz + kyz − 1)(cosαsinφcosγ + sinαsinγ)cosαcosφcosγ
2
+ (1 − kxz + kyz )cosαcosφ(−cosαsinφ)
+ (1 + kxz − kyz )(cosαsinφsinγ − sinαcosγ)(cosαcosφsinγ)] = 0
(5.250)

Satellite: Yaw(γ)

−e1 kxz α00 sinφ + e1 kxz γ 00 − e2 kxz [−(1 + α0 )sinφ + γ 0 ]


+ e1 [(1 − kyz )cos2γ − kxz ](1 + α0 )φ0 cosφ
+ e1 (1 − kyz )[(1 + α0 )2 cos2 φ − φ02 ]sinγcosγ
− 3(1 − kyz )(cosαsinφcosγ + sinαsinγ)
× (cosαsinφsinγ − sinαcosγ) = 0 (5.251)

where e1 = 1 + ecosθ and e2 = 2esinθ.


258 CHAPTER 5. DYNAMICS II

5.4 Summary
In this chapter, we have described Euler’s and Lagrange’s equations of motion
and several example problems are presented. The important results discussed
in this chapter are summarized as follows.
5.4. SUMMARY 259

Newton’s Second Law of Motion


d~
p
F~ =
dt
Euler Equation of Motion
~˙ = H
T~ = H ~˙ xyz + ω ~
~ ×H
(for body-fixed x, y, and z principal axes)
Tx = Ixx ω̇x − (Iyy − Izz )ωy ωz
Ty = Iyy ω̇y − (Izz − Ixx )ωz ωx
Tz = Izz ω̇z − (Ixx − Iyy )ωx ωy

Lagrange’s Equation of Motion


d ∂L − ∂L = Q + Pm λ ∂fj , k = 1, 2, . . . , n;
 
dt ∂ q˙k ∂ qk k j=1 j ∂qk
∂~r
Lagrangian function, L = T − U ; Generalized force:Qk = pj=1 F~j · j
P
∂qk
Lagrange multiplier λj corresponding to fj constraint:
fj (q1 , q2 , . . . , qn ) = 0, j = 1, 2, . . . , m
(for U independent of q̇k )
 
d ∂T ∂fj
− ∂T + ∂U = Qk + m
P
dt ∂qk ∂qk j=1 λj ∂qk , k = 1, 2, . . . , n
∂ q˙k
Gravity Force on Dumbbell System
( )
µ 3
F~g = − 3 M R~− 2 2
2R Me L [2(îo · î)î − 5(îo · î) îo + îo ]
R
where M = m1 + m2 ; Me = m1 m2 /(m1 + m2 )

( Gravity Torque on Rigid Body



T~g = − 3 (Iyy + Izz − Ixx )[(îo · î)(îo × î)]
2R
)
+(Izz + Ixx − Iyy )[(îo · ĵ)(îo × ĵ)] + (Ixx + Iyy − Izz )[(îo · k̂)(îo × k̂)]
260 CHAPTER 5. DYNAMICS II

References

1. F. P. J. Rimrott, Introductory Attitude Dynamics, Springer-Verlag, New


York, 1989.

2. William E. Wiesel, Spaceflight Dynamics, McGraw-Hill, New York, sec-


ond edi- tion, 1997.

3. V. A. Chobotov, Spacecraft Attitude Dynamics and Control, Krieger


Publishing Co., Malabar, FL, 1991.

4. Peter C. Hughes, Spacecraft Attitude Dynamics, John Wiley & Sons,


New York, 1986.

Problem Set 5

5.1 Derive the 3-dimensional attitude equations of motion of a dumbbell


system as described in Example 5.2 using Euler’s method and La-
grange’s method.
5.2 Derive the in-plane attitude equations of motion of a given system of
a spacecraft m1 and two bodies m2 and m3 connected through rigid
massless cables of length L1 and L2 (Fig. 5.1).
m3
β2
Orbit L2
Local Vertical
m2
L1 β1
Local Vertical
S m1
Y R

θ
E
X

Figure 5.1: System undergoing in-plane libration.


5.4. SUMMARY 261

5.3 Derive the in-plane translational (orbital) and rotational (attitude) mo-
tion of the dumbbell system in Problem 5.1 using Newton’s method and
Lagrange’s method.
5.4 Derive the three-dimensional translational (orbital) and rotational (at-
titude) motion of a dumbbell system using Newton’s method and La-
grange’s method.
262 CHAPTER 5. DYNAMICS II
Chapter 6

Mathematical Analysis
and Simulation

This chapter deals with the mathematical analysis and numerical simulation
of the system equations of motion derived in Chapters 4 and 5. In mathe-
matical analysis we focus on equilibrium analysis, followed by linearization
of the system equations of motion about equilibrium states. Next, stability
analysis is explained, followed by derivation of closed-form solutions. Finally,
numerical methods to solve system nonlinear equations of motion and the use
of Matlab and Maple are explained.

6.1 Introduction
To understand the dynamics of a given system with regard to variations in
system parameters and external disturbances, it is important to first analyze
the given system analytically and then verify the analytical results with nu-
merical simulations. There are several methods to analyze the given system.
Here we focus on the linear system analysis as the theory for a linear system
is well developed and understood. Also, it is easier for beginners to under-
stand linear system analysis first and then read theory on nonlinear systems.
Furthermore, several softwares are now available to study the kinematics and
dynamics of a system. These softwares include Matlab, Maple, and Mathe-
matica. Maple and Mathematica are quite similar. We will only explain the
use of Maple in this chapter, in addition to Matlab.
To begin with linear system analysis, we first study the equilibrium con-
ditions of a given system. Next we explain the procedure to derive linear
equations of motion for a given system nonlinear equations of motion. Sta-
264 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

bility analysis followed by a closed-form solution is explained next. Then, the


numerical simulation of a given nonlinear system is studied. Several meth-
ods of numerical integration and the use of Matlab and Maple are illustrated
through examples.

6.2 Equilibrium Condition


A general nonlinear system with a state vector X can be written as

Ẋ = f (X, t) (6.1)

If the function f (X, t) does not depend explicitly on time t , then the sys-
tem is said to be autonomous. Otherwise the system is called nonautonomous.
For a feedback control system with control input vector u = g(X, t), we can
express the system by

Ẋ =f (X, u, t) (6.2)
Y =h(X, u) (6.3)
Here Y is the system output vector.
For a given state of the system, say X = Xe , at t = t0 if we have

f (Xe , t) = 0, ∀t > t0 (6.4)

then we call such states Xe as equilibrium states of the system. As per


Eq. (6.4), we have Ẋe = Ẍe = 0 and Xe may not be zero. If Xe = 0, Xe is
called the null solution of the system (6.1). For the case in which Ẍe = 0,
i.e., the force and/or torque acting on the system is zero, but Ẋe 6= 0, we call
the system in steady state motion.
In the case of the feedback control system, we have the following condition
of equilibrium

f (Xe , ue ) = 0. (6.5)
Here ue is a input vector at the equilibrium state of the system.
While using the Lagrangian method in deriving the equations of motion
of the system, we may use the equilibrium condition as
∂(V − T0 )
= 0, k = 1, 2, . . . , n (6.6)
∂qk
where T0 represents the part of T independent of the velocities.
6.3. LINEARIZATION 265

6.3 Linearization
As a first step in examining a nonlinear system, we linearize the system about
the equilibrium state discussed previously. The linerized system obtained
from the nonlinear system (6.1)

δ Ẋ = AδX (6.7)

and for the system (6.2-6.3), the linearized system is

δ Ẋ =AδX + Bδu (6.8)


δY =CδY + Dδu (6.9)

∂f
where, δX = X − Xe , δu = u − ue , δY = Y − Ye , A = ,
∂X (Xe ,ue )
∂f
B= , C = ∂h , D = ∂h .
∂u (Xe ,ue ) ∂X (Xe ,ue ) ∂u (Xe ,ue )
The eigenvalues of the system can be found by solving the characteristic
equation given by

|A − sI| = 0 (6.10)

The eigenvalues and eigenvectors of the system can be found using Matlab
as
[eigv,eign]=eig(A)
where eigv and eign stand for eigenvectors and eigenvalues, respectively.

6.4 Stability
Next, we study the stability of the linear system. There are various definitions
of the stability. We consider the following two definitions:

(a) the ability of the system to return to the equilibrium state after an
arbitrary displacement away from it.
(b) the ability of the system to produce a bounded output for any bounded
input.

For a linear time-invariant system that has a characteristic equation with


constant coefficients given as
266 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

an sn + an−1 sn−1 + · · · + a1 s + a0 = 0 (6.11)

both definitions of stability are equivalent, but in the case of nonlinear or


time varying systems, these definitions are distinct: a system may possess one
kind of the stability while the other may be absent. We examine the linear
time-invariant system. From the roots of the system characteristic Eq.(6.11),
the following conclusions can be made about its stability:

1. If Re(si ) < 0 for all i, then the system is asymptotically stable. Here
Re(si ) stands for the real part of si .

2. If Re(si ) > 0 for some i, then the system is unstable.

3. If Re(si ) = 0 for some i = j, and

(a) si is a simple root for all such j, then the system is marginally
stable, but not asymptotically stable.
(b) sj is a multiple root for some such j, the system is unstable.

In order to find the roots of the characteristic equation, we can use rou-
tines such as the Jenkins-Traub algorithm or Laguerre algorithm. However,
we are not interested in the actual values of the roots of the characteris-
tic equation, but in how the system parameters affect the locations of the
roots. Computational root finding algorithms do not provide such informa-
tion. Approaches such as Routh-Hurwitz criterion, Nyquist criterion,
and Bode diagram do not involve root finding and they are used frequently
for checking the stability of the system. Here, we will discuss Routh-Hurwitz
Criterion only. Using this approach, we can obtain conditions relating sys-
tem parameters to the system stability and in some cases, it is quite easy to
state whether the system is stable or not. The Routh-Hurwitz Criterion is
as follows:
The necessary and sufficient conditions that all roots of Eq. (6.11) to lie
in the left half of the s-plane are

(a) all the coefficients (aj , j = 0, 1, . . . , n) have the same sign. Let it be
aj > 0, j = 0, 1, . . . , n.

(b) aj 6= 0, j = 0, 1, . . . , n.
6.4. STABILITY 267

(c) Determinants,

D1 =an−1

an−1 an−3
D2 =
an an−2

an−1 an−3 an−5


D3 = an an−2 an−4
0 an−1 an−3
···

an−1 an−3 an−5 ··· 0


an an−2 an−4 ··· 0
Dn = 0 an−1 an−3 ··· 0 (6.12)
.. .. .. ..
. . . .
0 0 0 ··· a0

called Hertwitz determinants, where the coefficients with indices larger


than n or negative are replaced with zeros, must satisfy the conditions

Dj > 0, j = 0, 1, . . . , n (6.13)

To verify the conditions (6.13), we can use Routh’s table as explained


below. There are n+1 rows and n/2+1 columns. The rows are arranged
in decreasing order starting with the nth row

sn Row : bn1 bn2 bn3 ···


n−1
s Row : b(n−1)1 b(n−1)2 b(n−1)3 ···

and ending at row 0. Here,

bij = a[i−2(j−1)] , i = n − 1, n; j = 1, 2, . . . , n/2 + 1 (if n is even);


j = 1, 2, . . . , (n + 1)/2 (if n is odd)

or,

bn1 = an , bn2 = an−2 , bn3 = an−4 , ···


b(n−1)1 = an−1 , b(n−1)2 = an−3 , b(n−1)3 = an−5 , · · ·
268 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

The other rows of Routh’s table are obtained using the relation
b(i+1)i b(i+2)(j+1) − b(i+1)(j+1) b(i+2)i
si Row : bij = ,
b(i+1)i
i = n − 2, n − 1, . . . , 0;
j = 1, 2, . . . , n/2 + 1 (if n is even);
j = 1, 2, . . . , (n + 1)/2 (if n is odd)

For the stability of the system, all the elements of the first column of
Routh’s table should have same sign i.e.,

bi1 > 0, i = 0, 1, . . . , n (6.14)

The number of changes of signs of bi1 , for all i, equals the number of
roots with positive real parts or in the right half s-plane.
Remarks.
1. If an element in the first column bi1 = 0, then the above procedure
for Routh’s tabulation cannot continue. In this situation, write
bi1 = , where  is a small positive number, and then complete
Routh’s table as suggested. However, this procedure may not give
correct results if the characteristic Eq. (6.11) has pure imaginary
roots.
2. If in a row, say row k, bkj = 0, for all j, then use the following
procedure:
Write an auxiliary equation from the coefficients of (k − 1)th row
of Routh’s table as

A(s) = b(k−1)1 sk−1 + b(k−1)2 sk−3 + b(k−1)3 sk−5 + · · · (6.15)

and take the derivative (dA(s)/ds). Replace the kth row with the
coefficients of (dA(s)/ds). The roots of the auxiliary Eq.(6.15) also
satisfy the characteristic Eq. (6.11) and thus, by solving the aux-
iliary Eq.(6.15), we get the roots of the characteristic Eq. (6.11).
The auxiliary Eq.(6.15) always has even powers of s. Therefore,
the roots of the auxiliary Eq.(6.15) all lie on the imaginary axis
in the s-plane. Such systems having pure imaginary roots are
marginally stable.

Hessian matrix. Let f (X) be a function of n variables such that f (X) ∈


C 2 , C 2 denotes the space of all real functions whose second order partial
6.4. STABILITY 269

derivatives are continuous. Then, the Hessian matrix of f (X) is a n × n


symmetric matrix of second order partial derivatives defined by

∂2f ∂2f ∂2f


 
 ∂x2 ···
1
∂x1 ∂x2 ∂x1 ∂xn 
2 2
∂2f 
 
 ∂ f ∂ f
···
 ∂x ∂x ∂x22 ∂x2 ∂xn 
 
H(X) =  2 1
.. .. ..

 

 . . ··· . 

 ∂2f 2
∂ f 2
∂ f 
···
∂xn ∂x1 ∂xn ∂x2 ∂x2n

Since f (X) ∈ C 2 , it follows that ∂ 2 f /∂xi ∂xj = ∂ 2 f /∂xj ∂xi ∀ i, j ∈ {1, 2, . . . , n}.
1. Matrix minor test. For any square matrix of order n, a principal
matrix minor means any submatrix of order m(≤ n) which contains the first
m elements of the principal diagonal. These principal matrix minors decide
the nature of f (X). This test is also known as Sylvester’s criterion. Thus,
f (X) is

(a) positive definite, if all the principal minor determinants of A are posi-
tive,
a11 a12
D1 = a11 > 0, D2 = > 0,
a21 a22

a11 a12 a13


D3 = a21 a22 a23 > 0, · · · , Dn > 0;
a31 a32 a33

(b) positive semi-definite, if D1 > 0, Di ≥ 0, at least one of the Di = 0 for


i = 1, 2, . . . , n;

(c) negative definite, if D1 < 0, D2 > 0, D3 < 0, · · · , (−1)i Di > 0, i =


2, 3, . . . , n;

(d) negative semi-definite, if D1 < 0, D2 ≥ 0, D3 ≤ 0, · · · , (−1)i Di ≥ 0,


and at least one of the Di = 0 for i = 2, 3, . . . , n;

(e) indefinite, if none of the above cases happens.

2. Eigenvalue test. Since matrix A is a real symmetric matrix in f (X),


it follows that its eigenvalues are real. Then f (X) is

(a) positive definite, if si > 0, i = 1, 2, . . . , n;


270 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

(b) is positive semi-definite, if si ≥ 0, and at least one of si = 0, i =


1, 2, . . . , n;

(c) negative definite, if si < 0, i = 1, 2, . . . , n;

(d) negative semi-definite, if si ≤ 0, and at least one of si = 0, i =


1, 2, . . . , n;

(e) indefinite, if A has both positive and negative eigenvalues.


We are required to conclude the behavior of the nonlinear system Eq.
(6.1) from the analysis of the linear system. The conclusions are

(a) If the linear system is asymptotically stable, then the nonlinear


system is locally asymptotically stable.
(b) If the linear system is unstable, the nonlinear system is also un-
stable.
(c) If the linear system is marginally stable, we can not say anything
about the stability of the nonlinear system as the neglected sec-
ond or higher order terms in the nonlinear system may render
instability of the nonlinear system.

6.5 Closed-form Solution


The equation of motion of a system can be written in the form

Ẋ = f (X, t) (6.16)

where X is the state vector and t is the time. The closed-form solution of
this equation (6.16) can be of the form

X = g(X0 , t) (6.17)

where X0 is the initial state vector (i.e., X = X0 at t = t0 ). In the case


the system described by Eq. (6.16) is a nonlinear system, except for a few
nonlinear systems, it is difficult or impossible to find the closed-form solution.
On the other hand, if the system is a linear system expressed in the form

Ẋ = AX (6.18)

it has a closed-form solution

X = X0 eAt (6.19)

where A may be a constant or time varying.


6.5. CLOSED-FORM SOLUTION 271

Most of the systems dealt in the previous chapters are second order sys-
tems. A second order system equation of motion can be written as
aẌ + bẊ + cX = 0 (6.20)
The closed-form solution is
X(t) = C1 es1 t + C2 es2 t (6.21)
where

−b ± b2 − 4ac
s1,2 = (6.22)
2a
are the system eigenvalues obtained from the system characteristic equation
as2 + bs + c = 0 (6.23)
and C1 and C2 are constants determined from initial conditions. Depending
upon the eigenvalues s1 and s2 , the general solution of the system is presented
in Table 6.1.

Table 6.1: General solution of Eq. (6.20) depending upon eigenval-


ues.

Eigenvalues General solution X

Distinct real
s 1 , s2 X = C1 es1 t + C2 es2 t
Real double
s1 = s2 = −α X = (C1 + C2 t)e−αt
Complex conjugate
s1 = −α + jω,s2 = −α − jω X = (C1 sinωt + C2 cosωt)e−αt

A second order system can be written in the standard form


ẍ + 2ζωn Ẋ + ωn2 x = 0 (6.24)
where ωn and ζ are the natural frequency and damping ratio of the system,
respectively. The solution of this standard second order system is presented
in Table 6.2.
The equations of motion can be of the type as follows:
aẌ + bẊ + cX = F (t) (6.25)
272 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

Table 6.2: Solution for the Standard Second Order Eq. (6.24).

0< ζ < 1 X = Ce−ζωn t sin(ωd t + φ)


p
where ωd = ωn 1 − ζ 2
ẋ0 + x0 ζωn 2 1/2
h   i
C = x20 + ω d
h x0 ωd i
φ = tan−1
ẋ0 + x0 ζωn
ζ =1 X = (C1 + C2 t)e−ζωn t
where C1 = x0 , C2 = ẋ0 + x0 ωn
ζ >1 X = (C1 eωt + C2 e−ωt )e−ζωn t
p
where ω = ωn ζ 2 − 1
ẋ + x ζω
h i
C1 = 12 x0 + 0 ω 0 n
ẋ + x ζω
h i
C2 = 12 x0 − 0 ω 0 n

where F (t) on the right hand side represent a forcing term. The solution of
this type of equation is the sum of the homogeneous solution, Xh (i.e., when
F = 0) and the particular solution, Xp :

X = Xh + Xp (6.26)

The choice of Xp depending upon F is presented in Table 6.3.

Table 6.3: Choice of Xp for solution of Eq. (6.25).

Term in F Choice of Xp

pekt P ekt
ptn P0 + P1 t + P2 t2 + . . . + Pn tn
psinωt P1 sinωt + P2 cosωt
pcosωt P1 sinωt + P2 cosωt
pekt sinωt ekt (P1 sinωt + P2 cosωt)
pekt cosωt ekt (P1 sinωt + P2 cosωt)
6.5. CLOSED-FORM SOLUTION 273

Example 6.1
Derive the equilibrium and stability conditions of a dumbbell system un-
dergoing inplane librational motion. The equation of motion of the system
is given by

β 00 + 3sinβcosβ = 0 (6.27)

where ()0 denotes derivative with respect to reference angle θ.


Solution.
The system equation of motion is given by

β 00 + 3sinβcosβ = 0 (6.28)

Integrating Eq. (6.28) with an initial β 0 =β00 yields

β 02 + 3sin2 β = C = β002 (6.29)

where C, a constant of integration is determined from initial conditions:


t = 0, β=β0 =0, β 0 =β00 as C = β002 .
As per Eq.(6.29), β 0 reaches minimum at β = π/2. Thus, for the system
to librate about the local vertical,


β00 < 3. (6.30)

In the case β00 > 3, the system will have rotational motion.
Linearized Equation of Motion:
Let us consider a reference state: β = βr , β 0 = βr0 , β 00 = βr00 . Substituting
the reference state in Eq.(6.28), we get

3
βr00 + sin2βr = 0 (6.31)
2

Now we perturb the reference state and the perturbed state is given by:
β = βr + δβ, β 0 = βr0 + δβ 0 , β 00 = βr00 + δβ 00 . Putting the perturbed state in
Eq.(6.28), we have the following equation

3
βr00 + δβ 00 + sin(2βr + 2δβ) = 0 (6.32)
2
274 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

Subtracting Eq.(6.32) from Eq.(6.31), we get

3
δβ 00 + [sin(2βr + 2δβ) − sin2βr ] = 0 (6.33)
2
or,

δβ 00 + 3cos(2βr + δβ)sinδβ = 0 (6.34)

We carry out first order perturbation, i.e., ignore second and higher order
terms of the perturbed variable δβ. Then we get the resulting equation

δβ 00 + 3cos(2βr )δβ = 0 (6.35)

This is a linearized equation. We can obtain this linearized Eq.(6.35) by


simply applying differentiation as

δβ 00 + Aδβ = 0 (6.36)

where A = ∂ [ 3 sin2β] = 3cos(2βr )


∂β 2 βr

Let βr represents an equilibrium state of the system i.e., βr0 = 0, βr00 = 0.


We write βr = βe . Now, we examine the stability of the equilibrium state.
Let βe =0. Substituting this in Eq.(6.35), we get

δβ 00 + 3δβ = 0 (6.37)
√ √
The roots of the characteristic equation are s1 = j 3, s2 = −j 3. Since
the roots are complex conjugate, the system is marginally stable. Hence-
forth, we will call “marginally stable” as “stable” in this chapter, as in most
cases the system may not have damping forces and so the oscillations of the
perturbed variable never die asymptotically.
Next, we take βe = π/2. The Eq.(6.35) is

δβ 00 − 3δβ = 0 (6.38)
√ √
Here the roots of the characteristic equation are s1 = 3 and s1 = − 3.
As one of the roots is positive, the system is unstable. Thus, we can say
6.5. CLOSED-FORM SOLUTION 275

that a dumbbell initially aligned along the local vertical is stable while if it
is aligned along the local horizontal initially, it is unstable.
We find the solution of Eq.(6.37) with the initial conditions: δβ = δβ0 ,
δβ 0 = δβ00 at θ = 0, as

δβ = Asin(ωθ + ψ) (6.39)

where A = [δβ02 + (δβ00 /ω)2 ]1/2 , ψ = tan−1 [(δβ0 ω)/δβ00 ], ω = 3
The time period of librational motion Tp of the dumbbell in dimensionless
form is
2π 2π
Tp = =√ (6.40)
ω 3
in and in dimensional form, is


Tp = √ sec (6.41)
3Ω

Here Ω is the orbital angular velocity. As per Eq. (6.40) or Eq.(6.41), the
time period of libration Tp is independent of the two masses (m1 , m2 ) and
the connecting length between masses L. However, if we write Tp in seconds
as in Eq.(6.41), then it depends on Ω. Thus, as the distance R increases Tp
also increases and√vice versa. We have to understand that this solution is
valid until δβ00 < 3.
Example 6.2
The 3-dimensional attitude motion of a rigid satellite is defined by its
equations of motion (Ref. Chapter 5: Eqs. (5.210-5.212)) as follows:
Satellite: Pitch(α)

α00 cosφcosγ − φ00 sinγ − (1 − kxz + kyz )(1 + α0 )φ0 sinφcosγ


− (1 + kxz − kyz )[(1 + α0 )γ 0 cosφsinγ + φ0 γ 0 cosγ]
+ (kxz − kyz )(1 + α0 )2 sinφcosφsinγ
+ 3(kxz − kyz )(cosαsinφsinγ − sinαcosγ)cosαcosφ = 0

Satellite: Roll(φ)

kyz α00 cosφsinγ + kyz φ00 cosγ − (1 − kxz + kyz )(1 + α0 )φ0 sinφsinγ
− (1 − kxz − kyz )[(1 + α0 )γ 0 cosφcosγ − φ0 γ 0 sinγ]
+ (1 − kxz )(1 + α0 )2 sinφcosφcosγ
+ 3(1 − kxz )(cosαsinφcosγ + sinαsinγ)cosαcosφ = 0
276 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

Satellite: Yaw(γ)

−kxz α00 sinφ + kxz γ 00 + [(1 − kyz )cos2γ − kxz ](1 + α0 )φ0 cosφ
+ (1 − kyz )[(1 + α0 )2 cos2 φ − φ02 ]sinγcosγ
− 3(1 − kyz )(cosαsinφcosγ + sinαsinγ)
× (cosαsinφsinγ − sinαcosγ) = 0

Determine the followings:

(a) Equilibrium state.


(b) Linear equations of motion with respect to null states.
(c) Stability conditions.
(d) Closed-form solution of the linear system.

Solution.

(a) Equilibrium state.


From the given Euler equations of motion, we get the equilibrium state
vector Xe = 0, i.e.,

αe = φe = γe = α0e = φ0e = γe0 = 0 (6.42)

(b) Linear equations of motion with respect to null states.


To derive the linear system, we perturb the state X of the system about
the equilibrium state Xe as

X = Xe + δX (6.43)

Using relations (6.43) and (6.42), and considering first order approxi-
mation for the system state, we have the linearized equations of motion
are as follows:
Satellite: Pitch(α)

δα00 − 3(kxz − kyz )δα = 0 (6.44)

Satellite: Roll(φ)

kyz δφ00 − (1 − kxz − kyz )δγ 0 + 4(1 − kxz )δφ = 0 (6.45)


6.5. CLOSED-FORM SOLUTION 277

Satellite: Yaw(γ)

kxz δγ 00 + (1 − kxz − kyz )δφ0 + (1 − kyz )δγ = 0 (6.46)

Alternatively, in matrix form we can write

X 0 = AX (6.47)

where
   


 δα 
 0 1 0 0 0 0
 3(k2 − k1 )

 
  
δα0  0 0 0 0 0

 
 
 1 − k1 k2

   

 
 
δφ 
   
  0 0 0 1 0 0 
X= , A= 
δφ0 
 








 0 0 −4k1 0 0 k1 − 1

 
δγ 
   








 0 0 0 0 0 1 
 
0 
δγ 

 0 0 0 1 − k2 −k2 0
(6.48)

Here kxz and kyz are replaced by k1 = (Iz −Ix )/Iy and k2 = (Iz −Iy )/Ix
as Kxz = Ix /Iz = (1 − k1 )/(1 − k1 k2 ) and kyz = Iy /Iz = (1 − k2 )/(1 −
k1 k2 ).
Remark. The same linear equations of motion are obtained if we
linearize Lagrange Eqs.(5.217-5.221).

(c) Stability conditions.


For stability analysis, we are required to find the system eigenvalues.
From Eq. (6.47), the pitch α equation is decoupled from the couples roll
and yaw equationsṪherefore, we can analyze the stability of pitch mo-
tion and the stability of roll and yaw motions separately. The stability
condition for the pitch motion can be written as

3(k2 − k1 )
< 0 ⇒ k1 > k2 (6.49)
1 − k1 k2
as 1 − k1 k2 > 0 since k1 < 1 and k2 < 1 by definition.
To analyze the eigenvalues for the roll and yaw motions, we rewrite the
equations of roll and yaw motions
278 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

Y 0 = BY (6.50)

where
   


 δφ  
 0 1 0 0

 
  
 δφ0 
  −4k1

0 0 k1 − 1

Y = , B=  (6.51)
 δγ 
   0

0 0 1 


 
  
 
 δγ 0 
  0 1 − k2 −k2 0

We obtain the characteristic equation by solving

−s 1 0 0
−4k1 −s 0 k1 − 1
|B − sI| = =0 (6.52)
0 0 −s 1
0 1 − k2 −k2 −s

as

s4 + (1 + 3k1 + k1 k2 )s2 + 4k1 k2 = 0 (6.53)

Now, applying Routh-Hurwitz criterion to Eq. (6.53) we can write the


conditions of stability as

1 + 3k1 + k1 k2 > 0
k1 k2 > 0 (6.54)
2
(1 + 3k1 + k1 k2 ) − 16k1 k2 > 0

Using Eqs. (6.49) and (6.54), we can write the conditions of pitch, roll
and yaw stability as
Lagrange region:

k1 > k2 , k1 > 0, k2 > 0 (6.55)

where 1 + 3k1 + k1 k2 > 0 and (1 + 3k1 + k1 k2 )2 − 16k1k2 > 0 are already


satisfied. In terms of the dimensional form,i.e., Ix , Iy and Iz , the above
stability condition is
6.5. CLOSED-FORM SOLUTION 279

Iz > Iy > Ix (6.56)

Note that this region is stable even in presence of damping.


DeBra-Delp region:

k1 > k2 , k1 < 0, k2 < 0, (1 + 3k1 + k1 k2 )2 − 16k1 k2 > 0 (6.57)

where (1 + 3k1 + k1 k2 ) > 0 is already satisfied. Note this region is


unstable in presence of damping.

(d) Closed-form solution.

We derive the solution of the linear system Eq. (6.47). The solution of the
pitch α motion can be written as

δα = Csin(ωα θ + ϕα ) (6.58)

2 0 2 1/2 −1 0
where
p C = [δα(0) + (δα (0)/ωα ) ] , ϕα = tan [(δα(0)ωα )/δα (0)],
ωα = 3(k1 − k2 )/(1 − k1 k2 ). The time period of pitch libration is
q
Tp = 2π/ωα = 2π/ 3(kyz − kxz ) (6.59)

To obtain the solution of the roll and yaw motions, we assume δφ = Desθ
and δγ = Eesθ (as roll and yaw are coupled). Substituting these into the
equations for roll and yaw motions Eqs. (6.50), we get

esθ [As2 + (1 − k1 )Bs + 4k1 A] = 0 (6.60)


sθ 2
e [Bs + (k2 − 1)As + k2 B] = 0 (6.61)

Considering the fact that esθ 6= 0, we have

E s2 + 4k1
= (6.62)
D (k1 − 1)s

Assuming the solutions of the roll and yaw motions are periodic, i.e., s =
jω and thereby substituting it in Eq. (6.62), we get
280 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

−ω 2 + 4k1
λ= (6.63)
(k1 − 1)ω

where λ = Ej/D
Thus, we can write the solution as

δφ = Desθ (6.64)

δγ = −jDλe (6.65)

In fact, the system involving roll and yaw motions are two degrees of free-
dom of motion and thus have two frequencies of oscillations. The frequency
equation of the system are obtained from Eq. (6.53) by putting s = jω

ω 4 − (1 + 3k1 + k1 k2 )ω 2 + 4k1 k2 = 0 (6.66)

Solving, we get

 1/2 " s #1/2


1 + 3k1 + k1 k2 16k1 k2
ω1,2 = 1∓ 1− (6.67)
2 (1 + 3k1 + k1 k2 )2

Thus, the general solution of the equations of motion is composed of two


harmonic motions of the frequencies ω1 and ω2 ; they are the fundamental
and first harmonic. The solution can be expressed as

δφ = D1 sin(ω1 θ + ψ1 ) + D2 sin(ω2 θ + ψ2 ) (6.68)


δγ = D1 λ1 sin(ω1 θ + ψ1 ) + D2 λ2 sin(ω2 θ + ψ2 ) (6.69)

where

−ω12 + 4k1 −ω22 + 4k1


λ1 = , λ2 = (6.70)
(k1 − 1)ω1 (k1 − 1)ω2

and the four constants D1 , D2 , ψ1 and ψ2 are to be evaluated by the


four initial conditions δφ(0), δφ0 (0), δγ(0), δγ 0 (0). For the case when φ(0) =
γ(0) = 0, we obtain the initial conditions
6.5. CLOSED-FORM SOLUTION 281

ψ1 = ψ2 = 0
D2 = [φ0 (0)λ1 − γ 0 (0)]/[ω2 (λ1 − λ2 )]
D1 = [−D2 ω2 + φ0 (0)]/ω1

Example 6.3
The pitch equation of motion of a rigid satellite in an elliptic orbit is given
by

(1 + ecosθ)α00 − 2esinθ(1 + α0 ) − 3(kxz − kyz )sinαcosα = 0 (6.71)

Derive the linear equation of motion assuming low eccentricity and first
order perturbation of the system state variables. Also, obtain the closed-form
solution of the linear system.
Solution.
Considering small amplitude librations and low eccentricities, and ignor-
ing the second and higher order terms in α, α0 , and e, the resulting equation
of motion of the satellite is

α00 − 3(kxz − kyz )α = 2esinθ (6.72)

The solution of the linear system is

−2esinθ
α = Asin(ωα θ + ϕ) + (6.73)
1 + 3k3

where A = [α20 + (α00 /ωα )2 ]1/2 , ϕ = tan−1 [(α0 ωα )/α00 ], ωα = 3k3 , and
k3 = kyz − kxz . Note here the first term is the homogeneous solution while
the second term corresponds to particular solution of the system. When we
compare the solutions of the linear and nonlinear systems, we find that the
linear solution follows the nonlinear solution until e < 0.01. In order to
obtain more accurate solution, we can assume α for a particular solution as

X mθ
α= Am,n sin (n = 1, 2, . . .) (6.74)
m=1
n

2π 2πn
where period is T = ω = m . For an illustration, we can consider

α = A1,1 sinθ + A2,1 sin2θ + A3,1 sin3θ (6.75)


282 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

The periodic solution exists upto e = 0.3. Referring to the pitch equation of
motion for e=0, we have

α00 + 3k3 sinαcosα = 0 (6.76)

Integrating,

α02 + 3k3 sin2 α = Cc (6.77)

To have a periodic solution:

Cc < −3k3 (6.78)

and

α(0) = 0, α0 (0) = (−3k3 )1/2 sinαmax (6.79)

6.6 Numerical Simulation


The equations of motion of a given system derived in Chapters 4 and 5 are
in general highly nonlinear, coupled differential or partial differential equa-
tions. The linearized equations obtained with suitable approximations in the
previous section can only have closed-form solutions. However, to under-
stand the underlying dynamics of the nonlinear systems we need to solve
them using numerical integration. There are several methods of numerical
integration depending upon the types of particular problems and associated
differential equations (DEs). The DEs can be of the types: ordinary dif-
ferential equations (ODEs), differential-algebraic equations (DAEs), partial
differential equations (PDEs) (including parabolic and elliptic partial differ-
ential equations), and delay differential equations (DDEs). The ODEs have
the following forms:
Explicit ODEs: Ẋ = f (X, t)
Linearly implicit ODEs: M (X, t)Ẋ = f (X, t) where M (X, t) is a matrix.
Fully implicit ODEs: f (Ẋ, X, t) = 0
The ODE solvers in general accept only first-order differential equations.
Any ordinary differential equation

X n = f (X, Ẋ, . . . , X (n−1) , t) (6.80)


6.6. NUMERICAL SIMULATION 283

can expressed as an equivalent system of 2n first-order ODEs of the form

Ẋ = f (X, t) (6.81)

We basically have two types of problems: initial value problems (IVPs)


and boundary value problems (BVPs). In IVPs, initial state conditions are
given, whereas in BVPs the initial and final state conditions are known. The
ODE solvers for IVPS are summarized in Table 6.1.

Table 6.1 ODE Initial Value Problem Solvers.


Solver Types of Problems Method Order
ode45 Nonstiff DEs1 Runge-Kutta Medium
ode23 Non-stiff DEs Runge-Kutta Low
ode113 Non-stiff DEs Adams Variable
2 3
ode15s Stiff DEs NDFs or BDFs Variable
and DAEs
ode23s Stiff DEs Rosenbrock Low
ode23t Moderately stiff Trapezoidal rule
ODEs and DAEs
ode23tb Stiff DEs TR-BDF2 Low
ode15i Fully implicit DEs BDFs Variable

1
Differential Equations includes ordinary and partial differential equations.
2
Numerical differentiation formulas. 3 Backward differentiation formulas also
known as Gear’s method.
Remarks. If ode45 failed or was inefficient, apply ode15s.
Example 6.3
Simulate a dumbbell system undergoing inplane librational motion given
by the equation of motion
µ
β̈ + 3 3 sinβcosβ = 0
r
where µ is the Earth’s gravitational constant and r is the orbital radius.
Assume r=6878 km and initial β0 =-90 deg, 30 deg, 90 deg and β̇0 =0.01Ω,
0.1Ω, -0.1Ω (Ω = µ/r3 ), respectively. Compare this simulation with the
closed-form solution of the corresponding linear system obtained in Example
6.1.
284 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

% Main File Name dumbbell inp.m


% Example 6.3: Dumbbell System (Inplane Motion)
% Global Variables Declaration
global mu r
% Orbit Parameters
mu=398500; % Earth’s Gravitational Parameter (km)
r=6878; % Orbital Radius (km)
thetadot=sqrt(mu/rˆ3); % Orbital angular velocity (rad/s)
deg2rad=pi/180; % Degree to Radian Conversion
% Initial conditions of state Vector
x01=-90*deg2rad; % Initial beta (rad)
x02=thetadot*0.01; % Initial beta dot (rad/s)
x0=[x01;x02]; % Initial state vector
tstart=0; % Simulation start time (in seconds)
nor=2; % Simulation end time (in orbits)
tperiod=2*pi/thetadot; % Orbital time period (sec)
tend=2*pi*nor/thetadot; % Simulation end time (in seconds)
tstep=1; % Time Step
t=tstart:tstep:tend; % Simulation Time
% Analytical Solution
omega=sqrt(3)*thetadot % System natural frequency (rad/s)
A=sqrt(x01ˆ2+(x02/omega)ˆ2); % Magnitude
phase=atan(x01*omega/x02); % Phase
beta anal=A*sin(omega*t+phase) % beta analytical
betadot anal=A*omega*cos(omega*t+phase) % beta dot analytical
% Numerical Solution
fname=’dumbbell inp sub’; % Name of the rhs File (.m is assumed)
options=odeset(’abstol’,1e-9,’reltol’,1e-9); % Tolerance
[t,x]=ode45(fname,t,x0); % Numerical integration using ode45
6.6. NUMERICAL SIMULATION 285

% Plots
figure
zoom on
theta=thetadot*t;
subplot(2,1,1),plot(theta/(2*pi),x(:,1)/deg2rad,theta/(2*pi),beta anal/deg2rad);
set(findobj(gca,’Type’,’line’,’Color’,[0 0 1]),’LineWidth’,1)
set(findobj(gca,’Type’,’axes’),’Fontsize’,12)
xlabel(”);
ylabel(’\fontsize{16}\alpha (deg)’);

subplot(2,1,2),plot(theta/(2*pi),x(:,2)/deg2rad,theta/(2*pi),betadot anal/deg2rad);
set(findobj(gca,’Type’,’line’,’Color’,[0 0 1]),’LineWidth’,1)
set(findobj(gca,’Type’,’axes’),’Fontsize’,12)
xlabel(’\fontsize{16}Orbits’);
ylabel(’\fontsize{16}\alpha dot (deg/s)’);

% Subroutine for Example 6.3


% Subroutine File Name dumbbell inp sub.m
function xdot=rhs(t,x)
% Global Variables Declaration
global mu r
% Equations in First Order Form
xdot1=x(2); % beta dot (derivative of x(1))
xdot2=-3*(mu/rˆ3)*sin(x(1))*cos(x(1)) % beta ddot (derivative of x(2))
xdot=[xdot1;xdot2]; % xdot vector
286 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

6.7 Maple
Like Matlab, Maple is another powerful tool that can be used for several
tasks including

(a) Deriving system position, velocity, and acceleration (as explained in


Chapter 2),
(b) Deriving system momentum, kinetic energy, and potential energy (as
explained in Chapter 2),
(c) Deriving system equations of motion (as explained in Chapter 5),
(d) Simulating the system dynamics.

We will explain the usage of Maple through several examples.


Example 6.4
For a given position vector of a system,

RI = [R1 cosθ, R1 sinθ, 0] (6.82)

in the inertial frame. Find the velocity and acceleration vectors of the
system in the inertial frame using Maple. Assume R1 and θ vary with time
and they represent the radial position and the angle with reference to the
inertial X axis, respectively.
Solution.
The velocity and acceleration vectors of the system are derived using
Maple program (see Fig. 6.1). VI and AI stand for the inertial velocity and
acceleration vectors, respectively.
6.7. MAPLE 287

2 3
2 3
0 1
ª
«¬ 0 1 0 10 1
¹
0 1 º»

0 1
ª§ ¶
Ǭ
©
·
¸
0 1 0 1
§ ¶ § ¶
0 1 ¨
©
· ¨
¸ ©
·
¸

0 1 0 1 0 1
§ ¶ ¹
¨ · º
© ¸ »
0 1
ª§ ¶
Ǭ
©
·
¸
0 1 0 1
§ ¶ § ¶
0 1 ¨
©
· ¨
¸ ©
·
¸

0 1 0 1 0 1
§ ¶ ¹
¨ · º
© ¸ »

Figure 6.1: Maple program for Example 6.4

Example 6.5
For a given position vector of a system,
R1 = [h, 0, 0] (6.83)
in the rotating, non-inertial frame. Find the velocity and acceleration vectors
of the system in the rotating, non-inertial frame using Maple. The rotating,
non-inertial frame is obtained from the inertial frame by rotation of θ angle
about the z-axis. Here h represents the radial position.
Solution.
The velocity and acceleration vectors of the system are derived using
Maple program (see Figs. 6.2-6.3). V 1 and A1 represent the velocity and
acceleration vectors in the rotating frame (called frame 1), respectively. V 2
and A2 stand for the velocity and acceleration vectors in the inertial frame
(frame 2), respectively. C12 is the rotation matrix from the frame 1 to the
frame 2 while C21 is the rotation matrix from the frame 2 to the frame 1.
288 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

ª ¹
« º
« º
«« ºº
¬ »
ª ¹
« º
« º
«« ºº
¬ »
0 1

ª
«¬ 0 1 0 1
¹
0 1 0 1 º
»
ª ¹
« º
« º
«« ºº
¬ »
ª ¹
« º
« º
«« ºº
¬ »

ª
«¬

00 1 0 11 00 1
0 11
00 1 0 11
¹
00 1 0 11 º»

ª ¹
«¬ 0 1 º»

Figure 6.2: Maple program for Example 6.5 (contd.).


6.7. MAPLE 289

ª
«
¬
§ ¶
¨
©
·
¸
0 1 0 1 0 1
§ ¶
¨ ·
© ¸
§ ¶
¨
©
·
¸
0 1 0 1 0 1
§ ¶ ¹
¨ · º
© ¸ »

ª
«
¬
§§ ¶
¨¨
©©
·
¸
0 1 0 1 0 1
§ ¶¶ §§ ¶
¨ ·· ¨¨ ·
© ¸¸ ©© ¸

0 1 0 1 0 1
§ ¶¶
¨ ··
© ¸¸
§§ ¶
¨¨
©©
·
¸
0 1 0 1
§ ¶¶
0 1 ¨
©
··
¸¸
§§ ¶
¨¨ ·
©© ¸

0 1 0 1 0 1
§ ¶¶ ¹
¨ ·· º
© ¸¸ »

ª § ¶ ¹
«
¬
0 1 0 10 1 ¨
©
·
¸
º
»

Figure 6.3: Maple program for Example 6.5.

Example 6.6
Solve Example 5.1 using Maple.
Solution.
The state vectors of the system are expressed in terms of xi , i=1,2,3,4:
290 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

X = [R, Ṙ, θ, θ̇] = [x1 , x2 , x3 , x4 ] (6.84)


The Maple program is shown in Fig. 6.4. Eq1 and Eq2 represent the equa-
tions of motion for R and θ degrees of freedom, respectively.

0 1
0 1

0 1

0 1

0 1

0 1 0 1
0 1
0 1

0 1

0 1 0 1
0 1 0 1

Figure 6.4: Maple program for Example 6.6.


6.7. MAPLE 291

Example 6.7
Solve Example 5.2 using Maple.
Solution.
The state vectors of the system are expressed in terms of xi , i=1,2:

X = [β, β̇] = [x1 , x2 ]

The Maple program is shown in Fig. 6.5. W represents θ̇ and Eq1 represents
equation of motion for β degree of freedom.

0 1
0 1

0 0 11
0 0 11

0 0 11
0 1

0 1
0 1

0 1
0 1

0 1 0 1
0 1 0 1
0 1

0 0 1 0 1 1
00 1 0 1 0 11

Figure 6.5: Maple program for Example 6.7.


292 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

Example 6.8
Solve Example 5.2 and numerically simulate the system equation of mo-
tion using Maple.
Solution.
The state vectors of the system are expressed in terms of xi , i=1,2:

X = [β, β̇] = [x1 , x2 ]

The Maple program is shown in Figs. 6.6-6.7. W represents θ̇ and Eq1


represents equation of motion for β degree of freedom.
6.7. MAPLE 293

0 1
0 1
0 0 11
0 0 11

0 0 11
0 1
0 1
0 1
0 1
0 1
0 1 0 1
0 1 0 1
0 1
0 1 0 1
0 1 0 1
0 1
0 1
0 1 0 1
0 1
0 0 1 1
§ ¶ 0 1 0 1
¨ ·
© ¸
0 1
§ ¶
¨ ·
© ¸

04 5 1

Figure 6.6: Maple program for Example 6.8 (contd.).


294 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

1.0

x 0.5

0.0
0 2,500 5,000 7,500 10,000
t
-0.5

-1.0

Figure 6.7: Maple program for Example 6.8.

6.8 Summary
In this chapter, mathematical analysis of a nonlinear system including equi-
librium analysis, stability analysis and closed-form solutions is presented.
Numerical simulation of a given system and the use of Matlab and Maple are
6.8. SUMMARY 295

explained next. Some of the important results discussed in this chapter are
summarized as follows.

Nonlinear System Equation

Second Order Form ẍi = ψ(x1 , x2 , · · · , xn ), i = 1, 2, . . . , n


First Order Form ẋi = ψ(x1 , x2 , · · · , x2n ), i = 1, 2, . . . , 2n

Linear System Equation


Pn ∂ψ
Second Order Form δẍi = i=1 ∂xi δxi xi =xri , i = 1, 2, . . . , n
First Order Form ẋ = Ax
Characteristic Equation |A − sI| = 0, s = eigenvalue; I = identity matrix

References

1. E. Kreyszig, Advanced Engineering Mathematics, John Wiley & Sons,


8th edition, 1999.
2. H. Goldstein, Classical Mechanics, Addison-Wesley, Reading, MA, sec-
ond
3. E.T. Whittaker, A Treatise of Analytical Dynamics of Particles and
Rigid Bodies, Dover, 1944 (1904).
4. Leonard Meirovitch, Methods of Analytical Dynamics, McGraw-Hill,
New York, 1970.
5. Marshall H. Kaplan, Modern Spacecraft Dynamics & Control, John
Wiley & Sons, New York, 1976.
296 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

Problem Set 6

6.1 The motion of a dumbbell system undergoing 3-dimensional librational


motion is described by equations of motion

β 00 − 2(1 + β 0 )η 0 tanη + 3sinβcosβ = 0 (6.85)


00 0 2 2
η + (1 + β ) sinηcosη + 3sinηcosηcos β = 0 (6.86)

where ()0 denotes derivative with respect to reference angle θ.


Find the system equilibrium conditions and derive the system linear
equations of motion with respect to null states and corresponding closed-
form solution. Furthermore, determine stability conditions of the given
system.
6.2 The equations of motion of a system of an inverted pendulum fixed to
a moving cart are obtained as

(M + m)ẍ + mlcosθθ̈ = −cẋ + mlsinθθ̇2 + u


(J + ml2 )θ̈ + mlcosθẍ = −mglsinθ

where c is the damping coefficient.


Is the system linear or nonlinear? If the system is nonlinear, derive the
linear system by linearizing the equations of motion about the equilib-
rium (i.e., θ = π).
Using simulink/matlab, compare the results of the linear and nonlinear
systems assuming any initial conditions.
Hint. In the given problem, State: x, θ, ẋ, θ̇; Input: u; Output: y=x.
6.3 The differential equation of motion of a system is obtained as

α̈ + 5α̇ + 25α = 0

Find the following:


(a) ωn , natural frequency
(b) ζ, damping ratio
(c) ωd , damped natural frequency
6.4 Given the second-order differential equation

θ̈ + 2θ̇ + 5θ = −δ
6.8. SUMMARY 297

(a) Rewrite the equation in state space form:

ẋ = Ax + Bu

(b) Determine the eigenvalues of the A matrix.


6.5 For a given linear system

ẋ1 = 2x1 + x2
ẋ2 = −x2

find eigenvalues and eigenvectors by hand calculation.


6.6 For a given linear system

ẍ + 4ẋ + 2x = 0

find eigenvalues and eigenvectors by hand calculation.


6.7 Given the differential equations

ẋ1 + 0.5x1 − 10x2 = −1δ


ẋ2 − x2 + x1 = 2δ

where x1 and x2 are the state variables and δ is the forcing input to
the system:
(a) Rewrite these equations in state space form; that is

ẋ = Ax + Bu

(b) Find the free response eigenvalues (i.e., u=0).


(c) What can you tell about the response of the system based on the
eigenvalues.
6.8 Given the following differential equation

...
x + ẍ − 4ẋ + 6x = r

(a) Rewrite the equation in state space form:


298 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

ẋ = Ax + Bu

(b) If the characteristic equation is given by

(λ + 3)(λ2 − 2λ + 2) = 0

describe the free response modes of the motion.


6.9 For a given linear equations of motion of a system

ẋ = Ax + Bu (6.87)

where x=state vector; A and B are constant coefficient matrix; u=control


input vector.
Prove that the time for doubling or halving of the amplitude and the
corresponding number of cycles are

0.693
t2 or t1/2 =
|η|
|ω|
N (cycles)2 or N (cycles)1/2 = 0.110
|η|

6.10 For a given characteristic equation of a system

λ4 + 5.05λ3 + 13.2λ2 + 0.67λ + 0.59 = 0

state about the stability of the system, i.e., whether the system will be
dynamically stable or unstable.
For the given system, find T (period), t1/2 (time to half amplitude),
and N1/2 (number of cycles to half amplitude).
6.11 For a second order system with mass m, damping coefficient c, and
stiffness k, write the characteristic equation of the system. Define the
natural frequency ωn and the damping ratio ζ of the system. Draw the
location of the roots on the complex s-plane as ζ is varied from −∞
to ∞ keeping ωn constant. When do you call the system as negatively
damped, undamped, underdamped, critically damped and overdamped.
Draw the sketches of the system response with input as unit impulse
for the cases of ζ < −1, −1 < ζ < 0, ζ = 0, 0 < ζ < 1, ζ = 1, ζ > 1.
6.8. SUMMARY 299

6.12 Consider the system described in Problem 6.11, draw the sketches of the
system response with input as unit step for the cases of ζ < −1, −1 <
ζ < 0, ζ = 0, 0 < ζ < 1, ζ = 1, ζ > 1. For an underdamped system 0 <
ζ < 1, what is the time constant of the system? Derive the relationships
for rise time (tr ), delay time (td ), settling time (ts ), time at which the
maximum overshoot occurs (tp ), and the corresponding overshoot (Mp ).
What is the definition of tr for the overdamped system? What are the
approximate relationships for tr , td , and ts applicable for designing the
controller later in Chapter 6?
6.13 The characteristic equation is of order n. Prove that the coefficient of
λn−1 is the negative of the sum of the real parts of all the roots, and
hence is aptly termed ”the sum of the dampings.” For example, you
can take n=3 as well as n=2.
6.14 The equations of motion of a system are given by

α̈ + a11 β̈1 + a12 α = 0


α̈ + a21 β̈1 + a22 β̇1 = 0

Find the order of the characteristic equation. In case one of the eigen-
values is zero, then explain why?. Here aij , i,j=1,2 are constant coef-
ficients.
6.15 The equations of motion of a system are given by

φ̈ + b11 β̈3 + b12 β̇2 + b13 γ̇ + b14 φ = 0


γ̈ + b21 β̈2 + b22 β̇3 + b23 φ̇ + b24 γ = 0
φ̈ + b31 β̈3 + b32 β̇3 + b33 γ̇ = 0
γ̈ + b41 β̈2 + b42 β̇2 + b43 φ̇ = 0

Find the order of the characteristic equation. In case some eigenval-


ues are zero, then explain why? Here bij , i,j=1,2,. . . ,4 are constant
coefficients.
6.16 Simulate the full nonlinear equations of motion of a spacecraft using
Simulink/matlab. Use the inertia properties of a typical spacecraft
model:
Ix =0.1147 kg-m2 , Iy =0.0576 kg-m2 , Iz =0.1712 kg-m2 , Ixy =0 kg-m2 ,
Ixz =0.0015 kg-m2 , Iyz =0 kg-m2
Compare this result with the linear equations of motion of the sys-
tem. Take initial attitude angles and rates as 10 degree and 3 deg/sec.
Convince yourself that the dynamic response is reasonable.
300 CHAPTER 6. MATHEMATICAL ANALYSIS AND SIMULATION

6.17 The equations of motion for a spacecraft in polar coordinates are


µ
r̈ − rθ̇2 = − + ar
r2
rθ̈ + 2ṙ θ̇ = aθ

where ar and aθ are the radial and transverse perturbing acceleration.


Convert these into a set of first-order equations suitable for numerical
integration.
6.18 Solve Example 5.3 using Maple.
6.19 Solve Example 5.4 using Maple.
Chapter 7

Control System

In previous chapters, we studied the dynamics of space systems. However,


even though a given space system is precisely positioned and oriented at
the beginning, it deviates in time from its preferred position and orienta-
tion under the influence of environmental forces or torques such as gravity
gradient, solar radiation pressure, magnetic forces, aerodynamic forces, and
free molecular reaction forces. Internal disturbances such as those within the
payload, e.g., motion of astronauts or sloshing of propellants and coupling
of the attitude dynamics with the orbital and flexural mechanics may also
lead to undesirable librational motion which must be controlled for successful
completion of a space mission.
This chapter introduces the concept of designing controllers for space
systems, especially those that account for gravity gradient forces. Linear
controllers are explained first, followed by design of nonlinear controllers. In
nonlinear controllers, the focus is on the design of intelligent controllers based
on fuzzy logic, neural networks, and neuro-fuzzy controllers. In addition, the
controller can be the combination of linear and nonlinear controllers.
The tuning of gains of the controllers is an important task for achieving
desired system performance. A control problem can be treated an optimiza-
tion problem. Linear optimal control theory based on the application of the
linear quadratic regulator (LQR) method is applied to develop optimal con-
trol laws. A Lyapunov approach can also be utilized to design control laws.
There are several near optimal methods including evolutionary optimizations
(such as genetic algorithms, simulated annealing, and genetic programming).
We will discuss these methods in this chapter.
Note that this chapter deals with the design of continuous controllers.
However, in practice for controlling on-orbit space systems, they have to be
discrete controllers. The principles of designing continuous controllers can be
302 CHAPTER 7. CONTROL SYSTEM

extended to design discrete controllers.

7.1 Introduction
As we have seen in the previous chapter on Dynamics, the equations of motion
of space systems are mostly second order systems expressed as

ẍ = ψ(x) + Td (7.1)

where ψ(x) can be nonlinear functions of system state x, and Td is an ex-


ternal disturbance force or torque vector. The preceding equations can be
alternatively written in first order form as

ẋ = φ(x) + Td (7.2)

Here x = [x1 , x2 , · · · , xn ]T and n denotes number of system state variables.


As we found in the previous chapter, the system response does not follow a
desired system state. For example, the inplane attitude motion of a dumbbell
system (see Example 4.1) has an undamped librational motion as shown in
Fig. (7.1). However, for a particular mission it is required to have β within

15

10

5
β (deg)

−5

−10

−15
0 0.5 1 1.5 2 2.5 3 3.5 4
Orbits

Figure 7.1: Inplane attitude motion of a dumbbell system (R=6878 km,


β0 = 10o , β̇0 = 0.01θ̇).

±0.01o. To achieve such a performance, we must apply an external torque


termed control input using thrusters or other actuators. The system equations
of motion with control input can be written as

ẋ = φ(x) + Td + u (7.3)
7.2. LINEAR CONTROLLERS 303

The control input u is to be designed to follow a prescribed profile in order


to achieve desired system performance. It can be a function of system states
as

u = g(x) (7.4)

If g(x) is a linear function, given by

g(x) = k(x − xr ) (7.5)

where xr denotes a desired or reference state and k is a controller gain matrix.


This controller is called a linear controller.
If g(x) is a nonlinear function, the controller is known as a nonlinear
controller. If g(x) is based on the concepts of fuzzy logic, neural networks, and
neuro-fuzzy, the corresponding controllers are called fuzzy logic controllers,
neural networks controllers, and neuro-fuzzy controllers, respectively.

7.2 Linear Controllers


The controllers mostly used in industry are of a linear type. They have been
proven to be robust, simple, and stable for many practical applications. The
general structure of linear controllers is shown in Fig. 7.2. The most devel-
oped part of control theory deals with linear systems and powerful techniques
for designing controllers for such systems are available in literature. How-
ever, as applications become more complex, the processes to be controlled
are increasingly characterized by poor models, distributed sensors and actu-
ators, multiple subsystems, high noise levels, etc. The difficulty envisioned in
designing controllers for such processes can be broadly classified under three
groups: 1) complexity, 2) nonlinearity, and 3) uncertainty. In the presence of
uncertainties in the plant, actuator dynamics, sensor dynamics, environmen-
tal conditions, and so on, linear controllers have been found to be inadequate
to provide the desired performance. Robust linear controllers are designed
to control such systems. It is a well known paradigm in control literature
that in order to achieve robustness, some part of performance has to be com-
promised. This can be viewed as an inherent difficulty with robust linear
controller design. Here, by ‘controllability’ we mean the ability of a con-
troller to transfer the system from any initial state x(t0 ) to any other state
in a finite time interval, whereas, 0 robustness0 is the ability of a controller
to maintain the system’s desired performance even with changes in internal
structure or external environment.
In designing linear controllers, the system equations of motion are to be
linearized with respect to a reference system state and it is expressed in state
304 CHAPTER 7. CONTROL SYSTEM

Reference Linear Input Output


System
xr Controller x

Figure 7.2: Linear controller for the system.

space form as

ẋ = Ax + Bu + GTd (7.6)
y = Cx + Du + HTd (7.7)

where A, B, C, D, G, and H are constant matrixes. x and y are input and


output state vectors, respectively, while, u and Td denote control input and
disturbance force or torque vector, respectively.
For example, we consider a dumbbell system undergoing inplane attitude
motion (see Example 5.1) given by the system equation of motion
3
β̈ + Ω2 sin2β = 0 (7.8)
2

Here Ω = θ̇ denotes system angular velocity.


Let us consider a reference state x = xr ,i.e., β = βr , β̇ = β̇r , β̈ = β̈r .
Substituting the reference state in Eq. (7.8), the resulting equation is
3
β̈r + Ω2 sin2βr = 0 (7.9)
2

Next the reference state is perturbed by δx as

x = xr + δx (7.10)

Putting the perturbed state in Eq.(7.8), we have the following equation


3
β̈r + δ β̈ + Ω2 sin(2βr + 2δβ) = 0 (7.11)
2

Subtracting Eq.(7.11) from Eq.(7.9), we get


3
δ β̈ + Ω2 [sin(2βr + 2δβ) − sin2βr ] = 0 (7.12)
2
7.2. LINEAR CONTROLLERS 305

or
3
δ β̈ + Ω2 {sin(2βr )[cos(2δβ) − 1] + cos(2βr )sin(2δβ)} = 0 (7.13)
2
or

δ β̈ + 3Ω2 cos(2βr + δβ)sinδβ = 0 (7.14)

We consider a first order perturbation, i.e., ignore second and higher order
terms of the perturbed state δβ. Using the Taylor series expansion for sin(x)
and cos(x)

1 3 1
sin(x) = x − x + x5 + . . . (7.15)
3! 5!
1 2 1
cos(x) = 1 − x + x4 + . . . (7.16)
2! 4!

in Eq. (7.14) and applying the first order perturbation, the resulting
equation is

δ β̈ + 3Ω2 cos(2βr )δβ = 0 (7.17)

The preceding Eq. (7.17) is a linearized equation of motion of the given


system described by the nonlinear Eq. (7.8).
The linearized equation of motion can be alternatively derived by simply
applying differentiation to the nonlinear Eq. (7.8) of the system as

δ β̈ + Ω2 Aδβ = 0 (7.18)
h i
where A = ∂ 3 sin2β = 3cos(2βr )
∂β 2 βr
The procedure of differentiation is rather easy and it already assumes a
first order perturbation of the perturbed state.
The reference state in general is taken as an equilibrium state xe defined
as
ẋ = φ(x) = 0 (7.19)

Here x = [β, β̇]T .


Substituting the preceding conditions (7.19) in the equations of motion

β̈ + 3Ω2 sinβcosβ = 0 (7.20)


306 CHAPTER 7. CONTROL SYSTEM

we get

3Ω2 sinβe cosβe = 0

or

sin2βe = 0 (7.21)
as Ω 6= 0.
Thus,

βe = 0, π/2, π, . . . , nπ/2 (7.22)

Here n is an integer.
Let us design a proportional-derivative (PD) controller (denoted by u) for
this linear system
β̈ + 3Ω2 cos(2βr )β = u (7.23)
Note that we write δ β̈ and δβ as β̈ and β just for convenience. We consider
the desired or the reference state as βr = βe = 0, and β̇r = β̇e = 0. The PD
control input can be expressed as
u = −kp (β − βr ) − kd (β̇ − β̇r ) (7.24)
Substituting the control input along with the reference state in the preceding
equation as βr = βe = 0, and β̇r = β̇e = 0, we have
β̈ + kd β̇ + (3Ω2 + kp )β = 0 (7.25)

The above equation is compared with the standard second-order system:


ẍ + 2ζωn ẋ + ωn2 x = 0 (7.26)
where ωn is the undamped natural frequency and ζ is the damping ratio.
Considering desired ωn of the system with the control input be equal to
ωn of the free librational motion of the system
ωn2 = 3Ω2 (7.27)
and the damping ratio ζ = 0.7, we have
3Ω2 + kp = ωn2 = 3Ω2 (7.28)
kd = 2ζωn (7.29)
7.2. LINEAR CONTROLLERS 307

and thus the control gains are obtained as


kp = 0, kd = 2 × 0.7 × 3Ω = 2.425Ω (7.30)

Taking the system orbiting at R=6878 km, µ=3.986× 105 km3 /s2 , we have
Ω=0.00111 rad/sec and the corresponding derivative gain is

kd = 0.00269 (7.31)

Applying this PD controller to the nonlinear system described by Eq. (7.8),


we obtain the system response as shown in Fig. 7.3.

12

10

6
β (deg)

−2
0 0.5 1 1.5 2 2.5 3 3.5 4
Orbits

Figure 7.3: Inplane attitude motion of a dumbbell system with control input
(R=6878 km, β0 = 10o , β̇0 = 0.01Ω, kp =0, kd =0.00269).

Example 7.1
Design a PD controller for a rigid spacecraft undergoing three-dimensional
attitude motion. The desired performance is specified by closed-loop frequen-
cies and damping ratios of ωq , q = α, φ, γ and ζq , q = α, φ, γ, respectively.
Solution.
Referring to Chapters 5 and 6, the equations of motion of a rigid spacecraft
are given by
308 CHAPTER 7. CONTROL SYSTEM

Satellite: Pitch(α)
α00 cosφcosγ − φ00 sinγ − (1 − kxz + kyz )(1 + α0 )φ0 sinφcosγ
− (1 + kxz − kyz )[(1 + α0 )γ 0 cosφsinγ + φ0 γ 0 cosγ]
+ (kxz − kyz )(1 + α0 )2 sinφcosφsinγ
+ 3(kxz − kyz )(cosαsinφsinγ − sinαcosγ)cosαcosφ = 0 (7.32)

Satellite: Roll(φ)
kyz α00 cosφsinγ + kyz φ00 cosγ − (1 − kxz + kyz )(1 + α0 )φ0 sinφsinγ
− (1 − kxz − kyz )[(1 + α0 )γ 0 cosφcosγ − φ0 γ 0 sinγ]
+ (1 − kxz )(1 + α0 )2 sinφcosφcosγ
+ 3(1 − kxz )(cosαsinφcosγ + sinαsinγ)cosαcosφ = 0
(7.33)

Satellite: Yaw(γ)
−kxz α00 sinφ + kxz γ 00 + [(1 − kyz )cos2γ − kxz ](1 + α0 )φ0 cosφ
+ (1 − kyz )[(1 + α0 )2 cos2 φ − φ02 ]sinγcosγ
− 3(1 − kyz )(cosαsinφcosγ + sinαsinγ)
× (cosαsinφsinγ − sinαcosγ) = 0 (7.34)

and the corresponding linearized equations of motion of the spacecraft


with respect to the null state vector are
k2 − k1
δα00 − 3 δα =0 (7.35)
1 − k1 k2
δφ00 + (1 − k1 )δγ 0 + 4k1 δφ =0 (7.36)
00 0
δγ + (k2 − 1)δφ + k2 δγ =0 (7.37)

To design a linear controller (i.e., PD controller), we consider the linear


equations of motion with control input uq , q = α, φ, γ, as
k2 − k1
δα00 − 3 δα =uα (7.38)
1 − k1 k2
δφ00 + (1 − k1 )δγ 0 + 4k1 δφ =uφ (7.39)
00 0
δγ + (k2 − 1)δφ + k2 δγ =uγ (7.40)

and comparing these linear equations with the desired closed-loop re-
sponse
δα00 + 2ζα ωα δα0 + ωα2 δα =0 (7.41)
δφ + 2ζφ ωφ δφ + ωφ2 δφ
00 0
=0 (7.42)
δγ 00 + 2ζγ ωγ δγ 0 + ωγ2 δγ =0 (7.43)
7.3. NONLINEAR CONTROLLERS 309

results in a PD controller given as


k2 − k1
uα = − 2ζα ωα δα0 − ωα2 δα − 3 δα (7.44)
1 − k1 k2
uφ = − 2ζφ ωφ δφ0 − ωφ2 δφ + (1 − k1 )δγ 0 + 4k1 δφ (7.45)
0
uγ = − 2ζγ ωγ δγ − ωγ2 δγ 0
+ (k2 − 1)δφ + k2 δγ (7.46)

7.3 Nonlinear Controllers


In this section, we will discuss Fuzzy logic controllers and neural networks
controllers.

7.3.1 Fuzzy Logic Controllers


Fuzzy logic was first proposed by Lotfi A. Zadeh of the University of California
at Berkeley in 1965. There have been numerous applications of the fuzzy-
logic controllers including washing machines, refrigerators, subway trains,
elevators, and space systems. The general structure of the fuzzy logic con-
troller is shown in Fig. 7.4.

Reference Fuzzy Input Output


System
xr Controller x

Figure 7.4: Fuzzy Logic controller for the system.

In designing fuzzy controllers, we have the following steps:


1) Fuzzification
The system input and output states are considered as linguistic variables
and represented in linguistic terms such as “Low”, “Medium”, and “High”.
The membership functions for linguistic terms of the input and output states
are then assumed. These membership functions can have various shapes
including triangular, rectangular, trapezoidal, sinusoidal, bell-shaped, and
monotonic. In general, numerical values of input and output states should
be between -1 and 1 or 0 and 1, and to achieve such a range, they are scaled
by appropriate input and output scaling factors, respectively.
310 CHAPTER 7. CONTROL SYSTEM

2) Rules
The rules between the input and output states in linguistic terms are
prescribed.
3) Defuzzification
The output state in linguistic terms or fuzzy sets is finally defuzzified
and expressed in numerical values using Mean of Maximum (MOM) method
or center of area (COA) method. The MOM method is also called Sugeno
method.

7.3.2 Neural Networks Controllers

Recently, neural networks (NN) have emerged as a means of explicitly ac-


counting for uncertainties in plant dynamics. Their on-line learning and
functional approximation capabilities make NN an excellent candidate for
this type of application. In recent years NN have been proposed for iden-
tification and control of linear and nonlinear dynamic systems [1-6]. The
recent development in NN theory provides us with new tools for control of
complex systems. The NN with their self-organizing and learning ability,
are now being utilized as promising tools for such purposes. The inclusion
of semilinear sigmoidal functions in the hidden neurons allows the network
to learn and solve control problems. The NN can be trained to learn any
function provided that enough information is given during the training pro-
cess coupled with judiciously selected neural models. Thus, this self-learning
ability of the NN eliminates the use of complex and difficult mathematical
analyses which is dominant in many traditional adaptive and optimal control
methods. Neural networks control or neuro-control is a control scheme using
NN.
The NN controller is required to be trained in order to have a desired
performance. The manner in which the NN is trained is an important prob-
lem for the practical realization of controllers for complex dynamical systems.
The backpropagation algorithm which is based on the steepest descent (gra-
dient) method, is the most popular algorithm for training neural networks’
parameters, and generally converges to the local optimum. To obtain the
global optimum, the genetic algorithm (GA)[7-9] is one of the most powerful
techniques. The GA is extremely useful for control problems in which the sys-
tem’s dynamics or the performance measures are discontinuous, non-smooth,
or non-convex, as it is not gradient-based and can operate regardless of the
complexity of the problem dynamics or performance specifications. The GA
involves a directed, discontinuous search over a class of candidate solutions
with the intent of finding the optimum performance measure. With the GA
technique, a near-optimal solution can be obtained. However, this approach
involves high computational efforts especially as it approaches the optimal
7.3. NONLINEAR CONTROLLERS 311

solution and therefore, some modifications have been suggested to solve this
problem.
A neural network controller can be of a PID type or based on past inputs
and outputs. The NN controller of PID type (Fig. 7.5) has control input
u(t) as
Z t
u(t) = N [xr , x, (x − xr )dτ, ẋ] (7.47)
0

whereas the NN controller based on past inputs and outputs (Fig. 7.6) has
the control input u(t) as
u(k) = N [x(k), · · · , x(k − n + 1), xr , u(k − 1), · · · , u(k − n + 1)] (7.48)

Reference Neural Networks Input Output


System
xr Controller x

Figure 7.5: Neural networks controller for the system.

Reference Neural Networks Input Output


System
xc Controller x

Figure 7.6: Neural networks controller based on past inputs and outputs of
the system.

Neural networks can be of three types: multiple neural networks (MNN)


controller, fully forward neural networks controller (FNN) and sparse fully
forward neural networks controller (SFNN). An n-layer multilayer neural net-
works (MNN) with p as input vector and u as output can be written as,
u = Γ [Wn Γ [Wn−1 · · · [W1 p + b1 ] + · · · + bn−1 ] + bn ] (7.49)
312 CHAPTER 7. CONTROL SYSTEM

where Wi is the weight matrix associated with the ith layer, the vectors bi
(i = 0, 1, 2, , n) represent the threshold values of each neuron in the ith layer
and Γ[.] is an operator with Γ[x]=[γ(x1 ), γ(x2 ), · · · , γ(xn )]T where γ(x) is the
activation function that can be a linear or nonlinear function. The structure
of MNN is shown in Fig. 7.7. The structure of neural networks with n-number
of hidden layers can be described by R − S 1 − S 2 − S 3 − · · · − S j − · · · Sn T .
Here, R is number of inputs. T is the number of outputs. Sj is the number
of neurons in jth hidden layer.
The equations for fully forward connected neural networks (FNN) are
given by

x1 = p (7.50)
 
Xi−1
xi = γi  wij xj  , yi = xi , 2 ≤ i ≤ n (7.51)
j=1

u = xn (7.52)

where γi is the activation function at the ith layer and the input neuron is a
linear neuron.
Input Layer Hidden Layer Output Layer

W1 W1
R x1
S1x R + f
1
S2 x S1 + f2
1 1
1 b 1 b
1 2
R 1
S x1 S 2
S x1 S

Figure 7.7: Multi-layered neural networks structure [f 1 (n) = (1 − e−n )/(1 +


e−n ); f 2 (n) = n].
7.4. SUMMARY 313

7.4 Summary
In this chapter, we have discussed linear controllers and their applications
on space systems are illustrated through example problems. Next, nonlinear
controllers based on fuzzy logic, neural networks, and genetic algorithms are
explained. The summary of important results is presented in a table form as
follows.

Nonlinear System Equation

Second Order Form ẍi = ψ(x1 , x2 , · · · , xn ), i = 1, 2, . . . , n


First Order Form ẋi = ψ(x1 , x2 , · · · , x2n ), i = 1, 2, . . . , 2n

Linear System Equation


Pn ∂ψ
Second Order Form δẍi = i=1 ∂xi δxi xi =xri , i = 1, 2, . . . , n
First Order Form ẋ = Ax
Characteristic Equation |A − sI| = 0, s = eigenvalue; I = identity matrix

Linear PD Controller

Linear System δẍi = f (δx1 , δx2 , · · · , δxn ) + ui , i = 1, 2, . . . , n


Desired Performance δẍi + 2ζi ωi δ ẋi + ωi2 δxi = 0, i = 1, 2, . . . , n

References

1. Arthur E. Bryson, Jr., Control of Spacecraft and Aircraft, Princeton


University Press, Princeton, 1994.

2. Marshall H. Kaplan, Modern Spacecraft Dynamics & Control, John


Wiley & Sons, New York, 1976.

3. J. R. Wertz, editor, Spacecraft Attitude Determination and Control, D.


Reidel, Dordrecht, Holland, 1978.
314 CHAPTER 7. CONTROL SYSTEM

4. Marcel J. Sidi, Spacecraft Dynamics and Control: A Practical Engi-


neering Ap- proach, Cambridge University Press, Cambridge, 1997.

5. K. S. Narendra and K. Partasarathy, Identification and Control of Dy-


namical Systems using Neural Networks, IEEE Transactions on Neural
Networks, 1(1), 1990, 4-27.

6. K. J. Hunt, D. Sbarbaro, R. Zbikowski, and P. J. Gawthrop, Neural


Networks for Control Systems - A Survey, Automatica, 28(6), 1992,
1083-1097.

7. K. S. Narendra, Neural Networks for Control: Theory and Practice,


Proc. of IEEE, 84(10), 1996, 1385-1406.

8. K. S. Narendra, Neural Networks for Intelligent Control, American


Control Conference, Workshop No. 4, 1997.

9. M. T. Hang, and H. B. Demuth, Neural Networks for Control, Proc.


American Control Conf., 1999, 1642-1656.

10. S. Ferrari, and R. F. Stengel, Classical/Neural Synthesis of Nonlinear


Control Systems, Journal of Guidance, Control and Dynamics, 25(3),
2002, 442-446.

11. S. Mondolni, A Genetic Algorithm for Determining Optimal Flight Tra-


jectories, Proc. AIAA Guidance, Navigation, and Control Conf., 1998,
1646-1656.

12. L. S. Crawford, V. H. L. Cheng, and P. K. Menon, Synthesis of Flight


Vehicle Guidance and Control Laws using Genetic Search Methods,
AIAA-99-4153, Proc. AIAA Guidance, Navigation, and Control Conf.,
1999, 1123-1133.

13. D. E. Goldberg, Genetic Algorithm in Search, Optimization, and Ma-


chine Learning (Addison-Wesley Publishing Company, Inc, 1989).

14. L. R. Ray, and R. F. Stengel, Application of Stochastic Robustness to


Aircraft Control Systems, Journal of Guidance, Control and Dynamics,
14(6), 1991, 1251-1259.

15. C. I. Marrision, and R. F. Stengel, Synthesis of Robust Control Systems


for a Hypersonic Aircraft, Proc. 33rd Conf. Decision and Control, 1994,
3324-3329.

16. C. I. Marrision, and R. F. Stengel, Design of Robust Control Systems


for Hypersonic Aircraft, Journal of Guidance, Control and Dynamics,
21(1), 1998, 58-63.
7.4. SUMMARY 315

17. W. M. Schubert, and R. F. Stengel, Parallel Synthesis of Robust Con-


trol Systems, IEEE Transactions on Control Systems Technology, 6(6),
1998, 701-706.
18. Y. Miyazawa, and T. Motoda, Stochastic Gain Tuning Method Ap-
plied to Unmanned Space Vehicle Flight Control Design, AIAA-99-
4309, Proc. AIAA Guidance Navigation and Control Conf. ,1999.
19. T. Motoda, R. F. Stengel, and Y. Miyazawa, Robust Control System
Design Using Simulated Annealing, Journal of Guidance, Control and
Dynamics, 25(2), 2002, 267-273.
20. D. L. Carroll, Fortran GA - Genetic Algorithm Driver V1.7, Users
Guide, 1998.
21. Maxwell Noton, Spacecraft Navigation and Guidance (Advances in In-
dustrial Control)
22. Gene F. Franklin, J. David Powell, Abbas Emami-Naeini, Feedback
Control of Dynamic System, Fourth Edition
23. Gene F. Franklin, David J. Powell, Michael L. Workman, Dave Powell,
Digital Control of Dynamic Systems.
316 CHAPTER 7. CONTROL SYSTEM

Problem Set 7

7.1 NASA has proposed a mission to send a probe to the Mars to investigate
life on it. The mission involves the landing of the probe on the surface
of the Mars. However, during descending of the probe towards the sur-
face of the Mars, its attitude may get disturbed as the system neither
possesses rotational stiffness nor damping and thereby, the whole mis-
sion may be a failure. For the success of the mission, the attitude of
the probe must be controlled precisely. The gas jets are provided for
this purpose and they apply moments about the probe’s center of mass.
Take the moment of inertia of probe as I=100 kg-m2 .

(a) Draw the block diagram of the system. Consider an open-loop con-
trol, write the system characteristic equation and find the roots of the
characteristic equations.

(b) Determine the system attitude response in the presence of initial atti-
tude angle θ0 and rate θ̇0 errors. Is the system stable?

(c) What is the type of input? (Is it step or ramp or parabolic?)

(d) In order to satisfy the performance requirements, it was desired to apply


feedback control. Answer the following questions:

(d.1) Apply a proportional controller, find the roots of the characteristic


equation? Is the system stable? Is the system marginally stable? (i.e.,
roots on the imaginary axis.) If yes, what should the range of propor-
tional controller gain Kp ? What would be the system response if the
initial attitude error is θ0 ?

(d.2) Can you apply final value of theorem of Laplace Transform to find the
steady state error? Explain the reason.

(d.3) If you consider a derivative control action only, what would be the roots
of the characteristic equation? Is the system stable? If yes, determine
the range of the derivative controller gain Kd ? Can you apply final
value of theorem of Laplace Transform to find the steady state error?

(d.4) Is it possible to apply a proportional-derivative (PD) control to convert


the marginally stable system considered above to a stable system (i.e.,
roots having negative real parts.)? Determine the steady state error.
Is it possible to meet the desired performance requirements with a PD
controller? If yes, what would be the range of the proportional gain Kp
and derivative gain Kd ? (Hint. by looking at the location of roots on
the s-plane or applying the Routh-Hurwitz criterion)
7.4. SUMMARY 317

7.2 A ground-based system comprises of a radar tracking dish, a sensor,


and a motor actuator. The system is required to track an airplane
moving across the sky at the fixed angular rate of 3o /s. The following
data about the system and the performance requirements are available:

Table 1
System Data Moment of inertia Damping Coefficient
20 kg-m2 5 N-m-sec/rad
Performance Steady-State Tracking Error System Time Constant
≤ 0.1o ≤ 0.5 sec

(a) Draw the block diagram of the system.


(b) Design a PD controller to meet the desired performance.
(c) If Kp =10 and Kd =1, can the desired performances be met?
7.3 A system is modelled as a mass m restrained by a linear spring of
stiffness k and a linear damper having damping constant c. The sys-
tem responds very sluggishly to a step input. In order to produce
quicker response, a PD controller is added, draw the block diagram of
the controlled system. The following data about the system and the
performance requirements are available:

Table 2
System Data Mass (m) Stiffness (k) and
Damping Coefficient (c)
1 kg k=1 N/m,c=4 N-sec/m
Performance Settling time (ts ) Peak Overshoot (Mp )
≤1 s ≤ 20%

7.4 A spacecraft is desired to maintain a specified pitch orientation in the


presence of aerodynamic disturbance torques Td . Reaction wheels are
rotated to produce desired roll moments. The system has pitch moment
of inertia I=100 kg-m2 and damping c=10 N-m-sec/rad. The pitch
stiffness is not present. If Td of unit impulse producing an initial angular
318 CHAPTER 7. CONTROL SYSTEM

velocity θ̇0 is acted on the system, determine the time histories θ(t) and
θ̇(t) taking θ0 =0. Determine the final value of θ(t) as t tends to ∞.
Apply a proportional controller, find the eigenvalues of the controlled
and uncontrolled system. What would be stability of the system if
proportional controller gain Kp =50?

7.5 The open-loop transfer function of the pitch motion of a spacecraft


with control torque u as an input and pitch angle of the spacecraft θ
as output is
θ(s) 1.151s + 0.1774
= 3
u(s) s + 0.739s2 + 0.921s

The step input of u=2 rad is applied. Answers the following questions:

(a) Checking the location of the roots of the characteristic equations


of the system, state whether the open loop response is stable?
What would happen if the input is null?
(b) For the closed-loop control, write the transfer function applying a
proportional controller? Is the system response stable? What are
the range of the proportional controller gain Kp . Is the system
response stable for Kp =2? (Hint. by looking at the location of
roots on the s-plane or applying the Routh-Hurwitz criterion)
(c) What would happen to the stability of the system if you apply
proportional-plus-derivative (PD) control with Kp =9 and Kd =8?
(d) Could you expect improvement in the system response if proportional-
integral-derivative (PID) control with Kp =2, Ki =4, and Kd =3 is
considered?
(e) What is the effect on the stability of the system if proportional-
plus-integral (PI) control is applied?

7.6 An open-loop transfer function of the system is

X(s) 1
= 2
F (s) s + 10s + 20

(a) For an unit step input, design a suitable controller ( i.e, P or


PD or PI or PID) to meet the performance requirements of rise
time (tr )≤ 0.8 sec, peak overshoot (Mp )=0, and steady state error
ess =0.
(b) Checking the location of the roots of the characteristic equations
of the system, state whether the open-loop response is stable. Find
ess .
7.4. SUMMARY 319

(c) For the closed-loop control, write the transfer function applying
proportional controller? Is the system response stable? What
are the range of the proportion controller gain Kp . Is the system
response stable for Kp =300? (Hint. by looking at the location of
roots on the s-plane or applying Routh-Hurwitz criterion)
(d) What would happen to the stability of the system if you ap-
ply proportional-plus-derivative (PD) control with Kp =300 and
Kd =10?
(e) Could you expect improvement in the system response if proportional-
integral-derivative (PID) control with Kp =350, Ki =300, and Kd =50
is considered?
(f ) What is the effect on stability of the system if proportional-plus-
integral (PI) control with Kp =30, Ki =70 is applied?
7.7 Consider a transfer function
K(s + z1 )(s + z2 ) . . . (s + zm )
st (s + p1 )(s + p2 ) . . . (s + pn )
with gain K, m number of zeroes, and n number of poles. Write the
steady state errors for unity feedback system of type 0 (t=0), type 1
(t=1), and type 2 in the presence of step input/ramp input/parabolic
input.
7.8 What are lead and lag controllers? Add a lag controller with a transfer
function of
s+1
s + 0.1
to the forward loop of DC motor speed control with the open-loop
transfer function of
θ K
=
V (Js + c)(Ls + R) + K 2
where J=0.01 kg-m2 , c=0.1 N-s/m, K=0.01, R=1 Ω, L=0.5 Henry
Take step input of 1 rad/s. How is the closed-loop response with regard
to the steady state error (ess ) and the settling time (ts ) with only pro-
portional controller? Design the suitable proportional controller gain
Kp to satisfy the following desired requirements: ts < 2 sec, Overshoot
(Mp ) < 5%, and ess < 1%. How is the response with Kp =10? How-
ever, you will observe that ts and Mp requirements can be met but the
requirement on ess can not be satisfied with the proportional controller.
To meet ess requirement, adding a lag controller can be effective. What
would be the effect on adding a lag controller on ess and ts with the
proportional gain of 50? Does it reduce ess by a factor of 100 and
increase the settling time?
Note. Adding a lag controller in general reduce ess .
320 CHAPTER 7. CONTROL SYSTEM

7.9 The open-loop transfer function of a spacecraft is expressed as a third-


order system:

1.5 × 107 K
G(s) =
s(s2 + 3408.3s + 1, 204, 000)

However, to understand the system dynamics and design a suitable


control gain K, the system is approximated as a second-order system:

4500K
G(s) =
s(s + 361.2)

Answer the following questions:


(a) For the second-order system, find the natural frequency and damp-
ing ratio of the closed-loop system.
(b) What is the range of K for which the second-order approximation
is valid in relation with the closed-loop response of the system?
(c) For K=14.5, the roots of the characteristic equation of the third-
order closed-loop system are

s1 = 186.53 + j192, s2 = 186.53 − j192, s3 = −3035.2

which root will dominate the transient response and what will be the
corresponding equivalent damping ratio.
7.10 A space system is in a 6878 km circular orbit about a spherical Earth
(Fig. 7.8). It comprises a spacecraft m1 and two bodies m2 and m3
connected through rigid massless cables of length L1 and L2 . The
masses of the two bodies are very small in comparison to the mass of
the spacecraft and therefore, the system center of mass is considered to
coincide with the center of mass of the spacecraft. Assuming m1 , m2 ,
and m3 as point masses and L1 = 10m and L2 = 5m, design two PD
controllers (one for β1 motion and another for β2 motion) to stabilize
the system along local vertical (β1 = β2 = 0) within ±0.01o in half an
orbit and examine its performance in the following situations:
(a) linear equations of motion, 5o initial attitude error in β1 , no dis-
turbance
(b) linear equations of motion, 5o initial attitude error in β2 , no dis-
turbance
(c) nonlinear equations of motion, 5o initial attitude errors in β1 and
β2
7.4. SUMMARY 321

m3
β2
Orbit L2
Local Vertical
m2
L1 β1
Local Vertical
S m1
Y R

θ
E
X

Figure 7.8: System undergoing in-plane libration.

(d) nonlinear equations of motion, 135o initial attitude errors in β1


and β2 , 1 Nm disturbance torques along β1 and β2

7.11 A spacecraft is in a 6878 km circular orbit about the spherical Earth,


and its principal moments of inertia are Ix =40 kg-m2 , Iy =30 kg-m2 ,
Iz =50 kg-m2 . Design a PID controller with ωn = ωp , ζ = 0.7, and
T = 10/(ζωn ). The frequency ωp is defined to be the natural frequency
of the open-loop pitch motion of a spacecraft in the same orbit, but with
Ix =40 kg-m2 , Iy =50 kg-m2 , Iz =30 kg-m2 , which in fact characterizes
the stable system motion.
Examine the controller performance in the following situations and dis-
cuss the results:
(a) linear pitch equation, 10o initial attitude error, no disturbance
(b) linear pitch equation, 135o initial attitude error, 1 Nm disturbance
torque
(c) nonlinear pitch equation, 10o initial attitude error, no disturbance
(d) nonlinear pitch equation, 135o initial attitude error, 1 Nm distur-
bance torque
Hint. The system equation with the PID controller is
Z t
ẍ = −kp x − kI x(τ )dτ − kd ẋ (7.53)
0
322 CHAPTER 7. CONTROL SYSTEM

The closed-loop characteristic equation is

s3 + kd s2 + kp s + kI = 0 (7.54)

and we want this polynomial to factor as

(s2 + 2ζωn s + ωn2 )(s + 1/T ) = 0 (7.55)

where ωn is the natural frequency, ζ is the damping ration, and T is the


integral time constant. Relating the gains to these parameters leads to

kp = ωn2 + 2ζωn /T (7.56)


kI = ωn2 /T (7.57)
kd = 2ζωn + 1/T (7.58)

Take these gains for designing a PID controller.


7.12 The control laws for a rigid spacecraft undergoing three-dimensional
attitude motion are considered to be

uα = − kpα δα − kdα δα0 (7.59)


0
uφ = − kpφ δφ − kdφ δφ (7.60)
uγ = − kpγ δγ − kdγ δγ 0 (7.61)

where α, φ, and γ are the pitch, roll, and yaw attitude of the space-
craft. kpq and kdq , q = α, φ, γ are proportional and derivative gains,
respectively.
Derive the closed-loop characteristic equation of the system given by
k2 − k1
δα00 − 3 δα = 0 (7.62)
1 − k1 k2
δφ00 + (1 − k1 )δγ 0 + 4k1 δφ = 0 (7.63)
δγ 00 + (k2 − 1)δφ0 + k2 δγ = 0 (7.64)

and state about the controller gains for stable response.


7.13 Design a fuzzy controller for an inplane attitude motion of a dumbbell
system described in Example 7.1.
7.14 Design a fuzzy controller for the problem in Example 7.1.
7.15 Design a neural networks controller for an inplane attitude motion of a
dumbbell system described in Example 7.1.
7.16 Design a neural networks controller for Example 7.1.
Chapter 8

Formation Flying

This chapter presents formation flying of satellites beginning with formation


classification. This is followed by the Leader-Follower type of formation. The
chapter concludes with a case of a leader satellite in an elliptic orbit.

8.1 Introduction
Satellite formation flying is defined as two or more satellites flying in a co-
operative manner in prescribed orbits at a fixed separation distance for a
given period of time. Very close coordination among satellites comprising
the formation is essential. In addition, autonomous operation will be nec-
essary. The simplest example could be two satellites in a circular orbit of
same orbital radius, i.e., same period. Their separation between them can
be specified by phase difference or mean anomaly difference. Note that the
relative distance between them remains constant if there exists only central
gravitational force acting (no J2 perturbation; and other forces). In other
words, we can say that the satellites are in formation. In another situation, if
the satellites are in elliptic orbits of same semi-major axes, i.e., same period,
their separation is specified by phase difference of mean anomaly difference.
Even in the presence of only central gravitational force (no J2 or other per-
turbations), the relative distance between them will change. The separation
will increase at apogee and decrease at perigee. However, for a particular
application, say taking images of the Earth’s magnetic field, equal distance
between neighboring satellites is required when they are at perigee. There
is indeed an interesting solution to this problem. If the axis of the nodes of
each satellite orbit should be separated by 1.34 deg, it is possible to main-
tain the constant distance at perigee. However, if there exist environmental
disturbances, these solution can only be achieved using thrusters.
324 CHAPTER 8. FORMATION FLYING

Formation flying is significantly different from Constellation wherein two


or more satellites are designed to fulfill a common mission objective that
can not otherwise be effectively or efficiently accomplished by just one space-
craft operating alone. These satellites are relatively independent and tightly
coupled. Typical examples are LEO global communication satellite constel-
lations, and the Global Positioning System (GPS).
The formation flying of satellites has been identified as an enabling tech-
nology for many future space missions [1]-[8] including Air Force TechSat
21[3] and NASA ST5 Nanosat Constellation Path Finder [4], Ionospheric Ob-
servation Nanosatellite Formation (ION-F)[4] and NASA’s Terrestrial Planet
Finder (TPF) mission[6]. Compared to a single spacecraft mission, this ap-
proach has several advantages including the ability to enhance and/or enable
missions through longer baseline observations and high failure tolerance. It
is real-time reconfigurable, adaptable to highly dynamic demands, and has
lower life cycle cost[7]. However, its development involves tremendous chal-
lenges ranging from spacecraft formation initialization to reconfiguration, co-
ordination, and formation trajectory generation. Satellite mass, power, fuel,
and communications are significant constraints and the guidance, naviga-
tion, and control (GNC) tasks become highly complicated for larger forma-
tions. The satellite formation may experience environmental disturbances
from gravitational perturbation, atmospheric drag, solar radiation pressure
and electromagnetic forces. With a view to tackle these challenges, innova-
tive methods for achieving satellite formation flying with minimum station
keeping requirements are sought.
The formation flying dynamics has been studied using orbital elements[8]-
[12] as well as relative Cartesian coordinates[13]-[16]. Several investigations[13]-
[16] considered linearized relative motion equations in a Cartesian coordinate
frame, called the Hill’s Equations[13] or the Clohessy-Wiltshire equations[14].
Satellite formation initialization is an important step in achieving a desired
formation. Several researchers including Sabol et al.[16], Inalhan et al.[17],
and Vaddi et al.[18] have discussed this problem with reference to relative
Cartesian coordinates. Inalhan et al.[17] described an initialization proce-
dure for formation with eccentric reference orbits. The initial conditions
accommodating corrections due to nonlinearity as well as eccentricity have
been presented by Vaddi et al.[18]
The problem of satellite formation keeping and reconfiguration has been
examined based on impulsive control[8]-[11],[19],[20] as well as continuous
thrust control [15],[21]-[25]. Schaub and Alfriend[9] proposed impulsive feed-
back control laws for establishment of a desired set of mean-element differ-
ences. A two-impulse solution is suggested by Vaddi et al.[10] for achieving
the desired formation characterized by orbital-element differences. Vassar
and Sherwood[19] presented impulsive control for leader and follower satellites
in a circular orbit using a Cartesian coordinate model. Wiesel[20] described
8.2. FORMATION CLASSIFICATION 325

an optimal impulse control of relative satellite motion and solved the re-
sulting optimization problem numerically. Sparks[11] analyzed discrete-time,
linear feedback control for satellite formationkeeping. Irvin and Jacques[21]
compared linear and nonlinear feedback control laws, as well as continuous
and discrete burn techniques for the satellite formation reconfiguration. Yeh
et al.[22] derived a tracking control design using sliding mode techniques
to control a desired satellite formation. Mitchell and Richardson[23] ap-
plied a method based on invariant manifold tracking for controlling first-
order nonlinear Hill’s equations. The application of fuzzy for formation
control is considered by Qingsong et al.[15]. de Quierpoz[24] presented a
Lyapunov-based, nonlinear, adaptive controller for multiple spacecraft for-
mation flying. The satellite formation using no radial thrust has also been
examined[12],[25],[26]. This investigation might have been prompted because
of fuel efficiency, propulsion system simplifications, and weight reduction in
achieving desired formation.
We begin with formation classification.

8.2 Formation Classification


Formation flying can be classified in a number of ways. Based on command
or control architecture, we have
Centralized Formation. The formation is called a centralized forma-
tion wherein there is a leader satellite and all other satellites follow the leader
satellite. It is also known as Leader-Follower formation.
Decentralized Formation. Decentralized formation is those formation
where each satellite acts individually and the formation is achieved by con-
trolling an individual satellite.
Based on the arrangement of satellites in a formation, we have the fol-
lowing classifications:
Inplane Formation. The in-plane formation is the simplest of all for-
mation configurations. The motion of all the satellites in this formation are
confined to the same orbital plane and the in-plane spacing among them are
defined by mean anomaly difference. The main advantage of the in-plane
formation is its simplicity in design, deployment, and control.
Circular Formation and Projected Circular Formation. In the
circular formation of a Leader-Follower type formation, the follower satellites
remain at a constant distance from the leader satellite and their paths make
a sphere. However, in the projected circulation formation, the projection of
the formation in a plane is a circle. These formations are the generalization
of a general elliptical formation and they are the most useful for imaging
applications. Using these types of formations, a plane of interest can be filled
326 CHAPTER 8. FORMATION FLYING

with an arbitrary number of satellites whose relative geometry is fixed and


whose apparent relative motion is a periodic rotation about the formation
center.
Triangular Formation and Tetrahedron Formation. In Triangu-
lar formation, satellites in formation are located at the vertices of a triangle
while in Tetrahedron formation, they are situated at the vertices of a tetra-
hedron. Similarly, we can define rectangle, pentagon, hexagon, and other
similar formations.
We have another classification of formation as well depending upon whether
the formation as a whole is spinning or non-spinning about its formation cen-
ter. If it is spinning, it is called a Rotating Formation.

8.3 Leader Satellite in Circular Orbit


Assuming the leader and follower satellites are of point masses under the
Earth’s central gravitational force of attraction, we can write their equations
of orbital motion referring to Fig. 8.1 as

G(Me + mc )
~r¨c = − r~c (8.1)
rc3
G(Me + m)
~r¨ = − ~r (8.2)
r3
where ~rc and ~r are the position vectors of the leader and follower satellite
with respect to the Earth center. The nomenclature Me , mc , and m denote
the masses of the Earth, leader satellite, and follower satellite, respectively.
Considering mc , m  Me and knowing that ρ ~ = ~r − ~rc , the relative equation
of motion of the follower satellite with respect to the leader or chief satellite
using Eqs. (8.1-8.2) can be written as
  µ µ
~¨ = ~r¨ − ~r¨c = − 3 (~rc + ρ
ρ ~) + 3 ~rc (8.3)
r rc

where µ = GMe . The above relative equation of motion in the vector form is
to be expressed in a scaler form. To do so we fix a cartesian coordinate frame
S − xyz on the mass center of the leader satellite (Fig. 8.1). We first find
the left-hand side of the preceding equation that defines the kinematics of
the relative motion. Note that ρ~¨ is the inertial acceleration as per Newton’s
second law of motion. However, ρ ~ is defined with respect to the non-inertial
frame S − xyz and so we write the inertial acceleration ρ~¨ as

~¨ = ρ
ρ ~¨xyz + 2(~ ~˙ xyz ) + ω
ω×ρ ~ × (~
ω×ρ ~˙ × ρ
~xyz ) + ω ~xyz

where ω
~ is an angular velocity vector of the S − xyz frame.
8.3. LEADER SATELLITE IN CIRCULAR ORBIT 327

Follower Satellite
y
ρ
S x
Inertial Y
r Leader Satellite
Reference
Frame
rc z
θ
Reference Line
Earth O
X

Z
Figure 8.1: Formation of Leader and Follower Satellite.

Defining ρ
~ = xî + y ĵ + z k̂, we can further rewrite

ρ~¨ =ẍî + 2ẋ(~


ω × î) + x[ω~˙ × î + ω
~ × (~ω × î)]
+ ÿ ĵ + 2ẏ(~
ω × ĵ) + y[ω~˙ × ĵ + ω
~ × (~
ω × ĵ)]
+ z̈ k̂ + 2ż(~ ~˙ × î + ω
ω × k̂) + z[ω ~ × (~
ω × k̂)] (8.4)

Assuming the leader satellite is in a circular orbit and its motion is in the
orbit plane, we have

ω
~ = θ̇k̂ (8.5)
~˙ = θ̈k̂ = 0
ω (8.6)

where θ defines
p the position of the leader satellite with respect to the reference
line and θ̇ = µ/rc3 is its orbital angular velocity. Using the above relations
in Eq. (8.4) lead to

ρ~¨ = [ẍ − 2θ̇ẏ − xθ̇2 ]î + [ÿ + 2ẋθ̇ − y θ̇2 ]ĵ + z̈ k̂ (8.7)

The position of the follower satellite, ~r, can be written in the S − xyz frame
as

~r = ~rc + ρ
~ = (rc + x)î + y ĵ + z k̂ (8.8)

and the corresponding magnitude |~r| is given by

r = [(rc + x)2 + y 2 + z 2 ]1/2 (8.9)


328 CHAPTER 8. FORMATION FLYING

Using Eqs. (8.7) and (8.8) in Eq. (8.3), the relative equations of motion
of the follower satellite with respect to the leader satellite can be obtained as

µ µ
ẍ − xθ̇2 − 2ẏθ̇ = − (rc + x) + 2 (8.10)
r3 rc
µ
ÿ − y θ̇2 + 2ẋθ̇ = − 3 y (8.11)
r
µ
z̈ = − 3 z (8.12)
r

p
where r = [(rc +x)2 +y 2 +z 2 ]1/2 and θ̇ = µ/rc3 , a constant. The coordinates
x, y, and z are also called radial, along-track, and cross-track components
of motion. Note that the preceding equations are nonlinear equations as
the right-hand side of the equations have nonlinear gravitational force terms
and so we can not find a closed-form solution of these equations. We can
obtain the solution only using the numerical simulation applying six initial
conditions (i.e., at t=0, x = x0 , ẋ = ẋ0 , y = y0 , ẏ = ẏ0 , z = z0 , ż = ż0 ).
However, with suitable approximations, we can derive the linear equations of
relation motion and for which we have closed-form solution.
In order to linearize the relative equations of motion Eqs. (8.10-8.12), we
substitute r from Eq.(8.9) into Eqs. (8.10-8.12) and write as

−3/2
x2 y2 z2

µ(rc + x) x
ẍ − xθ̇2 − 2ẏθ̇ = − 1 + 2 + + +
rc3 rc rc2 rc2 rc2
µ
+ 2 (8.13)
rc
−3/2
x2 y2 z2

µy x
ÿ − y θ̇2 + 2ẋθ̇ = − 3 1 + 2 + 2 + 2 + 2 (8.14)
rc rc rc rc rc
−3/2
x2 y2 z2

µz x
z̈ = − 3 1 + 2 + 2 + 2 + 2 (8.15)
rc rc rc rc rc

Assuming x  rc , y  rc , z  rc , we apply Binomial series expansion

n n(n − 1) 2 n(n − 1)(n − 2) 3


(1 + X) = 1 + nX + X + X + · · · if |X| < 1
1×2 1×2×3

of the terms inside the brackets in the right-hand side of the preceding equa-
8.3. LEADER SATELLITE IN CIRCULAR ORBIT 329

tions and neglect the second and higher order terms of x, y, and z:
 
µ(rc + x) x µ
ẍ − xθ̇2 − 2ẏθ̇ = − 1 − 3 + 2
rc3 rc rc
µ µ 2µ
= − 3 [rc − 2x] + 2 = 3 x (8.16)
rc r rc
  c
µ x µ
ÿ − y θ̇2 + 2ẋθ̇ = − 3 1 − 3 y = − 3y (8.17)
r rc rc
 
µ x µ
z̈ = − 3 1 − 3 z = − 3z (8.18)
rc rc rc
p
Knowing θ̇ = µ/rc3 leads to

ẍ − 2ẏθ̇ − 3θ̇2 x = 0 (8.19)


ÿ + 2ẋθ̇ = 0 (8.20)
2
z̈ + θ̇ z = 0 (8.21)
The preceding equations of motion are the linear equations of relative motion
and they are also called the Hill’s Equations or the Clohessy-Wiltshire (CW)
equations[13, 14].
To find the solution of the linear Eqs. (8.19-8.21), we change the variable
of differentiation t to θ, an angle measured from the reference line using the
following transformation:
dq dq dθ
q̇ = = × = q 0 θ̇ (8.22)
dt dθ dt
d2 q d(q 0 θ̇) dq 0 dθ̇ dq 0 dθ
q̈ = 2 = = θ̇ + q 0 = × θ̇ + q 0 θ̈ = q 00 θ̇2 (8.23)
dt dt dt dt dθ dt
where q 0 = dq/dθ. Note that θ̈=0. Applying these relations in Eqs. (8.19-
8.21) lead to the following relative equations of motion:
x00 − 2y 0 − 3x = 0 (8.24)
y 00 + 2x0 = 0 (8.25)
z 00 + z = 0 (8.26)
where the derivative terms are with respect to θ.
The last Eq. (8.26) can be solved and its solution with the initial condi-
tions of z = z0 and z = z00 at θ=0 is given by
z = z0 cosθ + z00 sinθ (8.27)
Next integrating Eq. (8.25) leads to
y 0 + 2x = C = y00 + 2x0 (8.28)
330 CHAPTER 8. FORMATION FLYING

where C, a constant and y 0 = y00 and x = x0 are initial conditions at θ=0.


Substituting y 0 from the preceding equation into Eq. (8.24) yields

x00 − 2(C − 2x) − 3x = 0

or

x00 + x = 2C (8.29)

The solution of the preceding equation can be written as

x = x00 sinθ + (x0 − 2C)cosθ + 2C (8.30)

where x = x0 and x0 = x00 are initial conditions are θ=0. Putting the above
expression of x into Eq.(8.28), we have

y 0 = −2x00 sinθ − 2(x0 − 2C)cosθ − 3C (8.31)

Integrating with initial conditions y0 and y00 at θ=0 leads to

y = 2x00 cosθ − 2(x0 − 2C)sinθ − 3Cθ + y0 − 2x00 (8.32)

We can summarize the solution of the linear equations of the relation motion
of the follower satellite with respect to the leader satellite as

x =x00 sinθ − (3x0 + 2y00 )cosθ + 2(2x0 + y00 ) (8.33)


y =(6x0 + 4y00 )sinθ + 2x00 cosθ − 3(2x0 + y00 )θ + y0 − 2x00 (8.34)
z =z0 cosθ + z00 sinθ (8.35)

Writing in a state-space form:

X(θ) = φ(θ)X(0) (8.36)

where X(θ), the system state vector, X(0), the initial system state vector,
and φ(θ), the state transition matrix are
   


 x  



 x0 


 
 
 

0 0
x x

 
 
 









 0 


   
 y 
   y 
 
0
X(θ) = , X(0) = , (8.37)


 y0 



 y00 



 
 
 

 z   z0 

  
 
 
  


   
 0  
 0 
z z0 
 
  
8.3. LEADER SATELLITE IN CIRCULAR ORBIT 331

and
 
−3cosθ + 4 sinθ 0 2(1 − cosθ) 0 0
 
 3sinθ cosθ 0 2sinθ 0 0 
 
 
 6sinθ − 6θ 2(−1 + cosθ) 1 4sinθ − 3θ 0 0 
 
φ(θ) =  
 6cosθ − 6 −2sinθ 0 4cosθ − 3 0 0 
 
 
0 0 0 0 cosθ sinθ
 

 
0 0 0 0 −sinθ cosθ
(8.38)
Note that in Eqs. (8.33) and (8.34), the terms 2xo + y00 and y0 − 2x00 lead to
the secular growth in the relative motion and therefore, for any formation to
be possible it is important that these terms must be removed, i.e.,
2xo + y00 = 0
y0 − 2x00 = 0
or
y0 =2x00 (8.39)
y00 = − 2x0 (8.40)
Thus, the preceding relations set the initiation conditions for achieving for-
mation. Applying these conditions in Eqs. (8.33-8.35), we get the solutions
of the relative motion as
x =x00 sinθ + x0 cosθ (8.41)
y = − 2x0 sinθ + 2x00 cosθ (8.42)
z =z0 cosθ + z00 sinθ (8.43)
This solution can also be expressed in the form:
q
x = x20 + x02
0 sin(θ + ϕx ) (8.44)
q
y = − 2 x20 + x02 0 cos(θ + ϕx ) (8.45)
q
z = z02 + z002 sin(θ + ϕz ) (8.46)

where the phases ϕx = tan−1 (x0 /x00 ) and ϕz = tan−1 (z0 /z00 ).
Using Eqs. (8.41) and (8.42) or Eqs. (8.44) and (8.45), we get the radial,
x, and along-track, y motion following the elliptical path
x2 y2
+ =1 (8.47)
C2 4C 2
332 CHAPTER 8. FORMATION FLYING
p
with 2
p semi-minor axis a = C = x0 + x0 and semi-major axis b = 2C =
02
2
2 x0 + x0 of an ellipse defined by
02

x2 y2
+ =1 (8.48)
a2 b2
In fact Eq. (8.47) defines an elliptical formation of radial and along-track
motion as
y2
x2 + = rf2
4
where rf =C, a constant, is the formation size.
Projected Circular Formation.
A projected circular formation is mathematically defined as
y 2 + z 2 = rf2 (8.49)
where rf , a constant, is formation size. Differentiating the preceding equation
with respect to θ yields
yy 0 + zz 0 = 0 (8.50)
Substituting the initial state vector, X0 as defined in Eq. (8.37), into the
preceding equation and applying the conditions (8.39) and (8.40), we have
y0 y00 + z0 z00 = 0 ⇒ −4x0 x00 + z0 z00 = 0 (8.51)
or
dz02 4dx20
= (8.52)
dθ dθ
Integrating leads to
z02 = 4x20 ⇒ z0 = ±2x0 (8.53)
Using this relation into Eq. (8.51) yields
−4x0 x00 + ±2x0 z00 = 0 ⇒ x0 [−4x00 + ±2z00 ] = 0 (8.54)
Assuming x0 6= 0, then
z00 = ±2x00 (8.55)

Putting X0 into Eq. (8.50) and applying the conditions (8.39) and (8.40)
along with the preceding condition (8.53), we get
y02 + z02 = rf2 ⇒ 4x02 2 2
0 + z0 = rf (8.56)
2 2 2
rf2
⇒ 4x02 02
0 + 4x0 = rf ⇒ x0 + x0 = (8.57)
4
(8.58)
8.3. LEADER SATELLITE IN CIRCULAR ORBIT 333

Thus, for achieving the projected circular formation we have the following
initial conditions of the state vector:

y0 =2x00 (8.59)
y00 = − 2x0 (8.60)
z0 = ± 2x0 (8.61)
z00 = ± 2x00 (8.62)
rf2
x20 + x02
0 = (8.63)
4
Substituting the preceding conditions in Eqs. (8.41)- (8.43), we obtain the
solution for the projected circular formation as

x =x00 sinθ + x0 cosθ (8.64)


y = − 2x0 sinθ + 2x00 cosθ (8.65)
z = ± (2x0 cosθ + 2x00 sinθ) (8.66)

Alternatively, we can also write the solution given by Eqs. (8.44)- (8.46) by
putting the conditions (8.59)-(8.63). We get the phases as

φx = tan−1 (x0 /x00 ) (8.67)


−1
φz = tan (z0 /z00 ) = tan −1
(x0 /x00 ) = φx (8.68)

and the solution is


rf
q
x = x20 + x020 sin(θ + ϕx ) = sin(θ + ϕx ) (8.69)
q 2
y =2 x20 + x020 cos(θ + ϕx ) = rf cos(θ + ϕx ) (8.70)
q q
z = z02 + z002 sin(θ + ϕz ) = 2 x20 + x020 sin(θ + ϕx ) = rf sin(θ + ϕx )
(8.71)
or
   


 x 

 sin(θ + ϕ)
  r  
f 
y =  2cos(θ + ϕ) (8.72)

2 


 
 
 z 
  2sin(θ + ϕ)

where the phase ϕ = ϕx = tan−1 (x0 /x00 ) = tan−1 (z0 /z00 ).


Circular Formation.
The circular formation is mathematically defined as

x2 + y 2 + z 2 = rf2 (8.73)
334 CHAPTER 8. FORMATION FLYING

where rf , a constant is formation size. Differentiating the preceding equation


with respect to θ yields

xx0 + yy 0 + zz 0 = 0 (8.74)

Substituting the initial state vector, X0 as defined in Eq. (8.37), into the
preceding equation and applying the conditions (8.39) and (8.40), we have

x0 x00 + y0 y00 + z0 z00 = 0 ⇒ x0 x00 − 4x0 x00 + z0 z00 = 0 (8.75)


⇒ −3x0 x00 + z0 z00 = 0 (8.76)

or
dz02 3dx20
= (8.77)
dθ dθ
Integrating leads to

z02 = 3x20 ⇒ z0 = ± 3x0 (8.78)

Using this relation into Eq. (8.76) yields


√ √
−3x0 x00 + ± 3x0 z00 = 0 ⇒ x0 [−3x00 + ± 3z00 ] = 0 (8.79)

Assuming x0 6= 0, then

z00 = ± 3x00 (8.80)

Putting X0 into Eq. (8.73) and applying the conditions (8.39) and (8.40)
along with the preceding condition (8.78), we get

x20 + y02 + z02 = rf2 ⇒ x20 + 4x02 2 2


0 + z0 = rf (8.81)
rf2
⇒ x20 + 4x02 2 2 2 02
0 + 3x0 = rf ⇒ x0 + x0 = (8.82)
4
(8.83)

Thus, for achieving the circular formation we have the following initial con-
ditions of the state vector:

y0 =2x00 (8.84)
y00 = − 2x0 (8.85)

z0 = ± 3x0 (8.86)

z00 = ± 3x00 (8.87)
rf2
x20 + x02
0 = (8.88)
4
8.3. LEADER SATELLITE IN CIRCULAR ORBIT 335

Substituting the preceding conditions in Eqs. (8.41)- (8.43), we obtain the


solution for the circular formation as

x =x00 sinθ + x0 cosθ (8.89)


y = − 2x0 sinθ + 2x00 cosθ (8.90)
√ √
z = ± 3x0 cosθ + 3x00 sinθ (8.91)

We can also write the solution given by Eqs. (8.44)- (8.46) by putting the
conditions (8.59)-(8.63). We get the phases

φx = tan−1 (x0 /x00 ) (8.92)


−1
φz = tan (z0 /z00 ) = tan −1
(x0 /x00 ) = φx (8.93)

and the solution is


rf
q
x = x20 + x020 sin(θ + ϕx ) = sin(θ + ϕx ) (8.94)
q 2
y =2 x20 + x020 cos(θ + ϕx ) = rf cos(θ + ϕx ) (8.95)

3rf
q q
z = z02 + z002 sin(θ + ϕz ) = 3x20 + 3x020 sin(θ + ϕx ) = sin(θ + ϕx )
2
(8.96)
or
   


 x 

 sin(θ + ϕ)
  r  
f 
y =  2cos(θ + ϕ) (8.97)

2  √


 
 
 z 
 
3sin(θ + ϕ)

where the phase ϕ = ϕx = tan−1 (x0 /x00 ) = tan−1 (z0 /z00 ).


Note that the solution obtained above is an exact solution for the linear
system described by Eqs. (8.24)-(8.26) or Hill’s Eqs. (8.19)-(8.21). However,
this solution is an approximate solution for the nonlinear system described
by Eqs. (8.10)-(8.12) and is valid only in the following situations:

(i) The formation size is very small in comparison to the orbital radius of
the leader satellite, i.e., x/rc  1, y/rc  1, and z/rc  1.

(ii) Time t or θ of the numerical simulation of the nonlinear system is very


short.

Note that the preceding linear system can also be called a first order approx-
imation of the nonlinear system as in deriving the linear system we consider
336 CHAPTER 8. FORMATION FLYING

the first order terms of the state variables in the expansion of the right-hand
side of Eqs. (8.13)-(8.15). However, we can have better approximation of
the nonlinear system if we consider second order approximation as explained
below.
Second Order Approximation.
Referring to Eqs. (8.13)-(8.15), we assume x  rc , y  rc , z  rc , and
keep first and second order terms, and neglect higher order terms of x, y, and
z. The resulting second order equations are
 2
z2
 
2 3µ y 2
ẍ − 2ẏ θ̇ − 3θ̇ x = + −x (8.98)
r4 2 2
 c

ÿ + 2ẋθ̇ = xy (8.99)
r4
 c

z̈ + θ̇2 z = xz (8.100)
rc4

Changing the variable of differentiation t to θ as explained previously, we


write the preceding equations as

y2 z2
 
x00 − 2y 0 − 3x = + − x2 (8.101)
2 2
y 00 + 2x0 =xy (8.102)
z 00 + z =xz (8.103)

where  = 3/rc . The solution of the preceding equations can be assumed to


be of the form:

X = Xh + Xn (8.104)

where X is the state of the system, X = [xh , x0h , yh , yh0 , zh , zh0 ]T ; Xh is the
homogeneous solution of Eqs. (8.101-8.103) when =0; Xn is the particular
solution due to nonlinearity in the right-hand side of Eqs. (8.101-8.103).
Substituting X from Eq. (8.104) into Eqs. (8.101-8.103) leads to
"
(yh + yn )2 (zh + zn )2
x00h + x00n − 2yh0 − 2yn0 − 3xh − 3xn =  +
2 2
#
− (xh + xn )2 (8.105)

yh00 + yn00 + 2x0h + 2x0n = (xh + xn )(yh + yn ) (8.106)


zh00 + zn00 + zh + zn = (xh + xn )(zh + zn ) (8.107)
8.3. LEADER SATELLITE IN CIRCULAR ORBIT 337

Knowing that

x00h − 2yh0 − 3xh = 0


yh00 + 2x0h = 0
zh00 + zh = 0

and neglecting terms with 2 (as   1) and  6= 0, the resulting equations


become
yh2 z2
x00n − 2yn0 − 3xn = + h − x2h (8.108)
2 2
yn00 + 2x0n =xh yh (8.109)
zn00 + zn =xh zh (8.110)

The preceding equations can be written in the form:

Xn0 = AXn + Bu (8.111)

where
   


 xn 

 0 1 0 0 0

 
  
x0n 3 0 0 2 00

 
  

 


 
  
 y
 
 0 0

0 1 0 0

n
Xn = , A= ,
yn0 0 −2 0 0 0 0
   

 


 
  
zn 0 0 0 0 0 1

 
  

 


 
  
 z0
 
 0 0 0 0 −1 0
n
   
 1

 0 0 0 0 0 




 0 


 yh2 2
   


0

1 0 0 0 0 
  z h − x2












 2 + 2 h




   
 0

0 1 0 0 0 
 
 0


B= , u= (8.112)


 0 0 0 1 0 0 




 xh yh 



 
 
 

0 0 0 0 1 0 
   










 0 



   
 0

0 0 0 0 1 
 
 xh zh

In the preceding Eqs. (8.111), the matrix A and B are constant while vector u
changes as per the formation type (i.e., projected circular formation, circular
formation or others). The matrix A is the same as in the case of the linear
system (Eqs. (8.24)-(8.26)) denoted as

Xh0 = AXh
338 CHAPTER 8. FORMATION FLYING

Thus, Eqs. (8.111) are a set of linear constant coefficient non-homogeneous


ordinary differential equations and we can write the analytical solution as
Z θ
Xn (θ) = φ(θ)Xn (0) + φ(θ − θt )Buh (θt )dθt (8.113)
0

where φ(θ) is the state transition matrix and it is given by Eq. (8.38) of the
linear system since the matrix A is the same in both linear and nonlinear
systems. Using the preceding relation 8.113 along with Eq. (8.104), the
complete solution of the system can be written
Z θ
X(θ) = φ(θ)Xh (0) + φ(θ)Xn (0) +  φ(θ − θt )Buh (θt )dθt (8.114)
0

For a given formation of projected circular formation or circular formation


or others as discussed previously, the complete solution given by Eq. (8.114)
can be obtained following the steps outlined here: (i) Assume initial Xn (0),
(ii) Substitute φ(θ) from Eq. (8.38) into Eq. (8.114), (iii) Substitute uh from
Eq. (8.112) into Eq. (8.114) using the known xh , yh , and zh solution for
a given formation, (iv) Substitute initial Xh (0) for a given formation, (iv)
Integrate the right-hand side of the integral of Eq. (8.114).

8.4 Leader Satellite in Elliptic Orbit


For the case of a leader satellite in an elliptic orbit, we can derive the equa-
tions of relative motion, referring to Eqs. (8.3)-(8.6) and taking ω ~˙ = θ̈ k̂, as
follows.
−µ µ
ẍ − xθ̇2 − 2ẏθ̇ − y θ̈ = 3 (rc + x) + 2 (8.115)
r rc
−µ
ÿ − y θ̇2 + 2ẋθ̇ + xθ̈ = 3 y (8.116)
r
−µ
z̈ = 3 z (8.117)
r
where θ is the true anomaly and r = [(rc + x)2 + y 2 + z 2]1/2 . In the preceding
Eqs. (8.115)-(8.117), we change the differentiation with respect to time t to
true anomaly θ referring to Eqs. (8.22) and (8.23):
q̇ = q 0 θ̇ (8.118)
00 2 0
q̈ = q θ̇ + q θ̈ (8.119)

where θ̇ and θ̈ are derived as explained next. Note these parameters refer to
the orbit of the leader satellite. The θ̇ is given by

h µp
θ̇ = 2 = 2 (8.120)
rc rc
8.4. LEADER SATELLITE IN ELLIPTIC ORBIT 339

where h is the orbital angular momentum, p is the semi-latus rectum, and rc


is the orbital radius. Differentiating Eq. (8.120) with respect to time t, we
have
h
θ̈ = −2 ṙc (8.121)
rc3

Knowing the relation p/rc = 1 + ecosθ and differentiating it with respect to


time, we get r˙c as
p rc esinθ
ṙc = 2
(esinθ)θ̇ = θ̇ (8.122)
(1 + ecosθ) 1 + ecosθ

Substituting ṙc into Eq. (8.121) and using Eq. (8.120) for θ̇ yield

hθ̇esinθ µ
θ̈ = −2 = −2 3 esinθ (8.123)
rc2 (1 + ecosθ) rc

Upon substitution of these relations into Eqs. (8.118) and (8.119) lead to

µp
q̇ = 2 q 0 (8.124)
rc

µp µpesinθ µ
q̈ = q 00 + q0 4 = 3 [(1 + ecosθ)q 00 − 2q 0 esinθ] (8.125)
rc4 rc (1 + ecosθ) rc
Applying the above expressions into the relative equations of motion given
by Eqs. (8.115)-(8.117) result in the following relative equations of motion
with respect to the true anomaly θ:

rc3
(1 + ecosθ)x00 − 2(x0 − y)esinθ − (x + 2y 0 )(1 + ecosθ) = (rc + x) + rc
r3
(8.126)
−rc3
(1 + ecosθ)y 00 − 2(x + y 0 )esinθ + (2x0 − y)(1 + ecosθ) = y (8.127)
r3
3
−r
(1 + ecosθ)z 00 − 2z 0 esinθ = 3c z (8.128)
r
Assuming x  rc , y  rc , z  rc , we apply Binomial series expansion of
the terms in the right-hand side of the preceding equations and neglect the
second and higher order terms of x, y, and z. The resulting first order or
linear system is

(1 + ecosθ)x00 − 2(x0 − y)esinθ − (x + 2y 0 )(1 + ecosθ) − 2x = 0 (8.129)


00 0 0
(1 + ecosθ)y − 2(x + y )esinθ + (2x − y)(1 + ecosθ) + y = 0 (8.130)
00 0
(1 + ecosθ)z − 2z esinθ + z = 0 (8.131)
340 CHAPTER 8. FORMATION FLYING

Rewriting the above equations in the form:

1
[(1 + ecosθ)2 x0 ]0 − (ecosθ)x − 2[(1 + ecosθ)y]0 = 0 (8.132)
1 + ecosθ
1
[(1 + ecosθ)2 y 0 ]0 − (3 + ecosθ)y + 2[(1 + ecosθ)x]0 = 0 (8.133)
1 + ecosθ
1
[(1 + ecosθ)2 z 0 ]0 + z = 0 (8.134)
1 + ecosθ

Taking

xp = (1 + ecosθ)x, yp = (1 + ecosθ)y, zp = (1 + ecosθ)z (8.135)

lead to

x00p − 2yp0 = 0 (8.136)


3yp
yp00 − + 2x0p = 0 (8.137)
1 + ecosθ
zp00 + zp = 0 (8.138)

Solving the preceding equations and writing in terms of x, y, and z, we have


the solution as
 
 2
 d2 e
x(θ) = d1 e + 2d2 e H(θ) sinθ − + d3 cosθ (8.139)
(1 + ecosθ)2
   
d4 d3
y(θ) = d1 + + 2d2 eH(θ) + + d3 sinθ
1 + ecosθ 1 + ecosθ
+ cosθ d1 e + 2d2 e2 H(θ)
 
(8.140)
   
d5 d6
z(θ) = sinθ + cosθ (8.141)
1 + ecosθ 1 + ecosθ

where dj , j = 1, 2, · · · , 6 are integration constants, and they are calculated


from initial conditions. The term H(θ) is

θ
cosθ
Z
H(θ) = dθ
θ0 (1 + ecosθ)3
 
2 −5/2 3eE 2 e
= −(1 − e ) − (1 + e )sinE + sinEcosE + dH (8.142)
2 2

where E is an eccentric anomaly given by cosE = (e + cosθ)/(1 + ecosθ). The


term dH is an integration constant calculated from H(θ0 )=0. For a typical
case when θ0 =0, dH =0.
8.4. LEADER SATELLITE IN ELLIPTIC ORBIT 341

Differentiating Eqs. (8.139)-(8.141) yields


x0 (θ) = d1 e + 2d2 e2 H(θ) cosθ + 2d2 e2 H 0 (θ) sinθ
   

2d2 e2 sinθ
   
d2 e
+ )2 + d3 sinθ − cosθ (8.143)
(1 + ecosθ (1 + ecosθ)3
 
d4 esinθ
y 0 (θ) = + 2d 2 eH 0
(θ)
(1 + ecosθ)2
   
d3 d3 esinθ
+ + d3 cosθ + sinθ
1 + ecosθ (1 + ecosθ)2
− d1 e + 2d2 e2 H(θ) sinθ + 2d2 e2 H 0 (θ) cosθ
   
(8.144)
   
d5 d5 esinθ
z 0 (θ) = cosθ + sinθ
1 + ecosθ (1 + ecosθ)2
   
d6 d6 esinθ
− sinθ + cosθ (8.145)
1 + ecosθ (1 + ecosθ)2
where
cosθ
H 0 (θ) = (8.146)
(1 + ecosθ)3
The initial and final states at θ=0 and θ=2π, respectively can be written as
a function of the integration constants:
e
x0 (0) = ed1 , x(0) = − d2 − d3 (8.147)
(1 + e)2
2e (2 + e)
y 0 (0) = 2
d2 + d3 (8.148)
(1 + e) 1+e
1
y(0) = (1 + e)d1 + d4 (8.149)
1+e
1 (2 + e)(1 + e)2 (1 + e)3 0
d1 = x0 (0), d2 = 2
x(0) + y (0) (8.150)
e e e2
1+e 0 2(1 + e)
d3 = y (0) − x(0) (8.151)
e e
(1 + e)2 0
d4 = (1 + e)y(0) − x (0) (8.152)
e
6e3 π(1 + e)
x0 (2π) = ed1 − d2 (8.153)
(1 − e2 )5/2
e
x(2π) = − d2 − d3 (8.154)
(1 + e)2
2e (2 + e)
y 0 (2π) = d2 + d3 (8.155)
(1 + e)2 1+e
1 6e2 π(1 + e)
y(2π) = (1 + e)d1 + d4 − d2 (8.156)
1+e (1 − e2 )5/2
342 CHAPTER 8. FORMATION FLYING

Referring to the preceding states, we can find a no-drift condition or peri-


odicity at θ0 = 0 (i.e., x(2π) = x(0), y(2π) = y(0), z(2π) = z(0)) when the
leader is at the perigee at the initial time as

y 0 (0) 2+e
=− (8.157)
x(0) 1+e

Writing in time domain:

ẏ(0) n(2 + e)
=− (8.158)
x(0) (1 + e)1/2 (1 − e)3/2

The preceding Eq. (8.157) or Eq. (8.158) states initial conditions for achiev-
ing bounded relative motion.

8.5 Summary
In this chapter, different models of satellite formation are presented with a
focus on a leader and follower type of formation. Initial conditions for a
stable formation are discussed.

References

1. Pete Aldridge, E. C. , Jr., Fiorina, C. S., Jackson, M. P., Leshin, L.


A., Lyles, L. L., Spudis, P. D., Tyson, N. d., Walker, R. S., and Zu-
ber, M. T., ”A Journey to Inspire, Innovate, and Discover,” Report of
the President’s Commission on Implementation of United States Space
Exploration Policy, June 2004.
2. NASA New Millennium Program, http://nmp.jpl.nasa.gov/.
3. Air Force Research Laboratory Space Vehicles Directorate, ”TechSat 21
fact sheet page,” http://www.vs.afrl.af.mil/factsheets/TechSat21.html.
4. Ticker, R.L, ”The New Millinnium Space Technolohy 5 (ST5) Project,”
Space 2000 conference, Albuquerque, NM, Feb. 28 - Mar. 2, 2000.
5. Campbell, M., Fullmer, R.R, and Hall, C.D., ”The ION-F Formation
Flying Experiments,” Advances in the Astronautical Sciences, Vol. 105,
2000, pp. 135-149.
6. Terrestrial Planet Finder Science Working Group, Terrrestrial Planet
Finder, edited by C.A. Beichman, N.J. Woolf, and C.A. Lindensmith,
NASA Jet Propulsion Lab, JPL Publ. 99-003, California Inst. of Tech-
nology, Pasadena, CA, 1999.
8.5. SUMMARY 343

7. Carpenter, J. R., Leitner, J. A., Folta, D. C., and Burns, R. D., ”Bench-
mark Problems for Spacecraft Formation Flying Missions,” AIAA Pa-
per 2003-5364, Aug. 2003.
8. Schaub, H., ”Relative Orbit Geometry Through Classical Orbit Ele-
ment Differences,” Journal of Guidance, Control, and Dynamics, Vol.
27, No. 5, 2004, pp. 839-848.
9. Schaub, H., and Alfriend, K. T., ”Impulsive Feedback Control to Estab-
lish Specific Mean Orbit Elements of Spacecraft Formations,” Journal
of Guidance, Control, and Dynamics, Vol. 24, No. 4, 2001, pp. 739-745.
10. Vaddi, S. S., Alfriend, K. T., Vadali, S. R., and Sengupta, P., ”For-
mation Establishment and Reconfiguration Using Impulsive Control,”
Journal of Guidance, Control, and Dynamics, Vol. 28, No. 2, 2005, pp.
262-268.
11. Sparks, A., ”Linear Control of Spacecraft Formation Flying,” AIAA
2000-4438, Aug. 2000.
12. Alfriend, K. T., Vaddi, S. S., and Lovell, T. A., ”Formation Mainte-
nance for Low Earth Near-Circular Orbits,” AAS/AIAA Astrodynam-
ics Specialist Conf., AAS Paper 03-652, Aug. 2003.
13. Hill, G. W., ”Researches in the Lunar Theory,” American Journal of
Mathematics, Vol. 1, 1878, pp. 5-26.
14. Clohessy, W., and Wiltshire, R., ”Terminal Guidance Systems for Satel-
lite Rendezvous,” Journal of the Astronautical Sciences, Vol. 27, No.
9, 1960, pp. 653-658.
15. Qingsong, M., Pengji, W., and Di, Y., ”Low-thrust Fuzzy Formation
Keeping for Multiple Spacecraft Flying, ” Acta Astronautica, Vol. 55,
No. 1, 2004, pp. 895-901.
16. Sabol, C., Burns, R., and McLaughlin, C. A., ”Satellite Formation
Flying Design and Evolution,” Journal of Spacecraft and Rockets, Vol.
38, No. 2, 2001, pp. 270-278.
17. Inalhan, G., Tillerson, M., and How, J. P., ”Relative Dynamics and
Control of Spacecraft Formations in Eccentric Orbits,” Journal of Guid-
ance, Control, and Dynamics, Vol. 25, No. 1, 2002, pp. 48-59.
18. Vaddi, S. S., Vadali, S. R., and Alfriend, K. T., ”Formation Flying:
Accommodating Nonlinearity and Eccentricity Perturbations,” Journal
of Guidance, Control, and Dynamics, Vol. 26, No. 2, 2003, pp. 214-223.
19. Vassar, R. H., and Sherwood, R. B., ”Formation Keeping for a Pair
of Satellites in a Circular Orbit,” Journal of Guidance, Control, and
Dynamics, Vol. 8, No. 2, 1985, pp. 235-242.
344 CHAPTER 8. FORMATION FLYING

20. Wiesel, W. E., ”Optimal Impulse Control of Relative Satellite Motion,”


Journal of Guidance, Control, and Dynamics, Vol. 26, No. 1, 2003, pp.
74-78.
21. Irvin, D. J., and Jacques, D. R., ”Linear vs. Nonlinear Control Tech-
niques for the Reconfiguration of Satellite Formations,” AIAA Paper
2001-4089, Aug. 2001.
22. Yeh, H., Nelson, E., and Sparks, A., ”Nonlinear Tracking Control for
Satellite Formations,” Journal of Guidance, Control, and Dynamics,
Vol. 25, No. 2, 2002, pp. 376-386.
23. Mitchell, J. W., and Richardson, D. L., ”Invariant Manifold Tracking
for First-Order Nonlinear Hill’s Equations,” Journal of Guidance, Con-
trol, and Dynamics, Vol. 26, No. 4, 2003, pp. 622-627.
24. de Queiroz, M. S., Kapila, V., and Yan, Q., ”Adaptive Nonlinear Con-
trol of Multiple Spacecraft Formation Flying,” Journal of Guidance,
Control, and Dynamics, Vol. 23, No. 3, 2000, pp. 385-390.
25. Starin, R. S., Yedavalli, R. K., and Sparks, A. G., ”Spacecraft For-
mation Flying Maneuvers Using Linear-Quadratic Regulation with No
Radiation Axis Inputs,” AIAA Paper 2001-4029, Aug. 2001.
26. Leonard, C. L., Hollister, W. M., Bergmann, E. V., ”Orbital Forma-
tionkeeping with Differential Drag,” Journal of Guidance, Control, and
Dynamics, Vol. 12, No. 1, 1989, pp. 108-113.
8.5. SUMMARY 345

Problem Set 8

8.1 For a given formation, the equations of motion are

ẍ − 2θ̇ẏ − 3θ̇2 x = 0
ÿ + 2θ̇ẋ = Fy

where the input force Fy is applied according to the following control


laws
(I)Fy = Kp (y − yc )
(II)Fy = Kp (y − yc ) + Kd (ẏ − ẏc )
(III)Fy = Kp1 (x − xc ) + Kd1 (ẋ − ẋc ) + Kp2 (y − yc ) + Kd2 (ẏ − ẏc )
Find the range of gains for the absolute stability of the system in each
case applying the Routh or Routh-Hurwitz criterion. Here θ̇ is a con-
stant and subscript ‘c’ denotes commanded value.
346 CHAPTER 8. FORMATION FLYING
Index

Potential energy, 65 polar, 42


rectangular, 40
Hubble space telescope, 1 spherical, 40
Coriolis acceleration, 44
Aerodynamic drag, 200
Apoapsis, 170 Definitive geomagnetic reference
Apsides, 170 field, 141
Argument of aries, 184 Descending node, 184
Argument of the periapsis, 184 Direction cosines, 71
Ascending node, 183
Assumed modes, 97 Earth’s oblateness, 200
Attitude determination and con- Eccentric anomaly, 179
trol subsystem, 11 Eclipse, 138
Attitude equations of motion Eigenvalue test, 269
rigid body, 254 Elastic deformation, 94
Attitude motion, 71 Elliptic orbit, 281, 338
Elliptical formation, 332
Backpropagation algorithm, 310 Energy, 64, 84
Bending stiffness, 102 Energy equation, 178
Bending stress, 98 Equinoctial elements, 198
Bode diagram, 266 Euler angles, 74
Boundary value problems, 283 Euler equations of motion, 225
Euler-Bernoulli beam, 99
Centered dipole model, 141 Explicit ODEs, 282
Centripetal acceleration, 44
Chandra X-ray observatory, 1 First point of Aries, 183
Circular formation, 333 Flexible body, 94
Classical orbital elements, 185 Flexural rigidity, 102
Command and data handling sub- Formation flying, 323
system, 11 Fully implicit ODEs, 282
Conic sections, 169
Coordinate frame Gauss’ variational equations, 199
cartesian, 40 Genetic algorithm, 310
Coordinates Geocentric reference frame, 45
cartesian, 40 Geometric boundary condition,
cylindrical, 40 97

347
348 INDEX

Guidance, navigation and con- Maple, 263, 286


trol system, 11 Mathematica, 263
Matlab, 263
Hamiltonian, 70 Matrix
Heliocentric orbit, 45 Hessian, 268
Hessian matrix, 268 Matrix minor test, 269
Hook’s law , 96 Mean anomaly, 179
Hooke’s law Minor axis, 172
shear, 98 Modulus of rigidity, 98
Molniya Orbits, 210
Inclination of the orbit plane, 184
Momentum, 60, 78
Inertia matrix, 80 angular, 60
Inertia tensor, 80
linear, 60
Inertial reference frame, 160
Initial value problems, 283 N-body problem, 168
International geomagnetic refer- Natural boundary condition, 97
ence field, 141 Neutral plane, 99
Neutral surface, 99
J2 perturbation, 204 Newton’s law of gravitation, 124
Jenkins-Traub algorithm, 266 Newton’s laws of motion, 160
Kepler’s first law, 173 Non-Keplerian orbit motion, 190
Kepler’s laws of orbit motion, 162 Numerical simulation, 282
Kepler’s third law, 182 Nyquist criterion, 266
Kepler’s time equation, 182 Octopole field model, 141
Keplerian orbit, 185 Orbit determination, 186
Kinetic energy, 65, 86 Orbital elements, 184
Orbital perturbations, 190
Lagrange reversion theorem, 182
Ordinary bending, 98
Lagrange’s equations of motion,
Osculating orbit, 195
226
Osculating orbital plane, 195
Lagrange’s planetary equations,
198 Periapsis, 170
Lagrangian function, 227 Planetary gravitational pertur-
Laguerre algorithm, 266 bations, 200, 211
Landsat, 211 Plastic deformation, 94
Latus rectum, 172 Poisson’s ratio, 97
Legendre polynomials, 124 Potential energy, 84, 124, 201
Line of apsides, 170 gravitational, 65
Line of nodes, 184 Power system, 11
Linearly implicit ODEs, 282 Projected circular formation, 332
Longitude of the ascending node, Pure bending, 98
184
Luni-Solar perturbation, 217 Quadrapole field model, 141

Major axis, 170 Reference frame


INDEX 349

inertial, 41
Restricted two-body problem, 165
Right ascension of the ascending
node, 184
Rigid body, 70
Rigid satellite, 275
Rotation matrix, 49
Routh-Hurwitz criterion, 266

Semilatus rectum, 172


Semimajor axis, 172
Semiminor axis, 172
Shear modulus of elasticity, 98
Solar radiation pressure, 133, 200
Spherical harmonic model, 141
Sputnik, 1
Strain
Longitudinal strain, 95
Normal strain, 95
Stress
Normal stress, 95
Structures and mechanisms, 11
Sugeno method, 310
Sunsynchronous, 6, 211
Sylvester’s criterion, 269

Telemetry tracking and control


subsystem, 11
Tesseral harmonics, 127
Theory of general relativity, 125
Thermal control subsystem, 11
Tilted dipole model, 141
Timoshenko beam Theory, 100
Triaxiality, 200
Two-body motion, 163
Two-body problem, 163

Vernal equinox, 183


Virtual work, 254
Vis-viva equation, 178

You might also like