You are on page 1of 77

Welcome To..

Satellite Tool Kit


Astronautics Primer
by

Jerry Jon Sellers


Based on

Understanding Space: An
Introduction to Astronautics
Copyright 1996 McGraw-Hill Inc.

by

Jerry Jon Sellers

Wiley J. Larson (editor)

Your Understanding of Space Starts Here!

COPYRIGHT NOTICE

McGraw-Hill owns the copyright to the book Understanding Space: An Introduction to


Astronautics. This primer was adapted from the book by Jerry Jon Sellers and Wiley J.
Larson. Analytical Graphics, Inc. retains copyright for this Primer. It is illegal to
reproduce this material without permission.

1
STK Astronautics Primer

OBJECTIVE

The objective of this primer is to provide you with a fundamental reference for key
concepts in astronautics, including:

♦ Space mission architecture


♦ History
♦ Dynamics
♦ Orbital mechanics
♦ Orbital elements
♦ Orbit propagation
♦ Ground tracks
♦ Satellite access

ABOUT THE AUTHOR

Jerry Jon Sellers was born in Harlan, Iowa. He has worked over 13 years at various
astronautics assignments including the NASA Johnson Space Center, where he
worked in Space Shuttle Mission Control (guidance and navigation) and the U.S. Air
Force Academy where he served on the faculty of the Department of Astronautics. He
was a distinguished graduate from the U.S. Air Force Academy in 1984 and has
earned a Master’s Degree in Physical Science from the University of Houston, Clear
Lake, a Master’s Degree in Aeronautics and Astronautics from Stanford University
and a Ph.D. in Satellite Engineering from the University of Surrey, UK. He currently
works as an international research and development liaison officer in London, UK,
and continues to write and consult on space mission analysis and design.

CONTENTS
Stepping Into Space
The big picture of why space is important and how the pieces fit together
Exploring Space
Some early “explorers” who’ve shaped our current understanding of orbits
An Introduction to Orbit Motion
Key concepts necessary for understanding orbit motion
Describing Orbits
Understanding orbital elements (two-line element sets), ground tracks and how they relate
to space missions

2
STK Astronautics Primer

Predicting Orbits
The “nuts and bolts” of predicting real-world satellite motion using orbit propagators
Satellite Access
The how, when and where of links between ground stations and satellites
Recommended Reading
Other great astronautics references

STEPPING INTO SPACE

Since the dawn of the Space Age only a few decades ago, we have come to rely more
and more on satellites for a variety of needs. Daily weather forecasts, instantaneous
world-wide communication, and a constant ability to keep an eye on not-so-friendly
neighbors are all examples of space technology that we’ve come to take for granted.

The purpose of this brief astronautics primer is to provide the reader with a
conceptual overview of important topics in orbital mechanics. Understanding these
key concepts will enhance your insight into the science behind Satellite Took Kit and
better equip you to apply these concepts to practical problems in space. We’ll begin
with a brief overview of space, space missions and space history. Then we’ll get into
the details of orbital mechanics to see how you can use STK to plot your path to the
stars.

Why is space so useful?


Getting into space is dangerous and expensive. So why bother? Space offers several
compelling advantages for modern society

♦ A global perspective—the ultimate high ground


♦ A universal perspective—from space we have a clear view of the heavens,
unobscured by the atmosphere
♦ A unique environment—free-fall and abundant resources make space the
true final frontier

Global Perspective

Space offers a global perspective. As you can see in Figure 1, the higher you are, the
more you can see. For thousands of years, kings and rulers took advantage of this
fact by putting lookout posts atop the tallest mountains to survey more of their realm
and fend off would-be attackers. Throughout history, many battles have been fought
to “take the high ground.” Space takes this quest for greater perspective to its
ultimate end. From the vantage point of space, we can view large parts of the Earth’s
surface. Orbiting satellites can thus serve as “eyes in the sky” to provide a variety of
useful services.

3
STK Astronautics Primer

Figure 1: Global perspective. From space, satellites can observe large-scale features
on the Earth, track weather patterns, monitor the environment and view widely separated
points simultaneously, allowing them to communicate.

Universal Perspective

Space offers a clear view of the heavens. When we look at stars in the night sky, we
see their characteristic twinkle. This twinkle, caused by the blurring of “starlight” as it
passes through the atmosphere, is known as scintillation. Not only is the light blurred,
but some of it is blocked or attenuated altogether. This attenuation is frustrating for
astronomers who need access to all the regions of the spectrum to fully explore the
universe. By placing observatories in space, we can sit above the atmosphere and
gain an unobscured view of the universe, as depicted in Figure 2. The Hubble Space
Telescope and the Gamma Ray Observatory are armed with sensors operating far
beyond the range of human senses. Already, results from these instruments are
revolutionizing our understanding of the cosmos.

Figure 2: Seeing beyond the clouds. Earth-based astronomy is obscured by the


atmosphere. Astronomers don’t like the “twinkle” of star light. Some wavelengths are

4
STK Astronautics Primer

completely blocked. Space-based astronomy opens the door to a whole new perspective on
the universe.

A Unique Environment

Space offers a unique environment with many advantages.

♦ Free-fall environment—enables developing advanced materials


♦ Abundant resources—solar energy and minerals

Gravity makes some manufacturing processes difficult if not impossible. To form


certain new metal alloys, for example, we must blend two or more metals in just the
right proportion. Unfortunately, gravity tends to pull the heavier metal to the bottom,
making a uniform mixture difficult to obtain. But space offers the solution. A
manufacturing plant in orbit is literally falling toward Earth but never hitting it. This
is a condition known as free-fall (NOT zero gravity). In free-fall there are no contact
forces on an object, so it is said to be weightless. Unencumbered by the weight felt on
the Earth’s surface, factories in orbit can create exotic new metals for computers or
other advanced technologies, as well as for promising new pharmaceutical products to
battle disease on Earth.

Figure 3: Early free-fall manufacturing. In the 16th century, Italian weapons makers
developed a secret way of making lead shot for muskets. By dropping liquid lead from a
“shot tower,” they found near-perfect spheres would form as the molten lead cooled and
hardened in free fall.

Space also offers abundant resources. While some on Earth argue about how to carve
the pie of Earth’s resources into smaller and smaller pieces, others have argued that

5
STK Astronautics Primer

example, is known to be rich in oxygen and aluminum. The oxygen could be used as
rocket propellant and for humans to breathe. Aluminum is an important metal for
various industrial uses. These resources, coupled with the human drive to explore,
mean the sky is truly the limit!

Space Applications
Let’s look at some important application of space that affect all of our lives today.

Communications Satellites

Science/Science Fiction writer Arthur C. Clarke first proposed putting satellites into
orbits with periods of 24 hours, 36,000 km above the equator, exactly matching the
rotation rate of the Earth. These geostationary orbits could serve as communication
hubs to link together remote parts of the planet. With the launch of the first
experimental communications satellite, Echo I, into Earth orbit in 1960, Clarke’s
fanciful idea showed promise of becoming reality. Although Echo I was little more
than a reflective balloon in low-Earth orbit, radio signals were bounced off it,
demonstrating that space could be used to broaden our horizons of communication.
An explosion of technology to exploit this idea quickly followed.

Satellites are now used for a large percentage of commercial and government
communications and for most domestic cable television. Through satellite technology,
relief workers can now stay in constant contact with their organizations, enabling
them to better distribute aid to refugees hungry for food. In addition, our modern
military now relies almost totally on satellites to communicate with forces deployed
world-wide. Without satellites, global communication as we know it today would not
be possible.

Remote Sensing Missions

Satellites operating from the global perspective of space have also made possible the
science of remote sensing. Remote sensing is the act of observing Earth and other
objects from space. For decades, military “spy satellites” have kept tabs on the
activities of potential adversaries using remote-sensing technology. This same
technology has been adapted for civilian uses such as
♦ monitoring Earth’s environment
♦ forecasting the weather
♦ managing resources
Satellites can now “spy” on crops, ocean currents, and natural resources to aid
farmers, resource managers, and planners on Earth. In countries where the failure of
a harvest may mean the difference between bounty and starvation, spacecraft have
helped planners manage scarce resources and head off potential disasters before
insects or other blights could wipe out an entire crop. Weather forecasting is a further
application of remote-sensing technology—one we’ve all come to rely on. Overall, we’ve
come to rely more and more on the ability to monitor and map our entire planet. As

6
STK Astronautics Primer

Satellites can now “spy” on crops, ocean currents, and natural resources to aid
farmers, resource managers, and planners on Earth. In countries where the failure of
a harvest may mean the difference between bounty and starvation, spacecraft have
helped planners manage scarce resources and head off potential disasters before
insects or other blights could wipe out an entire crop. Weather forecasting is a further
application of remote-sensing technology—one we’ve all come to rely on. Overall, we’ve
come to rely more and more on the ability to monitor and map our entire planet. As
the pressure builds to better manage scarce resources and assess environmental
damage, we’ll call upon remote-sensing spacecraft to do even more.

Figure 4: Satellite remote sensing. From the vantage point of space, we can plan
urban development and plot the course of dangerous storms.

Space-based Navigation

Early seafarers looked to the stars to guide their way. Modern seafarers look only as
far as satellites in Earth orbit. Systems such as the Global Positioning System (GPS),
developed by the U.S. military, tell you where you are, in what direction you’re
heading and how fast you’re going.

7
STK Astronautics Primer

Figure 5: Global Positioning System (GPS). GPS allows Earth-based users armed
with a simple, hand-held receiver to triangulate from a constellation of 24 satellites. They
can then determine their location to within a few meters and velocity, and a few m/sec
anywhere on Earth.

Exploration

Perhaps the most exciting missions are those which explore the unknown. Missions
such as the Magellan spacecraft that orbited Venus with a powerful radar to peel back
the clouds of this once mysterious planet are a good example. A computer-enhanced
image taken by Magellan is shown in Figure 6. These types of missions push back the
boundaries of human knowledge, giving us new insight into planetary formation,
weather and other important processes at work back here on Earth.

8
STK Astronautics Primer

Figure 6: Magellan at Venus. The powerful synthetic aperture radar on NASA’s


Magellan spacecraft pierced the thick clouds of Venus, giving us the first details of the
planet’s surface. (Photo courtesy of NASA.)

Describing Space Missions


Space missions seem complex, and they are to a certain extent, but if you look at
them logically, you’ll see many similarities. Let’s begin with some key definitions:

♦ Mission Objective - Why we’re going to space and what we’re going to do once
we get there.
♦ Users - The people or systems that use data or services provided by the
satellite or satellites.
♦ Operators -The people who manage and run the mission from the ground.
♦ Mission Operations Concept -How users, operators, ground and space
elements all work together to make a mission successful.

All these come together to form the tangible elements of what is collectively called the
Space Mission Architecture. These elements are depicted in Figure 7; each one is
defined in the subsections following.

9
STK Astronautics Primer

Communications Network

The Space Mission Trajectories & Orbits

Space Transportation
Payload Bus
Space Operations

Payload

Spacecraft

Figure 7: Space mission architecture. The key to understanding how missions are
built is to look at the space mission architecture that includes these critical elements.

Space Operations
The term space operations encompasses all activities needed to monitor and control
satellites and the other elements that make up a space mission. Space operations are
performed by teams of people located at tracking sites and control centers around the
world.

Spacecraft Bus and Payload


A spacecraft has two basic parts, a payload (or payloads) and a bus. The payload includes
space-borne people and instruments that perform the primary mission. The spacecraft bus
provides for the care and feeding of the payload—pointing, heating and cooling, structure,
transportation and power. A simple analogy of a spacecraft bus and its payload is a good
old-fashioned school bus, as shown in Figure 8. It contains all of the same types of
systems needed to support a spacecraft.

10
STK Astronautics Primer

Body & frame


Horn, radio & driver (Structures)
(communications & data handling)

Steering (space vehicle control)

Radiator, air conditioning &


heater (environment control &
life support )
engine & drive train
Battery & alternator (electrical power) (space transportation)

Passengers
(payload)

Figure 8: The spacecraft “bus.” A spacecraft has all the basic systems found in a
regular school bus.

Trajectories and Orbits


A trajectory is any path an object follows through space. An orbit is a special type of
repeating trajectory. The simplest way to imagine an orbit is to think of a “racetrack”
around the Earth which satellites “drive” around, as shown in Figure 9.

Figure 9: Orbit racetrack. An orbit can be thought of as a fixed racetrack around a


planet, where the size of the racetrack depends on the velocity of the object in orbit.

Depending on the altitude of the orbit, a satellite has different perspectives on the Earth.
The total fraction of the Earth a satellite can “see” using its onboard sensors is known as
the field of view. The projection of this field of view onto the Earth’s surface creates a
swath width for the sensor as it sweeps around the Earth on its orbit. These two
parameters are illustrated in Figure 10.

11
STK Astronautics Primer

field of view
swath width

Figure 10: Field of view and swath width. The height of the orbit and the sensor field
of view dictates the swath width that can be imaged on the ground.

There are a variety of different types of orbits that can be found in a typical space mission,
including parking orbits, transfer orbits and final mission orbits. These are illustrated in
Figure 11.

fin
al parking orbit
or
bit
r orbit
transfe

Figure 11: Orbit types. Different types of orbits include the parking orbit, the transfer
orbit and the final or mission orbit. A satellite normally begins its life in a temporary
parking orbit. From there, an upper stage rocket is used to boost the satellite onto a
transfer orbit. An additional boost places it into the final mission orbit.

Space Transportation
Space transportation includes all of the systems necessary to deliver our spacecraft to its
final mission orbit. Normally, this consists of a booster, such as the Space Shuttle or

12
STK Astronautics Primer

Ariane, an upper stage, such as the Inertial Upper Stage (IUS), and onboard thrusters for
final maneuvers and station keeping. The Space Shuttle, shown in Figure 12, is one type
of complete space transportation system.

Figure 12: The Space Shuttle. Space transportation includes the systems that put the
spacecraft in orbit, keep it there, and rotate and move it if necessary. Space
transportation systems develop the velocity needed to obtain and stay in orbit. Space
boosters are divided into stages that provide incremental changes in velocity and are
then discarded.

Communications Network

A space mission is more than just rockets and satellites. An entire system of ground and
on-orbit assets are needed to track, command and control all aspects of the mission. This
communications network ties together various links needed to deliver bus telemetry and
payload data to operators and users, as shown in Figure 13.

13
STK Astronautics Primer

Tracking & Data


Relay Satellite relay satellite
(TDRS)

primary aircraft

tracking site
mission control center tracking site/users

Figure 13: Communications network. The communications network is the “glue”


that holds the mission together. The network ties together space assets, ground
controllers and users in a complex web of links that transfers data among the various
mission element nodes.

14
STK Astronautics Primer

EXPLORING SPACE

Long before rockets and interplanetary probes escaped the Earth’s atmosphere,
people explored the heavens with just their eyes and imagination. Later, with the aid
of telescopes and other instruments, humans continued their struggle to bring order
to the heavens. With order came some understanding and a concept of our place in
the universe. Thousands of years ago, priests of ancient Egypt and Babylon carefully
observed the heavens to plan religious festivals, to control the planting and harvesting
of various crops, and to understand at least partially the realm in which they believed
many of their gods lived. Later, philosophers such as Aristotle and Ptolemy developed
complex theories to explain and predict the motions of the Sun, Moon, planets and
stars.

The theories of Aristotle and Ptolemy dominated the world of astronomy and our
understanding of the heavens well into the 1600s. Combining ancient traditions with
new observations and insights, natural philosophers such as Copernicus and Kepler
offered rival explanations from the 1500s onward. Using their models and Isaac
Newton’s new tools of physics, astronomers in the 1700s and 1800s made several
startling discoveries, including two new planets—Uranus and Neptune. Let’s briefly
explore some of these major contributors to our early understanding of space and
orbits.

Copernicus

With the Renaissance and humanism came a new emphasis on the accessibility of the
heavens to human thought. Nicolaus Copernicus (1473–1543), a Renaissance
humanist and Catholic clergyman, reordered the universe and enlarged man’s
horizons. He placed the Sun at the center of the solar system, as shown in Figure 14,
and had the Earth rotate on its axis once a day while revolving about the Sun once a
year.

Copernicus further observed that, with respect to a viewer located on the Earth, the
planets occasionally appear to back up in their orbits as they move against the
background of the fixed stars. Ptolemy and others resorted to complex combinations
of circles to explain this backward motion of the planets, but Copernicus cleverly
explained that this motion was simply the effect of the Earth overtaking, and being
overtaken by, the planets as they all revolved about the Sun.

However, Copernicus’ heliocentric system had its drawbacks. He couldn’t prove the
Earth moved, and he couldn’t explain why the Earth rotated on its axis while
revolving about the Sun. He also adhered to the Greek tradition that orbits follow
uniform circles, so his geometry was complex and somewhat erroneous. In addition,
Copernicus wrestled with the problem of parallax—the apparent shift in the position
of bodies when viewed from different distances. If the Earth truly revolved about the
Sun, critics observed, a viewer stationed on the Earth should see an apparent shift in

15
STK Astronautics Primer

position of a closer star with respect to its more distant neighbors. Because no one
saw this shift, Copernicus’ Sun-centered system was suspect. In response,
Copernicus speculated that the stars must be at vast distances from the Earth, but
such distances were far too great for most people to contemplate at the time, so this
idea was also widely rejected.

Figure 14: Copernicus redefines the center. Polish astronomer Copernicus reordered
our view of the universe. He promoted a heliocentric (Sun-centered) universe, a simpler,
more symmetric approach with all of the planets in circular orbits about the Sun.
Unfortunately, these ideas were widely rejected because they disputed religious
teachings of his day.

Kepler

Johannes Kepler (1571 - 1630) revolutionized our understanding of orbits. In Cosmic


Mystery, written before he was age 25, he calculated that the orbit of Mars was not
circular but elliptical. From this work, he developed three important laws of orbit
motion, described in Figures 15 through 17.

st
Figure 15: Kepler’s 1 law. The orbits of the planets are ellipses with the Sun at a
focus.

16
STK Astronautics Primer

Planetary Planetary
motion over Area 1 Area 2 motion over
30 days 30 days

Area 1 = Area 2

Figure 16: Kepler’s 2nd law. The orbits of the planets sweep out equal areas in equal
time.

Average distance

Figure 17: Kepler’s 3rd law. The square of the orbit period—the time it takes to go
around once—is proportional to the cube of the average distance to the Sun.

Galileo

In 1609, an innovative mathematician, Galileo Galilei (1564–1642), heard of a new


optical device that could magnify objects so they would appear to be closer and
brighter than when seen with the naked eye. Building a telescope that could magnify
an image 20 times, Galileo ushered in a new era of space exploration. He made some
startling telescopic observations of the Moon, the planets, and the stars, thereby
attaining stardom in the eyes of his peers and potential patrons. Observing the
planets, Galileo noticed that Jupiter had four moons or satellites (a word coined by
Kepler in 1611) that moved about it. This disproved Aristotle’s claim that everything
revolved about the Earth.

17
STK Astronautics Primer

Galileo also took on Aristotle’s physics. He rolled a sphere down a grooved ramp and
used a water clock to measure the time it took to reach the bottom. He repeated the
experiment with heavier and lighter spheres, as well as steeper and shallower ramps,
and cleverly extended his results to objects in free-fall. Through these experiments,
Galileo discovered, contrary to Aristotle, that all objects fall at the same rate
regardless of their weight, as shown in Figure 18.

Galileo further contradicted Aristotle as to why objects, once in motion, tend to keep
going. Aristotle held that objects in “violent” motion, such as arrows shot from bows,
keep going only as long as something is physically in touch with them, pushing them
onward. Once this push dies out, they resume their natural motion and drop straight
to Earth. Galileo showed that objects in uniform motion keep going unless disturbed
by some outside influence. He wrongly held that this uniform motion was circular,
and he never used the term “inertia.” Nevertheless, we applaud Galileo today for
greatly refining the concept of inertia as we know it today.

Figure 18: Galileo on gravity. Through application of the scientific method, Galileo put
Aristotle’s ideas to the test and proved Aristotle wrong—all objects fall at the same rate.

Newton

To complete the astronomical revolution, which Copernicus had almost unwittingly


started and which Kepler and Galileo had advanced, the terrestrial and heavenly
realms had to be united under one set of natural laws. Isaac Newton (1642–1727)
answered this challenge. Newton was a brilliant natural philosopher and
mathematician who provided a majestic vision of nature’s unity and simplicity. 1665
proved to be Newton’s “miracle year,” in which he significantly advanced the study of
calculus, gravitation, and optics. Extending the groundbreaking work of Galileo in
dynamics, Newton published his three laws of motion and the law of universal
gravitation in the Principia in 1687. With these laws, you could explain and predict
motion not only on Earth but also in tides, comets, moons, planets—in other words,
motion everywhere. Newton’s laws are explained more thoroughly in the next section.

18
STK Astronautics Primer

INTRODUCTION TO ORBITAL MOTION

Orbits are one of the basic elements of any space mission. Understanding a satellite
in motion may at first seem rather intimidating. After all, to fully describe orbital
motion we need some basic physics along with a healthy dose of calculus and
geometry. However, as we’ll see, the complex trajectories of rockets flying into space
aren’t all that different from the paths of baseballs pitched across home plate. In fact,
in most cases, both can be described in terms of the single force pinning you to your
chair right now—gravity.

Armed only with an understanding of this single pervasive force, we can predict,
explain and understand the motion of nearly all objects in space, from baseballs to
entire galaxies. Once we know an object’s position and velocity, as well as the nature
of the local gravitational field, we can predict exactly where the object will be minutes,
hours or even years from now.

Overview

Math for the Faint-hearted


♦ What is a vector?
♦ What is a derivative?
♦ What is an integral good for?

Key Concepts in Dynamics


♦ What is the difference between mass, inertia and weight?
♦ What is momentum?
♦ What is energy?
♦ What are Newton’s Laws of Motion?
♦ What is gravity?

Orbits Made Simple


♦ OK, forget the math, what is an orbit, really?
♦ How are energy and momentum conserved in an orbit?

Math for the Faint-hearted


Before delving too deeply into a discussion of dynamics, orbital mechanics and
propagators it’s useful to step back and briefly review a bit of math. BUT DON’T
PANIC! This section is designed to simplify a few basic concepts and describe the

19
STK Astronautics Primer

notation used. You should walk away from this section with a “Math Survival Kit”
that will make the rest of this Primer far more useful and enjoyable.

Vectors and Such

Scalar

A scalar is a quantity that has magnitude only. Speed, energy and temperature
are examples of scalars. None of these quantities has a unique meaning in any
certain direction. A single letter, such as E for Total Mechanical Energy, denotes a
scalar quantity.

Vector

A vector is a quantity that has both magnitude and direction. For example, if I ask
you where you drove in your car, you might answer: “I went south.” But this
wouldn’t tell me much. If I asked “How far?,” and you said “five miles,” I could put
together a better picture. By knowing you drove five miles south, I have both
v
magnitude and direction. A letter with an arrow over it, such as V for the velocity
vector is used to denote a vector quantity.

Unit Vector

A unit vector is a vector having a magnitude of one; it’s used to describe direction
only. For example, if I want to define where north is on a drawing, I could do so
with a unit vector indicating the direction. A letter with a caret or hat over it, such
as I$ for the I-direction denotes a unit vector.

Example

We generally describe the velocity of an object in orbit in terms of three unit


vectors, I$, J$ , K$ . Thus, a typical velocity vector could be written as:
v
. I$ + 21
V = 30 . J$ − 7.4 K$ km / sec

This means the velocity is 3.0 km/sec in the I-direction, 2.1 km/sec in the J-
direction and 7.4 km/sec in the K-direction.

Calculus—Just the important bits

Calculus was developed to analyze changing parameters. Let’s look at how those
changes are described.

20
STK Astronautics Primer

Derivative

A derivative represents the rate of change of one parameter with respect to


another. For example, if you’re traveling north in your car, your position is
changing over time. The rate at which your position changes over time is your
velocity. Thus, if you travel 30 miles north in 30 minutes, your velocity is:
v
v Change in position (R) 30 miles
Velocity = V = = = 60 mph north
Change in time (t) 30 minutes

Several methods are commonly used to denote a vector. In this primer, we use two
types of notation. The first is to represent a derivative as d. So the change in
position over time would be:
v
v dR
V =
dt
We also use a “dot” over a symbol to represent the derivative with respect to time. For
v&
example, R represents the derivative of the position vector or the rate of change of
&&v
position. R is the second derivative of position, or the rate of change of the rate of change
of position (i.e. the acceleration).

Integral

An integral represents the cumulative effect of one parameter changing with


respect to another. If we were to graph both changing parameters, the integral is
the area under the curve. For example, if you drive north at 30 miles per hour for
30 minutes, the integral of this velocity is your new position at the end of the time
(e.g., 15 miles north of where you started). In other words, you add up all the
changes in position over time to get the total change. An integral is the reverse of
a derivative. Because acceleration is the derivative of velocity over time, the
integral of acceleration over time is velocity.

Key Concepts

Mass

Mass is a measure of how much matter or “stuff” an object possesses. For example, a
volleyball and a cannon ball are about the same size, but the cannon ball has far
more mass because it is made of a more dense material.

Inertia

Inertia describes how hard it is to move an object. It is much easier to push a baby
carriage than a bulldozer because the bulldozer, being more massive, has more
inertia.

21
STK Astronautics Primer

Weight

Weight actually describes the force produced by gravity acting on a mass. Your weight
in various situations is illustrated in Figure 19.

Figure 19: Weight. Weight describes a contact force caused by the effect of gravity on
mass. On Earth, your weight is one value. As you move further from the center of the
Earth, say in a penthouse at the top of a 250-mile-high skyscraper, your weight would be
slightly less. In orbit at 250 miles altitude, the gravitational force is still the same;
however, because you are in free-fall and not in contact with the Earth, your weight is
zero. In all cases, mass stays the same.

Momentum

Linear momentum describes the resistance a moving object has to changes in either
direction or speed. The more massive an object, or the faster it is moving, the harder
it is to stop or change its direction of motion. As a result, linear momentum is the
product of the mass and velocity of an object. Momentum for baby carriages and
bulldozers is shown in Figure 20.

m = 25 kg m = 25,000 kg
v= 1000 m/s v = 1 m/s
mv = 25,000 kg-m/s mv= 25,000 kg-m/s

Figure 20: Momentum, bulldozers and baby carriages. Linear momentum is the
product of mass and velocity. For a baby carriage to have the same linear momentum as
a bulldozer, it would have to be traveling at a much higher velocity.

22
STK Astronautics Primer

Angular momentum is a measure of the spinning properties of an object. As Figure 21


illustrates, a non-spinning top immediately falls over. However, a spinning top has
angular momentum, which allows it to resist the force of gravity pulling it over (until it
finally slows down due to friction).

angular momentum

non-spinning top spinning top

Figure 21: Angular momentum. A spinning top has angular momentum which keeps it
pointing upright even when pulled by outside forces such as gravity.

Energy

Potential energy is a function of an object’s position and mass. The greater the height
an object is raised to, the greater its potential energy. An object at the top of a deep
well, as shown in Figure 22, has more potential energy than an object at the bottom of
the well.

PE = 0 at R = infinite

R PE < 0 at R > 0

PE < 0 at R = 0

Figure 22: Potential energy. Because we normally define our coordinate systems as
positive outward from the center of the Earth, we measure potential energy “from the
bottom up.” For example, at the top of a deep well we would say the potential energy is
zero. As we get closer to the bottom of the well, the potential energy is less, or more
negative.

Kinetic energy is a function of an object’s mass and speed. Like momentum, the more
massive an object is or the faster it travels, the more kinetic energy it has. It is this

23
STK Astronautics Primer

energy that must be transformed in order to stop a speeding bulldozer, as shown in


Figure 23. This is accomplished by applying the brakes, which turns kinetic energy
into heat (those brakes get hot!).

m = 25,000 kg
v = 1 m/s
1/2 mv2 = 12,500 kg-m/s2

Figure 23: Newton’s 1st law. The kinetic energy of a bulldozer moving at only 1 m/s is
12,500 kg-m/s2.

Total energy is the sum of kinetic plus potential energy.

Total Energy = Kinetic Energy + Potential Energy

Newton’s Laws

Newton’s First Law

A body remains at rest or in constant motion unless acted upon by external forces.

In other words, if you were to pitch a baseball, it should continue on its path, in a
straight line forever, unless disturbed by an outside force such as gravity or air
resistance.

Newton’s Second Law

The time rate of change of an object’s momentum is equal to the applied force.

Change(momentum)
= Force Applied
Change(time)

Recall, momentum is the product of mass and velocity. Thus, as long as mass
stays constant (which it normally does as long as rockets aren’t firing) this
equation can be reduced to:

F = ma
or:

Force applied = mass (m ) times acceleration (a)

24
STK Astronautics Primer

The significance of this relationship can be felt every time you hit the brakes in you car.
The more force you apply (the harder you hit the brakes) the faster you stop (the faster
you decelerate). This principle is illustrated in Figure 24.

25,000 N
1 m/s

stops in 1
second

6.9 N
1 m/s

stops in 1 hour

Figure 24: Newton’s 2nd law. Force is proportional to acceleration (or deceleration). A
25,000 N force is needed to stop a 1 m/s bulldozer in 1 second, while much smaller 6.9
N force would take 1 hour to bring it to a stop.

Newton’s Third Law

For every action, there’s an equal and opposite reaction. This basic principle can
be illustrated by two roller-skating astronautics, as shown in Figure 25.

Figure 25: Newton’s 3rd law. If two people on roller skates push against each other,
they both move backward. Their acceleration is proportional to their mass.

25
STK Astronautics Primer

Newton’s Law of Universal Gravitation

The force of gravity between two bodies is proportional to the product of the
masses and inversely proportional to the square of the distance between them.
This is illustrated in Figure 26.

F
g m
F 2
g

m
1

Figure 26: Gravitational attraction. Two masses in space each exert a force on the
other. The magnitude of this force depends on the product of their masses and the
square of the distance between them.

Newton’s law can be summarized in equation form as follows:

GM1 M 2
Fgravity =
R2

where

Fgravity = Force of gravity (N)

G = Universal gravitational constant = 6.67 x 1011 (Nm/kg)

M1, M2 = Mass 1 and Mass 2 respectively (kg)

R = Distance between the two masses (m)

In other words, the more mass an object has, the more gravitational force it
generates. Furthermore, the farther apart two objects are, the less the force is, in
fact, the force decreases with the square of the distance as illustrated in Figure 27.

26
STK Astronautics Primer

R
2R

F
g m Fg /4 m
F 2 2
g
F /4
g
m
1 m
1

Figure 27: Gravity and distance. The force of gravitational attraction decreases with
the square of the distance (e.g., if you double the distance, the force decreases by one
fourth).

What exactly is gravity? The study of physics is still grappling to reconcile the
force of gravity with the other fundamental forces of nature. Already, extremely
weak “gravity waves” have been detected from distant galaxies. More sensitive
instruments are being built to understand and quantify this mysterious force.

How strong is gravity? Let’s look at the Earth-Moon system. The force of gravity
between the Earth and Moon is 1.98 x 1020 Newtons! To put this into perspective,
the Space Shuttle generates about 28 million Newtons thrust at lift-off. The Earth-
Moon gravity force is more than one trillion times as great as that of the Shuttle!

Regardless of what gravity really is, we know it’s a force that affects anything with
mass (and that’s pretty much everything!). While Galileo right in that the
gravitational force is greater on heavier objects than lighter objects, he was wrong
in predicting the affect this would have on the rate at which they fall. The
acceleration of an object in a gravitational field is independent of its mass.

Figure 28: All things fall at the same rate. It is important to note that all things
accelerate at the same rate within a gravitational field. For example, if you drop a

27
STK Astronautics Primer

hammer and a feather, both objects impact the ground at the same time (neglecting air
resistance). Of course, Galileo predicted this. Astronaut Dave Scott proved this with an
experiment on the Moon. He dropped a hammer and a feather at the same time. Both hit
the ground at the same instant (there is no air resistance on the Moon)!

Orbits Made Simple


What is an orbit? In the simplest sense, orbits are a type of “racetrack” in space that a
satellite “drives” around.

Figure 29: Orbits as racetracks. The simplest way to think of orbits is as giant, fixed
“racetracks” on which spacecraft “drive” around the Earth.

Baseballs and Satellites

But what makes these racetracks? Before diving into a complicated explanation, let’s
begin with a simple experiment that illustrates, conceptually, how orbits work. To do
this, we’ll arm ourselves with a bunch of baseballs and travel to the top of a tall
mountain. Imagine you were standing on top of this mountain prepared to pitch
baseballs to the east. As the balls sail off the summit, what would you see? The
baseballs would follow a curved path before hitting the ground. Why is this? The force
of your throw causes them to move outward, but the force of gravity pulls them down.
Therefore, the “compromise” shape of the baseball’s path is a curve.

28
STK Astronautics Primer

Figure 30: Baseballs in motion. A baseball-throwing astronaut can be used to illustrate


the simple motion of a satellite. Naturally, the harder the baseballs are thrown, the
further they travel before hitting the ground.

As Figure 30 illustrates, the faster you throw the balls, the farther they travel before
hitting the ground. This could lead you to conclude that the faster you throw them,
the longer it takes before they hit the ground. But is this really the case? Let’s try
another experiment to see. As you watch, two astronauts, standing on flat ground,
release baseballs. The first one simply drops a ball from a fixed height. At exactly the
same time, a second astronaut throws an identical ball horizontally as hard as
possible. What do you see? If the second astronaut throws a fast ball, it travels out
about 20 m (60 ft.) or so before it hits the ground. However, the ball dropped by the
first astronaut hits the ground at exactly the same time as the pitched ball, as Figure
31 shows!

Figure 31: Motion and gravity. Two astronauts each have a baseball held at the same
height above the ground. If the first astronaut drops her baseball while the second
astronaut throws his, both baseballs hit the ground at exactly the same time. Gravity acts
on both baseballs in the same way, independent of their horizontal motion.

29
STK Astronautics Primer

How can this be? To understand this seeming paradox, we must recognize that, in
this case, the motion in one direction is independent of motion in another. Thus,
while the second astronaut’s ball moves horizontally at 30 km/hr (20 m.p.h.) or so,
it’s still falling at the same rate as the first ball. This rate is the constant gravitational
acceleration of all objects near the Earth’s surface: 9.798 m/s2. Thus, they hit the
ground at the same time. The only difference is that the pitched ball, because it has
horizontal velocity, manages to travel some distance before intercepting the ground.

Now let’s return to the top of our mountain and start throwing our baseballs faster
and faster to see what happens. No matter how fast we throw them, the balls still fall
at the same rate. However, as we increase their horizontal velocity, they’re able to
travel farther and farther before they hit the ground. Because the Earth is basically
spherical in shape, something interesting happens. The Earth’s spherical shape
causes the surface to drop approximately 5 m for every 8 km we travel horizontally
across it, as shown in Figure 32.

8 km
5m

Figure 32: Our spherical Earth. We know the Earth is a nearly perfect sphere. For
every 8 km of horizontal distance, the Earth curves down about 5 m. In other words, if
you could lay an 8 km long board tangent to the Earth at one end, at the other end it
would be 5 m off the ground.

So, if we were able to throw a baseball at 8 km/s (assuming no air resistance), its
path would exactly match the rate of curvature of the Earth. That is, gravity would
pull it down about 5 m for every 8 km it travels, and it would continue around the
Earth at a constant height. If we don’t remember to duck, it will hit us in the back of
the head about 85 minutes later. (Actually, because of the rotation of the Earth, it
would miss your head.) A ball thrown at a speed slower than 8 km/s falls faster than
the Earth curves away beneath it.

If we analyze our various baseball trajectories, we see a range of different shapes.


Only at exactly one particular velocity do we get a circular trajectory. Any slower than
that and our trajectory impacts the Earth at some point. If we were to project this
shape into the Earth, we’d find the trajectory we see is really just a piece of an ellipse.
If we throw the ball a bit harder than the circular velocity, we also obtain an ellipse.
An object in orbit is literally falling around the Earth but, because of its horizontal
velocity, it never quite impacts the ground. If we throw the ball too hard, it leaves the
Earth altogether on a parabolic or hyperbolic trajectory, never to return.

30
STK Astronautics Primer

hyperbola

parabola

circle

ellipse

Figure 33: Baseballs in orbit. If we throw a baseball fast, but not quite fast enough,
eventually the ball will impact the ground (perhaps even on the other side of the Earth)
like an ICBM (Intercontinental Ballistic Missile). If we throw it at just the right speed,
gravity will cause the ball to fall 5 m for every 8 km of horizontal distance traveled. But,
since the Earth also curves down 5 m for each 8 km of horizontal distance, the ball will
stay at same the instantaneous height above the ground. We call this a circular orbit. If
we throw it faster than the circular orbit speed, the ball will be in an elliptical-shaped orbit.
If we throw the ball faster yet, it will escape the Earth’s gravity altogether on a parabolic or
hyperbolic trajectory.

Thus, it is important to note that no matter how hard we throw, our trajectory
resembles either a circle, ellipse, parabola or hyperbola. These four shapes are called
conic sections. Why conic sections? Because these are the shapes we get by slicing
through a cone with a plane at different angles, as illustrated in Figure 34.

circle ellipse hyperbola parabola

Figure 34: Conic sections. The four basic conic sections: circle, ellipse, hyperbola and
parabola.

31
STK Astronautics Primer

So how fast how fast do we have to throw our baseball to put it into a circular orbit?
Let’s play with some math. The velocity of a satellite in a circular orbit can be found
using:

GM Earth
Vcircular =
R

where

Vcircular = Satellite velocity in a circular orbit (km/sec)

G = Universal gravitational constant = 6.67 x 10-11 km2/sec3

MEarth = mass of the Earth = 5.98 x 1015 kg

R = distance from the center of the Earth = 6378 km at the surface

Thus, at the surface of the Earth, the velocity in a circular orbit would be 7.9 km/sec
(17,600 mph)! In other words, to move into a circular orbit that stays just above the
surface of the Earth (ignoring air drag) you’d have to throw the baseball at 17,600
mph. Notice that the circular orbit speed depends on your distance from the center of
the Earth. The lower you are, the faster you must travel to achieve a circular orbit.

Conservation of Energy & Momentum

Now that we’ve looked at the simple geometry of an orbit, we can consider how
conservation of energy and momentum affects the velocity of satellites. Gravity is a
conservation force, which means that an object moving in a gravitational field doesn’t
lose any energy through friction or heat, etc. Additionally, total energy is constant, or:

Total Energy = Kinetic (KE) + Potential Energy (PE) = constant

This basic principle is illustrated by the swinging astronaut in Figure 35.

32
STK Astronautics Primer

Maximum PE Maximum PE
KE=0 KE=0

Max KE
Min PE

PE+KE=constant

Figure 35: Trading kinetic and potential energy. The conservation of energy
(Potential Energy + Kinetic Energy = Constant) is illustrated by a simple swing.
Neglecting losses from friction, the total energy of the astronaut on the swing is constant.
At the low point in the swing, speed (kinetic energy) is greatest and potential energy is
lowest. As you swing, you trade kinetic energy (speed) for potential energy (height) with
the sum of the two constant.

When applied to an orbit, the same principle applies. Total energy must be conserved,
thus the orbit speed varies throughout the orbit as kinetic energy is traded for
potential energy. A satellite travels fastest at perigee—the lowest point in the orbit—
and slowest at apogee—the highest point in the orbit. This is shown in Figure 36.

high kinetic
low kinetic energy
energy

apogee perigee

high potential low potential


energy energy

Figure 36: Energy conservation in orbit. As a satellite moves closer to the Earth in an
orbit, it must speed up to conserve total energy. As it gets further away, the satellite
trades kinetic energy for potential energy and slows down.

Angular momentum is also always conserved. Ice skaters use this principle to speed
up or slow down as they spin, as shown in Figure 37. Just like a spinning ice skater,
an orbit has angular momentum. Because angular momentum is a vector quantity,

33
STK Astronautics Primer

the direction as well as the magnitude of this momentum stays the same. As a result,
even though the Earth rotates under the orbit and the Earth (and the orbit along with
it) moves around the Sun, the orbit itself stays fixed in respect to a stationary
reference.

Figure 37: Conservation of angular momentum. Ice skaters use this principle to
speed up or slow down as they spin. When their arms are extended, the moment of
inertia is low, so they spin more slowly. As they draw their arms in, the moment of inertia
decreases and the spin rate increases to keep total angular momentum constant.

34
STK Astronautics Primer

DESCRIBING ORBITS

Overview

Understanding Coordinate Systems


♦ What is a coordinate system?

The Geocentric-Equatorial Coordinate System


♦ What is the most common coordinate system used for satellites?

Classical Orbital Elements


♦ What do all those Greek letters tell me about an orbit?

Ground Tracks
♦ How do those squiggly lines on a map represent the path of a satellite?

Understanding Coordinate Systems


To be valid, Newton’s laws must be expressed in an inertial frame of reference,
meaning a frame that is not accelerating. Any reference frame is just a collection of
definitions that allow us to describe positions and velocities in a more meaningful
way. For example, if I simply told you a car is traveling south, you wouldn’t have very
much information. But if I first tell you that our a reference frame is centered on
Washington, DC, and then tell you that the car is 30 miles east of the city traveling
south at 60 mph, you’d know something far more useful.

In defining coordinate systems and describing position and velocities, we make


extensive use of vectors. We use vectors because we want to keep track of the
information contained in a particular parameter. Specifically, a vector is a parameters
having both magnitude and direction. For example, 60 mph tells you speed
(magnitude) without direction. But 60 mph south tells you both magnitude and
direction. In defining coordinate systems, we are sometimes only interested in
direction. In that case, we use unit vectors defined to have a specific direction and a
magnitude of 1.

Cartesian coordinate systems are laid out with three orthogonal unit vectors—vectors
at right angles to each other. For example, if the origin of a coordinate system were
chosen to be in one corner of a room, the floor would be the fundamental plane. We
could then describe the position of every piece of furniture in the room with respect to
this system. Such a collection of unit vectors allows us to establish the components of
other vectors in 3-D space.

35
STK Astronautics Primer

To create a coordinate system, we need to specify four pieces of information—an


origin, a fundamental plane, a principle direction, and a third axis, as shown in
Figure 38. The origin defines a physically identifiable starting point for the coordinate
system. The other two parameters fix the orientation of the frame. The fundamental
plane contains two axes of the system. Once we know the plane, we can define a
direction perpendicular to that plane. The unit vector in this direction at the origin is
one axis. Next, we need a principle direction within the plane. Again, we choose
something that is physically significant, like a star. Now that we have two directions,
the principle direction and an axis perpendicular to the fundamental plane, we can
find the third axis using the right-hand rule. The right-hand rule can be demonstrated
by pointing the thumb of your right hand in the direction perpendicular to the
fundamental plane. With you fingers pointing in the principle direction, curl your
fingers 90° so that your thumb is pointing in the direction of the third axis.

(1) pick origin (2) pick fundamental plane


& perpendicular to it

fundamental
plane

origin origin

(3) pick principal direction (4) find 3rd axis

fundamental
fundamental plane
plane
origin 3rd axis, found
origin using right-hand
rule

principal direction principal direction


Figure 38: Defining a coordinate system.

Remember—coordinate systems are defined to make our lives easier. If we choose the
correct coordinate system, developing the equations of motion can be simple. If we
choose the wrong system, it can be nearly impossible.

The Geocentric-Equatorial Coordinate System


For Earth-orbiting spacecraft, we’ll choose a tried-and-true system that we know
makes solving the equations of motion relatively easy. We call this system the
geocentric-equatorial coordinate system, shown in Figure 39. Here’s how it’s defined:

36
STK Astronautics Primer

Origin

The center of the Earth (hence the name geocentric).

Fundamental plane

Earth’s equator (hence geocentric-equatorial), where perpendicular to the fundamental


plane is the direction of the north pole.

Principle direction

Vernal equinox direction, , or the vector pointing to the first point of Aries. The
vernal equinox direction points at the zodiac constellation Aries and is found by
drawing a line from the Earth to the Sun on the first day of spring. While this
direction may not seem “convenient” to you, it’s significant to the astronomers who
originally defined the system. Unfortunately, the vernal equinox direction is not
perfectly constant. The Earth’s orbit precesses around the Sun and the Sun is moving
through the galaxy. Therefore, exactly when this direction is defined is extremely
important for the definition of the system. We can use two ways of defining these
directions. The first is to use the mean or average direction at some point in time. The
other is to use the true position at exactly one specific point in time. Various
combinations of these definitions using different dates gives us several possibilities for
coordinate systems used in STK:

J2000

Defines the mean vernal equinox direction and mean Earth rotation axis on
January 1 of the year 2000 at approximately 12:00:00.00 GMT.

Mean of date

Defines the mean vernal equinox direction and mean Earth rotation axis at the
orbit epoch time (the time for which the orbital elements being used is true).

Mean of epoch

Defines the mean vernal equinox direction and mean Earth rotation axis at the
coordinate epoch time (time at which the coordinate system being used is defined).

True of date

Defines the true vernal equinox direction and true Earth rotation axis at exactly
the orbit epoch time specified.

37
STK Astronautics Primer

True of epoch

Defines the true vernal equinox direction and true Earth rotation axis at the
coordinate epoch time specified.

B1950

Defines the mean vernal equinox direction and mean Earth rotation axis at the
beginning of the Besselian year 1950. It corresponds to 31 December 1949 at
22:09:07.20 Greenwich Mean Time (GMT).

Mean Equinox, True Equator

Defines mean vernal equinox direction and true Earth rotation axis for the orbit
epoch time specified.

Third axis

The third axis of the geocentric-equatorial coordinate system is found using the right-
hand rule.

^
K

^I ^
J

Figure 39: The geocentric-equatorial coordinate system. The system is defined by:
u Origin - Center of Earth; u Fundamental Plane - Equals equatorial plane;
u Perpendicular to Plane - north pole; u Principle Direction - vernal equinox direction.

Classical Orbital Elements


Three pieces of information are needed to fix any point in space; collectively, they’re
r
known as an object’s position vector, R . Three more pieces of information describe its
r
velocity vector, V . One additional item, time, tells us when the information provided
is valid. These elements are know as Cartesian elements. While it is often convenient
to describe orbit motion using simply position and velocity vectors in a Cartesian
coordinate system, especially for computational work, these vectors provide us little
insight into the orbit itself. For this reason, astronomers long ago developed orbital

38
STK Astronautics Primer

elements. Orbital elements give us a short-hand way of expressing orbit size, shape
and orientation, allowing us to tell at a glance the application for a given orbit. This
section describes the Classical orbital elements, which are sometimes referred to as
the Keplerian elements and are attributed originally to Kepler himself. Variations on
these elements, the commonly used two-line element sets, are described in a later
section.

Orbit Size

How big is an orbit? This depends on how fast we “throw” our satellite into orbit. The
faster we throw, the more energy an orbit has, and the bigger it is. We express the size
of an orbit in terms of its semimajor axis, a., as defined in Figure 40.

apogee perigee

2a = major axis

a = semimajor axis

Figure 40: Semimajor axis. The major axis of an elliptical orbit is the distance between
the point of closest approach (perigee) and furthest point (apogee). Semimajor axis is
one-half this distance.

We can express the semimajor axis in terms of the distance from the center of the
Earth to apogee (Rapogee) and perigee (Rperigee). Perigee is the point in an orbit that is
closest to the Earth. Apogee is the point where it is furthest away (apogee is undefined
for a parabolic or hyperbolic trajectory). The semimajor axis can be found using:

Rapogee + R perigee
a=
2

where

a = semimajor axis (km)

Rapogee = Distance from center of Earth to apogee (km)

39
STK Astronautics Primer

Rperigee = Distance from center of Earth to perigee (km)

The orbit’s period, P (i.e., how long the satellite takes to travel around the orbit one
time), is proportional to the orbit size:

a3
P = 2π
GM Earth

For example, a typical Space Shuttle orbit at an altitude of a few hundred kilometers
has a period of about 90 minutes. It orbits the Earth about 16 times each day! For
communications satellites in geosynchronous orbit at an altitude of 35,780 km, the
period is exactly 24 hours.

Orbit Shape

The less circular an orbit is, the more eccentric or “imperfect” it is. Eccentricity, e,
describes the shape of orbit with respect to that of a circle.

e>1
e=1 hyperbola
parabola

ellipse
circle
0<e<1 e=0

Figure 41: Eccentricity. A perfectly circular orbit has an eccentricity of 0. Elliptical


orbits have an eccentricity of less than 1. A parabolic orbit has an eccentricity of exactly
1. Hyperbolic orbits (or trajectories) have eccentricities of greater than 1. In practice, a
perfectly circular or parabolic orbit cannot be achieved.

40
STK Astronautics Primer

Orbit Orientation

How the orbit is tilted with respect to the equator is called its inclination, i. An orbit
that stays directly over the equator has an inclination of 0° and is called an equatorial
orbit. An orbit that goes directly over the north and south poles must have an
inclination of exactly 90° and is called a polar orbit.

Different classes of orbits have different inclinations, as shown in the table.

Inclination Orbit Type Diagram


0° = i = 180° Equatorial
i=0o

i = 90° Polar i=90o

0° < i < 90° Direct or posigrade


(moves in direction of
Earth’s rotation)
ascending
node

90° < i <180° Indirect or retrograde


(moves against the
direction of Earth’s
rotation) ascending
node

We measure how an orbit is twisted by locating its ascending node, the point where
the satellite crosses the equator moving south to north. This point is referenced to the
I-direction, which points in the Vernal Equinox direction. The angle between the I-
direction and the ascending node is called the right ascension of the ascending node,
RAAN, Ω.

41
STK Astronautics Primer

^
K

lp lane
a toria
equ


ascending
node
^
J
^
i

Figure 42: Right Ascension of the ascending node (RAAN), Ω , is the angular
distance from the vernal equinox direction to the ascending node. The ascending node
of an orbit is the point where it crosses the equatorial plane from south to north.

We describe the orbit’s orientation by locating perigee with respect to the ascending
node. This angle is called the argument of perigee, ω; it is measured positive in the
direction of satellite motion.

42
STK Astronautics Primer

^
K

lp lane
a toria
equ
ω

ascending
node
^
J
^
i

Figure 43: Argument of perigee. The argument of perigee, ω, is the angular distance
between the ascending node and perigee.

Finally, we describe a satellite’s instantaneous position with respect to perigee using


another angle, true anomaly, ν. It is the angle, measured positive in the direction of
motion, between perigee and the satellite’s position. Of the six orbital elements, only
true anomaly changes continually (ignoring perturbations).

V
R
υ

perigee

Figure 44: True anomaly. The true anomaly, ν, is the angular distance from perigee to
r
the orbit position vector, R .

43
STK Astronautics Primer

Summary of classical orbital elements

Recall, the reason we wanted to develop the orbital elements in the first place was to
give us a short-hand method of describing an orbit. We also wanted to use parameters
that would have some physical meaning we could more easily visualize. The six
classical orbital elements are summarized below.

Name Symbol Describes


Semimajor axis a Size (and energy)
Eccentricity e Shape (e = 0 for circle, 0> e >1 for
ellipse, e = 1 for parabola, e > 1 for
hyperbola)
Inclination i Tilt of orbit plane with respect to the
equator
Longitude of ascending node Ω Twist of orbit with respect to the
ascending node location
Argument of perigee ω Location of perigee with respect to the
ascending node
True anomaly ν Location of satellite with respect to
perigee

Satellite Missions

As we already know, varying missions require different orbits, which can be described
using Classical orbit elements. The table following shows various missions and their
typical orbits. Technically speaking, a geostationary orbit is a circular orbit with a
period of exactly 24 hours and an inclination of exactly 0°. A satellite in a
geostationary orbit appears to be stationary to an Earth-based observer.
Geosynchronous orbits are slightly inclined orbits with a period of 24 hours. In
practice, it is almost impossible to achieve an orbit with exactly a 24-hour period and
an inclination of 0°. Thus, the two terms are frequently used interchangeably. A semi-
synchronous orbit has a period of 12 hours. Sun-synchronous orbits are retrograde
low-Earth orbits (LEO) inclined 95° to 105°; they are typically used in remote-sensing
missions to observe Earth. A Molniya orbit is a semi-synchronous, eccentric orbit used
for some communication missions. Super-synchronous orbits are usually circular
orbits with periods longer than 24 hours.

Mission Orbit Type Semimajor Period Inclinatio Other


Axis (Altitude) n

♦ Communications Geostationary 42,158 km 24 ~0° e ≅0


hours
♦ Early Warning (35,780 km)
♦ Nuclear detection

♦ Remote sensing Sun- ~6500-7300 km ~90 min ~95° e ≅0


synchronous
(~150-900 km)

44
STK Astronautics Primer

Mission Orbit Type Semimajor Period Inclinatio Other


Axis (Altitude) n

♦ Navigation (GPS) Semi- 26,610 km 12 55° e ≅0


synchronous hours
(20,232 km)

♦ Space Shuttle Low-Earth ~6700 km ~90 min 28.5° or 57° e ≅0


obit (~300 km)

♦ Communication/ Molniya 26,571 km 12 63.4° ω = 270°


Intelligence (RP = 7971 km); hours
e = 0.7
(RA = 45,170 km)

Ground Tracks
Orbital elements allow us to visualize the shape of an orbit around the Earth. Because
we use satellites for missions involving specific points on Earth—taking pictures,
communications, navigation—we really would like to know what path the satellite
traces over the Earth’s surface. A satellite ground track is the orbit path (usually for
multiple orbits) projected onto a flat map of the Earth. These projections become
complex because we must account for the satellite circumnavigating the entire Earth
during each orbit while the Earth itself rotates at 1600 km/sec underneath it.

To visualize a satellite’s ground track, let’s begin by assuming the Earth doesn’t
rotate. Picture an orbit around this nonrotating Earth. Because the orbit plane must
pass through the Earth’s center, the ground track traces a great circle. By definition,
a great circle is any circle on a sphere that can be projected through the center. For
example, all lines of longitude are great circles. The equator is the only line of latitude
that is a great circle—No other line of latitude “slices through” the center of the Earth.

When the Earth is stretched out to a flat-map projection (called an equidistant


cylindrical projection), things start to look different. Imagine yourself on the ground,
watching the orbit pass by overhead. If the Earth didn’t rotate, the projection of the
ground track would always look the same—a sine wave over the surface of the Earth,.
(If you have trouble picturing why it is a sine wave, roll a piece of paper around a
soda can and draw an inclined circle around the can. When you unroll the paper,
you’ll see a sine wave just like an orbit ground track!)

45
STK Astronautics Primer

X
X

Figure 45: Ground track for a nonrotating Earth. If the Earth didn’t rotate, the ground
track would always look like a constant sine wave. As the figure shows, the ground track
would always have the same relative orientation with respect to a stationary observer on
the Earth (shown here as an X in the Pacific ocean).

Now let’s start the Earth rotating again. As you watch the orbit pass overhead,
something happens from one orbit to the next—the ground track shifts to the west!
What happened? The orbit plane is fixed in inertial space. This means the orbit stays
the same with respect to a stationary observer. However, because the Earth rotates at
15° per hour, an observer on the Earth is not stationary. As the Earth (and an Earth-
fixed observer) rotates to the east, the satellite ground track shifts to the west from
one orbit to the next, as shown in Figure 46. The amount it shifts depends on its
period. The longer the period, the more time the Earth has to rotate between
successive orbits.

X
X

Figure 46: Satellite ground tracks. Satellite orbits are fixed in space with respect to a
stationary observer. However, a stationary observer on the Earth is rotating to the east at
15° per hour. Thus, each successive orbit ground track shifts to the west.

Let’s look at the ground track of some very different orbits.

46
STK Astronautics Primer

E C B A

Figure 47: Ground tracks for orbits with different periods. Orbit A: Period=2.67 hr,
Orbit B: Period=8 hr, Orbit C: Period=18 hr, Orbit D: Period=24 hr, Orbit E: Period=24
hr.

47
STK Astronautics Primer

PREDICTING ORBITS

One of the most important problems in mission planning and satellite command &
control is being able to accurately predict orbital motion. To track satellites through
space, we need to know where they are now and where they’ll be later so that we can
predict sensor coverage and point our antennas at them to gather data. Although we
can easily predict this motion when the orbit is a circle, the problem becomes more
complicated when the orbit is an ellipse, and most orbits are at least slightly elliptical.

An orbit propagator is a mathematical algorithm for predicting the future position and
velocity (or orbital elements) of an orbit given some initial conditions and
assumptions. There are a wide variety of orbit propagation techniques available with
widely different accuracy and applications. Knowing the assumptions built into
different propagation schemes is key to knowing which one to use for a given
application.

Overview

Understanding Propagators
♦ How do propagators work?

The Two-Body Propagator


♦ What is meant by a “two-body” propagator?

Orbit Perturbations
♦ What are “J2” and those other things that affect an orbit?

Dealing with Perturbations


♦ How can I model orbit perturbations?

Two-line Element (TLE) Sets


♦ What are TLE set?

STK Propagators
♦ What propagators are available in STK?

Understanding Propagators
To understand the basic problem of orbit propagators, let’s return to the example of
our ball-throwing astronaut shown in Figure 48. What we’re after is a simple

48
STK Astronautics Primer

mathematical algorithm that will allow us to predict the ball’s position and velocity at
any point in time. Whether you’re analyzing the motion of baseballs or galaxies, the
fundamental approach is the same. This motion analysis process has three steps:

♦ define a convenient coordinate system


♦ list simplifying assumptions
♦ define initial conditions

Figure 48: A baseball motion propagator. The simple example of a thrown baseball
can be used to describe the basic problem of orbit propagation.

We can now apply the motion analysis process to describe, and eventually propagate,
the motion of the baseball. First, we select a simple, convenient coordinate system
with its origin at the point of release. The x-direction is defined to be positive to the
right in the picture. The y-direction is positive down. Next, we need to make some
assumptions to make our lives easier. The major assumption we’ll make is that
Earth’s gravity is the only force acting on the ball, a force which is constant over the
flight path we’re concerned with. That is, wind resistance and other forces (the
gravitational pull of the Moon and stars, solar pressure, etc.) are negligible. This
gravitational force can be expressed using Newton’s law, using vector notation, as:
v
Fgravity = mgy$ N

Note y$ is a vector notation indicating the force of gravity acts only in the y-direction, i.e.,
down. See the section describing vectors for further explanation.

49
STK Astronautics Primer

Finally, we need some initial conditions. Let’s pretend that the ball leaves the pitcher’s
hand at a velocity of 10 m/s on a horizontal path (i.e., all motion in the x-direction).
r r
Symbolically we would say the initial velocity ( Vinitial ) and position ( Rinitial ) are:

Vinitial = 10 x$ m / sec

The ball starts out traveling 10 m/sec horizontal to the right

r
Rinitial = 0x$ + 0y$ m

The ball starts out at the origin, in the pitcher’s hand

To derive an expression for the velocity and position of the ball as a function of time,
we begin by writing the acceleration as a function of time. Recall we assumed the only
force on the ball is due to gravity, which acts to accelerate the ball in the positive y-
direction. Thus, we have:
r
a = 0 x$ + gy$ m / sec 2

Acceleration is down at gravitational rate.

where g = gravitational acceleration at the Earth’s surface = 9.798 m/sec2

To obtain the instantaneous velocity at any time (t), we must integrate this equation
with respect to time. (Remember integrals from calculus? Basically, an integral is a
mathematical means of adding together lots of small changes over time to develop the
total change.) Thus,
r r
V (t ) = Vinitial + gty$ = 10 x$ + gty$ m / sec

This equation tells us that the ball will keep its initial horizontal velocity constant but
will speed up in the vertical direction due to gravity (which we already knew). To
obtain the instantaneous position of the ball at any time (t), we must once again
integrate this equation with respect to time so that:

r r r 1 1
R( t ) = Rinitial + Vinitial t + gt 2 y$ = 10tx$ + gt 2 y$
2 2

We can now use these relatively simple equations to propagate the motion of the
baseball. Using a simple spreadsheet, we can determine the position and velocity of
the baseball for each second for a total 10-second flight. These values and a graph
depicting the trajectory are shown in Figure 49. Notice the trajectory we derived is
exactly what we’d expect from experience. Often, people mistakenly refer to the

50
STK Astronautics Primer

baseball trajectories as “parabolic;” however, as we know, this is actually a small


section of an ellipse.

Time X (m) Y (m)


Trajectory of the Baseball
(sec) X (m)

0 0 0.0

100
10

20

30

40

50

60

70

80

90
0
1 10 4.9 0.0
50.0
2 20 19.6 100.0
3 30 44.1 150.0
4 40 78.4 200.0

Y (m)
250.0
5 50 122.5 300.0
6 60 176.4 350.0
7 70 240.1 400.0
450.0
8 80 313.5 500.0
9 90 396.8
10 100 489.9

Figure 49: Results of baseball propagator: Using the simple equations of motion we
derived for the baseball, we can use a spreadsheet to calculate the x and y positions at
each point in time for a 10-second flight. Plotting these on a graph, we see the shape of
the trajectory.

Note that the technique we developed here was for analytic propagation. An analytic
propagation technique has a close-form solution. In other words, given the initial
conditions, we solve directly for the position and velocity at any future time using a
straightforward “plug and chug” of the equations of motion. How accurate is this
propagation technique? As we shall see, this (and any other method) is only as good
as the assumptions we make. For example, we assumed no wind resistance, but we
know from experience that a sudden gust of wind could make this trajectory change
considerably. In the next section, we’ll apply this same basic technique to understand
the slightly more complicated motion of a satellite in orbit.

The Two-Body Propagator


Now we’ll develop a simple method we can use to propagate the position and velocity
of a satellite known at a given time to predict its position and velocity at some time in
the future. By “simple method” we mean the restrictions placed on the complexity of
the problem. In this case, one of the primary assumptions we will make is that there
are only two bodies concerned—the Earth and the satellite. Thus, we arrive at the
term used to describe this approach “the restricted two-body problem.”

51
STK Astronautics Primer

Motion Analysis Process

The approach we’ll take is exactly the same as the one we used to describe the motion
of the baseball in the previous section. Using the same motion analysis process
described, we can apply the three basic steps of our motion analysis model: define a
coordinate system, make assumptions and identify equations of motion. Our
coordinate system will be the geocentric-equatorial coordinate system described earlier.
The assumptions we make will “restrict” our solution to cases in which these
assumptions apply. Fortunately, this includes most of the situations we’ll encounter.
We’ll assume that:

♦ satellites travel high enough above the Earth’s atmosphere so that the drag
force is small.
♦ the satellite won’t maneuver or change its path, so we can ignore the thrust
force.
♦ we’re considering the motion of the satellite close to the Earth, so we can
ignore the gravitational attraction of the Sun, the Moon or any third body.
(That’s why we call this the two-body problem.)
♦ compared to Earth’s gravity, other forces such as those due to solar
radiation, electromagnetic fields, etc. are negligible.
♦ the mass of the Earth is much, much larger than the mass of the spacecraft.
♦ the Earth is spherically symmetrical with uniform density and can thus be
treated as a point mass.

After all these assumptions, we’re left with gravity as the only force affecting the
motion of a satellite for the restricted two-body problem. This can be expressed (using
vector notation) as:

v − GM Earth M satellite $
Fsatellite = 2
R
Rsatellite

The force due to gravity on a satellite depends on the mass of both the satellite and
Earth and the distance to the Earth’s center. The direction is down, in the minus-R
direction.

For convenience, we often combine GMEarth to derive an expression µEarth , known as the
Earth’s gravitational parameter.

While Newton’s law of gravity describes the force on the satellite, we can use Newton’s
2nd law of motion to describe the effect of that force to develop our equations of
motion. From Newton’s second law, the force on the satellite can be expressed as:

52
STK Astronautics Primer

Fsatellite = M satellite a satellite

Setting the two expressions for force on the satellite equal to each other, we develop
an expression for the satellite’s acceleration:

v −µ
a satellite = 2 Earth R$
Rsatellite

A satellite accelerates down [minus-R direction] due to gravity. The further away from
the Earth’s center, the smaller the gravitational force and, therefore, the smaller the
corresponding acceleration.

This equation says the motion of a satellite depends only on the distance between the
center of the Earth and the satellite. It is independent of satellite mass. Substituting
the more common notation for acceleration, we get the two-body Equation of Motion.

.. − m
v Earth R$
R= 2
R

where

R = distance from center of Earth (km) to satellite

&&v
R = 2nd derivative of position = acceleration (km/sec2)

µEarth = gravitational parameter (km3/sec2)

r
R$ = unit vector in direction of R

Two-Body Orbit Propagation

What can the two-body equation of motion tell us about the movement of a satellite
around the Earth? Unfortunately, in its present form—a second-order, non-linear,
vector differential equation—it doesn’t help us visualize anything about this
movement. So what good is it? To understand the significance of the two-body
equation of motion, we must first “solve” it using rather complex mathematical slight-
of-hand. When the smoke clears, we’re left with an expression for the position of an
object in space in terms of some variables we already know.

R=
(
a 1 − e2 )
1 + e cosν

53
STK Astronautics Primer

where
r
R = magnitude of R

a = semimajor axis (km)

e = eccentricity (dimensionless)

ν = true anomaly (deg or radians)

This equation represents the solution to the restricted two-body equation of motion
and describes the location, R, of a satellite in terms of a few constants and some
initial conditions. We can now make use of the solution to the two-body equation of
motion to propagate the position of a satellite to any point in time.

In nice circular orbits, determining how long a satellite takes to travel from an initial
position to a future position is simple, because the satellite is moving at a constant
speed. However, in an elliptical orbit this speed varies (recall a satellite travels fastest
at perigee and slowest at apogee, keeping total energy constant). As a result, we don’t
know how the true anomaly, ν, changes with time because it doesn’t change
uniformly. Here’s where Johannes Kepler came to the rescue. He developed this
technique to describe the orbit of Mars. To describe motion in an elliptical orbit,
Kepler began by defining the mean motion, n, which tells us the mean, or average,
speed in the orbit. The mean motion is defined as:

angle 2π µ
n= = =
time P a3

where

n = mean motion (rad/sec)

P = period (sec)

µ = gravitational parameter (km3/sec2)

a = semimajor axis (km)

Kepler figured out how to move n to a time in the future and, conversely, given a
future n, how to find out how long Mars would take to travel there. Kepler’s approach
was purely geometrical—he related motion on a circle to motion on an ellipse. To do
this, he had to invent a new angle called the mean anomaly, M, defined as:

M = nT (1)

where

M = mean anomaly (rad)

54
STK Astronautics Primer

n = mean motion (rad/s)

T = the time since last perigee passage (sec)

Mean anomaly is an angle that has no physical meaning and can’t be drawn in a
picture. We’ll have to describe it mathematically. Expressing this equation in terms of
two points in the same orbit:

Mfuture - Minitial = n(tfuture -tinitial) - 2kπ (2)

where

Mfuture = mean anomaly when the satellite is in the future position (rad)

Minitial = mean anomaly when the satellite is in the initial position (rad)

tfuture – tinitial = time of flight (TOF)

tfuture = time when the satellite is in the final position (e.g., 3:47 a.m.)

tinitial = time when the satellite is in the initial position (e.g., 3:30 a.m.)

k = the number of times the satellite passes perigee

To relate elliptical motion to circular motion, Kepler defined another new angle called
the eccentric anomaly, E, so that he could relate M to E and then E to n. With all of
these things defined, Kepler was able to develop his now-famous equation, commonly
called Kepler’s Equation. (For this equation to work, all angles must be in radians.)

M = E − e sin E (3)

where

E = eccentric anomaly (rad)

e = eccentricity

Kepler then related E to n using:

e + cos ν
cos E = (4)
1 + e cos E

where

ν = true anomaly (rad)

And related ν to E through:

55
STK Astronautics Primer

cos E − e
cos ν = (5)
1 − e cos E

Finally, we have all the equations needed to build a complete two-body propagator.
The first problem, and the easiest, is finding the time of flight between two points in
an orbit. Given νinitial and νfuture, we simply go through the following steps:

♦ Use Equation (4) to solve for Einitial and Efuture

♦ Use Equation (3) to solve for Minitial and Mfuture

♦ Use Equation (2) to solve for the time of flight (tfuture – tinitial)

Note that if n is between 0° and 180°, so are E and M.

The second problem we can solve using Kepler’s method is far more practical. This
involves determining a satellite’s position at some future time, tfuture, as shown in
Figure 50.

Time of Flight

υ Future
υ Initial

Figure 50: Time of flight on an elliptical orbit. The second problem Kepler tackled
was predicting the future position of a satellite knowing only its initial position.

This second problem is much trickier. We assume that we know where the satellite is
at time tinitial, so we know νinitial. We start by finding Einitial, using Equation (4). Then we
find Minitial using Kepler’s Equation (3).

M initial = E initial − e sin Einitial

Now, because we know tfuture, using Equation (2), we can find Mfuture:

56
STK Astronautics Primer

( )
M future − M initial = n t future − t initial − 2 kπ

Great. We’re now on our way to finding nfuture, which tells us where the satellite will
be. So we go to Kepler’s Equation again to find Efuture. Let’s rearrange this equation
and put E on the left side:

E future = M future + e sin E future (6)

OOPS! Efuture is on both sides of the equation. This is called a transcendental equation
and can’t be solved for Efuture directly. In fact, almost every notable mathematician
over the past 300 years has tried to find a direct solution to this form of Kepler’s
Equation without success. So we must resort to “math tricks” to solve for Efuture. The
“math trick” we’ll use is called iteration. To see how iteration works, think about the
kids’ game Twenty Questions. In this game, your partner thinks of a person, place, or
thing and you must guess what he’s thinking of. You’re allowed 20 questions
(guesses) to which your partner can answer only “yes” or “no.” In seeking the right
answer, a good player will systematically eliminate all other possibilities until only the
correct answer remains.

A mathematical application of this can be seen using another transcendental


equation:

y = cos(y)

Because we can’t solve for y using algebra (we can’t get the y out of the cosine
function to put all the ys on the left side), we must iterate. Begin by taking a guess at
the value for y, and take the cosine to see how close you were. Then take this as the
new value of y and use it for the next guess, and keep doing this iteration until the
new y equals the old y (or is, say, within 0.000001 radians of the old value).

Let’s try it to see what the answer for y really is. Take out your calculator and use π/4
radians as your first guess for y. (Remember to set your calculator to use radians, not
degrees.) Keep pressing the cosine function button and you’ll see the value slowly
converges to 0.739085 radians (about 43°). Presto—you’ve now solved the
transcendental equation y = cos(y) using iteration!

We can use this same iterative technique to solve Equation (6) for Efuture. It turns out
that the values for M and E are always pretty close together, even for the most
eccentric orbits, so let’s use Mfuture for our first guess at Efuture. Here’s the algorithm:

♦ Use Mfuture for the first E.

♦ Solve Equation (6) for a new Efuture.

♦ Use this new Efuture for the next guess for Equation (6).

57
STK Astronautics Primer

♦ Keep doing the previous step until Efuture doesn’t change by much (less than
about 0.0001 rad). At this point, the solution is said to have converged.

This brute force iteration method will solve Equation (6), but there are much better
methods in use in most standard propagators.

Let’s quickly summarize what we’ve learned. If we know where we are in an orbit and
where we want to be, we can use Kepler’s Equation to solve for the time it takes to
travel to the place we want to be. The solution is very straightforward. If, however, we
know where we are and want to know where we’ll be at some future time, we can use
Kepler’s Equation to find that location only by iterating a transcendental equation for
eccentric anomaly.

Orbit Perturbations
In deriving the two-body equation of motion, we had to assume that:

♦ gravity was the only force


♦ the Earth’s mass was much greater than the satellite’s mass
♦ the Earth was spherically symmetric with uniform density, so it could be
treated as a point mass

These assumptions led us to the restricted two-body equation of motion:

&&v + µ R
R $ =0 (7)
R2

The solution to this equation gives us the six classical orbital elements:

a = semimajor axis

e = eccentricity

i = inclination

Ω = longitude of the ascending node

ω = argument of perigee

ν = true anomaly

Under our assumptions, the first five of these elements remain constant for a given
orbit. Only the true anomaly, ν, varies with time as the satellite travels around its
fixed orbit. What happens if we now change some of our original assumptions? Other
classical orbital elements besides ν will begin to change as well. Any changes to these
classical orbital elements due to other forces are called perturbations. To see which
classical orbital elements will change and by how much, let’s look at our first
assumption—gravity is the only force.

58
STK Astronautics Primer

Atmospheric Drag

Just as a sudden gust of wind changes the course of a football, atmospheric drag can
affect satellites in low Earth orbit (below about 1000 km). Let’s look at how drag
affects the orbital elements.

Because drag is a nonconservative force, it takes energy away from the orbit in the
form of friction on the satellite. Thus, we expect the semimajor axis, a, to decrease.
The eccentricity also decreases, since the orbit becomes more circular. Let’s see why
this is so. When a satellite in an elliptical orbit is at perigee, it has a greater speed
than it would if the orbit were circular at that same altitude. The drag decreases the
speed, making it closer to the circular orbit speed. That’s exactly what we see in
Figure 51. It’s as if drag were giving the satellite a small negative velocity change, or
delta (∆) V, (slowing it down) each time it passes perigee.

successive orbits

∆V
drag

Earth’s
atmosphere original orbit

Figure 51: The effect of drag on an eccentric, low-Earth orbit. As a satellite passes
through the upper atmosphere at perigee, drag acts to gradually slow it down,
circularizing the orbit until it eventually decays.

Drag is very difficult to model because of the many factors affecting the Earth’s upper
atmosphere and the satellite’s attitude. The Earth’s day-night cycle, seasonal tilt,
variable solar distance, the fluctuation of Earth’s magnetic field, the Sun’s 27-day
rotation and the 11-year cycle for Sun spots make precise modeling nearly
impossible. The force of drag also depends on the satellite’s coefficient of drag and
frontal area, which can also vary widely, further complicating the modeling problem.

The uncertainty in these variables is the main reason Skylab decayed and burned up
in the atmosphere several years earlier than first predicted. For a given orbit,
however, we can approximate how the semimajor axis and the eccentricity change

59
STK Astronautics Primer

with time, at least for the short term. Different propagation techniques use different
methods of estimating drag, with widely varying accuracy.

Earth’s Oblateness—“J2”

Columbus was wrong! The Earth isn’t really round. From space, it looks like a big,
blue spherical marble, but if you take a closer look, it’s really kind of squashed. Thus,
it can’t most accurately be treated as a point mass, as it is treated in the two-body
assumption. We call this squashed shape oblateness. What exactly does an oblate
Earth look like? Imagine spinning a ball of jello around its axis and you can visualize
how the middle (or equator) of the spinning jello would bulge out—the Earth is fatter
at the equator than at the poles. This bulge can be modeled by complex mathematics
(which we won’t do here) and is frequently referred to as the J2 effect. J2 is a constant
describing the size of the bulge in the mathematical formulas used to model the oblate
Earth. Why “J2?” This term arises from the mathematical short-hand used to
describe Earth’s gravitational field. (Gravitational acceleration at any point on Earth is
commonly expressed as a geopotential function expressed in terms of Legendre
polynomials and dimensionless coefficients Jn—whew!). J2, J3 and J4 are the zonal
coefficients that depend on latitude. Of these, J2 is by far the most important; it is
roughly 1000 times greater than either J3 or J4. However, for more precise modeling
of the Earth’s oblateness, all three of these must be taken into account. In addition,
other, higher order terms can be included in the model. These terms serve to slice the
Earth into wedges that depend on longitude (sectoral terms) and slice it again into
regions of longitude and latitude (tesseral terms).

Let’s concentrate on the simplest and most profound case, J2. What effect does J2
have on the orbit? Let’s look at Figure 52. Here it’s shown exaggerated; actually the
bulge is only about 22 km thick. That is, the Earth’s radius is about 22 km longer
along the equator than through the poles.

60
STK Astronautics Primer

F
J2

F
J2

Figure 52: Diagram of Earth oblateness. The Earth’s oblateness, shown here as a
bulge at the equator (highly exaggerated to demonstrate the concept) causes a twisting
force on satellite orbits that change various orbital elements over time.

Let’s see if we can reason out how this bulge will affect the orbital elements. The force
caused by the equatorial bulge is still gravity. Recall that gravity is a conservative
force; therefore, the total mechanical energy in an orbit must be conserved. Total
mechanical energy depends on the orbit’s semimajor axis. Thus, as long as energy
remains constant (i.e., no drag or other forces adding or stealing energy), the
semimajor axis also remains constant. It turns out that the eccentricity, e, also
doesn’t change, although the explanation for this is beyond the scope of our
discussion here. Although you might expect the inclination to change because the
bulge pulls on our orbit, it doesn’t! However, it does affect the orbit by changing the
right ascension of the ascending node, Ω, and moving the argument of perigee, ω,
within the plane. That’s not very intuitive, but it’s like a force acting on a spinning
top. If you stand a nonspinning top on its point, gravity causes it to fall over. If you
spin the top first, gravity still tries to make it fall but, because of its angular
momentum, it begins to swivel—this motion is called precession. Let’s examine the
effect of precession on the ascending node and the argument of perigee more closely.

How J2 Affects the Right Ascension of the Ascending Node, Ω

The gravitational effect of this equatorial bulge slightly perturbs the satellite
because the force no longer originates from the center of the Earth. This causes
the plane of the orbit to precess (like the spinning top), resulting in a movement of
the ascending node, ∆Ω. This motion is westward for posigrade orbits (inclination
<90°) and eastward for retrograde orbits (inclination > 90°).

61
STK Astronautics Primer

Figure 54 shows this nodal regression rate, Ω& , as a function of inclination and
orbital altitude. Let’s look more closely at this figure. It shows that the higher the
satellite is, the less effect the bulge has on the orbit. This makes sense because
gravity decreases with the inverse square of the distance (see Newton’s Law of
Gravitation). It also says that if the satellite is in a polar orbit (center of the graph),
the bulge has no effect. The greatest effect occurs at low altitudes with low
inclinations. This makes sense, too, because the satellite travels much closer to
the bulge during its orbit, and thus is pulled more by the bulge. For low-altitude
and low-inclination orbits, the ascending node can move as much as 9° per day
(lower left corner and upper right corner of Figure 53).

Nodal Regression Rate as a Function of


Inclination and Eccentricity

10.0
Nodal Regression Rate (deg/day)

4000km x 100km
5.0 altitude elliptical orbit

0.0

-5.0 100km altitude circular


orbit
-10.0
0 20 40 60 80 100 120 140 160 180
Inclination (deg)

Figure 53 Nodal Regression Rate, Ω & . The nodal regression rate caused by the Earth’s
equatorial bulge. Positive numbers represent eastward movement; negative numbers
represent westward movement. The less inclined an orbit is to the equator, the greater
the effect of the bulge. The higher the orbit, the smaller the effect.

How J2 Affects the Argument of Perigee, ω

Figure 54 shows how perigee location rotates for an orbit with a perigee altitude of
100 km depending on the inclination for various apogee altitudes. This perigee
rotation rate, ω& , is difficult to explain physically, but it could be derived
mathematically from the equation for J2 effects on perigee location. With this
perturbation, the major axis, or line of nodes, rotates in the direction of satellite
motion if the inclination is less than 63.4° or greater than 116.6°. It rotates
opposite to satellite motion for inclinations between 63.4° and 116.6°.

62
STK Astronautics Primer

Perigee Rotation Rate as a Function of


Inclination and Eccentricity

20.0
Perigee Rotation Rate (deg/day)

100km altitude
15.0 circular orbit

10.0

5.0 4000km x 100km


altitude elliptical orbit
0.0

-5.0
0 20 40 60 80 100 120 140 160 180
Inclination (deg)

Figure 54: Affects of J2 on argument of perigee. The perigee rotation rate caused by
the Earth’s equatorial bulge depends on inclination and altitude at apogee.

Sun-Synchronous and Molniya Orbits

The effects of the Earth’s oblateness on the node and perigee positions give rise to
two unique orbits that have very practical applications. The first of these, the Sun-
synchronous orbit, takes advantage of eastward nodal regression at inclinations
greater than 90°. Looking at Figure 55, we see that the ascending node moves
eastward about 1° per day at an inclination of about 98° (depending on the
satellite’s altitude).

Coincidentally, the Earth also moves around the Sun about 1° per day (360° in
365 days), so at this Sun-synchronous inclination, the satellite’s orbital plane will
always maintain the same orientation to the Sun. This means the satellite can see
the same Sun angle when it passes over a particular point on the Earth’s surface.
As a result, the Sun shadows cast by features on the Earth’s surface won’t change
when pictures are taken days or even weeks apart. This is important for remote-
sensing missions such as reconnaissance, weather and monitoring of the Earth’s
resources, because they use shadows to measure an object’s height. By
maintaining the same Sun angle day after day, observers can better track changes
in weather, terrain or man-made features.

63
STK Astronautics Primer

Sun line
Sun angle orbit plane

Sun
angle Sun line
Sun line

Sun angle

Sun angle
Sun line

orbit plane
rotates at ~1 deg/
Earth moves around the
day due to Earth’s
Sun at ~1 deg/day
oblateness

Figure 55: Sun-Synchronous Orbit. Sun-synchronous orbits take advantage of the rate
of change in right ascension of the ascending node caused by the Earth’s oblateness. By
carefully selecting the proper inclination and altitude, we can match the rotation of Ω with
the movement of the Earth around the Sun. In this way, the same angle between the
orbit plane and the Sun can be maintained without using rocket engines to change orbit.
Such orbits are very useful for remote sensing missions that want to maintain the same
Sun angle on targets on the Earth’s surface.

The second unique orbit is the Molniya orbit, named after the Russian word for
lightning (as in “quick-as-lightning”). This is a 12-hour orbit with high eccentricity
(about e = 0.7) and a perigee location in the Southern Hemisphere. The inclination
is 63.4°—why? Because at this inclination, the perigee doesn’t rotate so the
satellite “hangs” over the Northern Hemisphere for nearly 11 hours of its 12-hour
period before it whips “quick as lightning” through perigee in the Southern
Hemisphere. Figure 56 shows the orbit and ground tracks for a Molniya orbit. The
Russians used this orbit for their communications satellites because they didn’t
have launch vehicles large enough to put them into geosynchronous orbits from
their far northern launch sites. Molniya orbits also offer better coverage of
latitudes above 80° north.

64
STK Astronautics Primer

Figure 56: Molniya orbit and ground tracks. Molniya orbits take advantage of the fact
that ω, due to Earth’s oblateness, is zero at an inclination of 63.4°. Thus, apogee stays
over the Northern Hemisphere, covering high latitudes for 11 hours of the 12-hour orbit
period.

Other Perturbations

Other perturbing forces can affect a satellite’s orbit and its orientation within that
orbit. These forces are usually much smaller than the J2 (oblate Earth) and drag
forces but, depending on the required accuracy, satellite planners may need to
anticipate their effects. These forces include:

♦ Solar radiation pressure, which can cause long-term orbit perturbations and
unwanted satellite rotation.
♦ Third-body gravitational effects (Moon, Sun, planets, etc.), which can perturb
orbits at high altitudes and on interplanetary trajectories.
♦ Unexpected thrusting caused by either out-gassing or malfunctioning
thrusters, which can perturb the orbit and cause satellite rotation.

Dealing with Perturbations


Understanding and modeling orbit perturbations is one of the primary activities of
astrodynamics. Even very early space pioneers such as Kepler and Newton spent
considerable effort grappling with the various forces that disturb a satellite from pure
two-body motion. Let’s begin by classifying perturbations with respect to their relative
effects on orbital elements. Perturbations can cause both secular and periodic
changes to orbital elements. Secular perturbations are those that cause elements to
steadily diverge over time. Periodic perturbations are those that impart a sinusoidal
variation in elements over time. Short-term periodic perturbations are those with a
period less than the orbit period. Long-term periodic perturbations are those with a

65
STK Astronautics Primer

period greater than one orbit period. We can now look at two techniques for modeling
both secular and periodic perturbations. The relative effects of these different types
are illustrated in Figure 57.

Long term effects on orbital elements for various


types of perturbations
9
8 Long-term
Secular
Orbital element variation

7 Periodic
(arbitrary units)

6
5
4
3
2 Short- term
1
Periodic
0
0

8
Orbit Periods

Figure 57: Types of orbit perturbations. Orbit perturbations are categorized based on
their long-term effects on orbital elements.

General Perturbations Techniques

General perturbations techniques are those that generalize the effects on orbital
elements in order to develop analytic expressions allowing for direct computation. In
the grossest sense, general perturbation techniques apply “fudge factors” to the
simple two-body solution to account for the effect of different perturbation sources.
For example, returning to our baseball-throwing astronaut scenario, we could model
the drag on the baseball using:

1
D= ρV 2 A
2

where

D = Drag (N)

ρ = air density (kg/m3)

V = baseball velocity (m/sec)

66
STK Astronautics Primer

A = baseball cross-sectional area (m2)

We could substitute this expression into our baseball equations to derive new
equations of motion that would account for the general effects of drag. Even with this
additional complexity, the equations could still be solved analytically.

One of the most widely used propagators was developed by the North American
Aerospace Defense Command (NORAD) to track the 8000-plus satellites and space
junk in orbit around the Earth. Called Merged Simplified General Perturbations-4,
MSGP-4, this technique uses the generalized approach to model orbit perturbations.

Special Perturbations Techniques

In contrast, special perturbation techniques are based on special case assumptions


about the orbit scenario that allow for more detailed modeling of individual
perturbation sources.

Two-line Element Sets


One of the most commonly used methods of communicating orbital parameters is the
2-line element sets generated by NORAD in Cheyenne Mountain, Colorado (literally, in
Cheyenne Mountain!). It is important to note that TLEs were developed specifically for
use with the MSGP-4 propagator! Using TLEs with any other propagator may
invalidate some of the built-in assumptions.

These elements contain many of the same elements as the classical orbital elements,
along with some additional parameters for identification purposes and for use in
modeling perturbations in the MSGP-4 propagator.

STK Propagators
In selecting the “best” propagator to use for a given application, it is important to
consider the assumptions on which they are based. The temptation is to use the most
“accurate” propagation model available. However, this can lead to false accuracy,
especially for very long term propagation over which time even the best models can
break down. The relative accuracy among various propagators can vary widely
depending on the scenario. For example, a geostationary spacecraft is well above most
atmospheric drag and J2 perturbations. Therefore, a short-term difference between
the two-body propagator and a more complex technique could be relatively small.
However, for a spacecraft in low-Earth orbit, the short-term differences between the
two solutions could be significant. The objective is to choose the most appropriate
propagation scheme for a given application. Unfortunately, any propagation technique
is simply an attempt to model events in the real world. Regardless of the technique
chosen, only frequent tracking of an orbit can guarantee that the predicted orbital
parameters will match the real world.

67
STK Astronautics Primer

Two-Body

The two-body propagator or Keplerian motion propagator uses the same basic
technique outlined in the two-body equation of motion development. This technique
assumes the Earth is a perfect sphere and the only force acting on a satellite is
gravity. This propagator doesn’t account for any perturbations.

J2

The J2 propagator accounts for the 1st order effects of J2 Earth oblateness. This effect
causes secular changes to the orbital elements over time.

J4

The J4 propagator accounts for 1st and 2nd order J2 effects as well as 1st order J4
effects. J3, which causes long-term periodic effects, is not modeled. Because the 2nd
order J2 and 1st order J4 effects are very small, you’ll see very little differences
between the J2 and J4 propagators for most orbits considered.

MSGP-4

MSGP-4 stands for Merged Simplified General Perturbations-4. It is one of the most
widely used propagators in the industry. This technique uses the generalized
approach to model orbit perturbations, including both secular and periodic variations
such as Earth oblateness, solar and lunar gravitational effects and drag.

It is important to understand the purpose for which MSGP-4 was developed. NORAD
wanted a simple propagator that would provide acceptable results for a wide variety of
tracking tasks, from tracking high-priority military satellites to keeping tabs on space
junk. Given the over 8000 objects NORAD must track, a technique was needed that
would not be computationally intensive (is was first developed back when computers
were much slower than today). Furthermore, there are far fewer ground tracking sites
than there are objects to track. Thus, it is important that the propagated solution be
good enough to ensure the tracking radar can find a specific object the next time it
gets around to tracking it (which, for some very low priority objects like pieces of
rocket boosters, may be days or even weeks).

Because MSGP-4 is a generalized approach, it is specifically tailored to a given set of


inputs: the TLE sets that contain parameters that make the analytic calculations
valid. For best results, MSGP-4 should always be used with TLEs. Likewise, NORAD-
generated TLEs should only be used in the MSGP-4 propagator.

68
STK Astronautics Primer

HPOP

HPOP is the High Precision Orbit Propagator. As its name implies, it uses a powerful
propagation technique to incorporate sophisticated orbit perturbation models. HPOP
uses a variety of high-fidelity models including:

♦ Joint Gravity Model (JGM) 2—a highly precise model of the Earth’s
oblateness.
♦ Lunar/solar gravitational effects—based on U.S. Naval Observatory data.
Accurate to within 0.03 arc seconds.
♦ Atmospheric drag effects— using either the 1971 Jacchia or the Harris
Priester model, which takes into account daily variations in the height of the
atmosphere due to solar heating among other parameters.
♦ Solar radiation pressure—yes, sunlight produces a small force on any exposed
surface. This force varies depending on how reflective the surface is—a
mirrored surface is more reflective than a black surface.

Depending on the application, HPOP can deliver accuracy on the order of 10 meters
per orbit. But beware of false accuracy, always remember—“garbage in, garbage out.”
To get this level of accuracy, your initial orbital elements must be at least this
accurate to start with. Putting NORAD-generated TLEs into HPOP will not necessarily
give you a better solution. No propagator can create accuracy, at best it can only
minimize the long term dispersions due to inherent limitations in our ability to model
the effects of perturbations.

Great Arc

The Great Arc propagator allows the user the model the flight path of a vehicle flying
close to Earth. By providing way points and speed, STK uses Great Arc to predict
where and when it will be next. The propagation scheme is essentially the same as the
Two-Body propagator, no perturbations are assumed.

Ballistic

The Ballistic propagator is a variation of the Two-Body propagator for use with
ballistic trajectories. These are the trajectories used by artillery shells, suborbital
sounding rockets and ballistic missiles, allowing the user to predict impact points or
determined required velocity to reach a certain point. The propagation “engine” is the
same as the Two-Body propagator, no perturbations are modeled.

LOP

The Long-term Orbit Predictor (LOP) allows accurate prediction of a satellite’s orbit
over many months or years. This is often used for long duration mission design, fuel
budget definition, and end-of-life studies. For performance reasons, it is impractical

69
STK Astronautics Primer

to compute the long-term variation in a satellite’s orbit using high accuracy, small
time step, propagators that compute a satellite’s position as it moves through its
orbit. LOP exploits a “variation of parameters” approach which integrates analytically
derived equations of motion computing the average effects of perturbations over an
orbit. This approach allows large multi-orbit time steps and typically improves
computational speed by several hundred times while still offering high fidelity
computation of orbit parameters.

Lifetime

Lifetime estimates the amount of time a low Earth orbiting satellite can be expected to
remain in orbit before the drag of the atmosphere causes reentry. While the
computational algorithms are similar to those implemented in the Long-term Orbit
Predictor, there are some important differences. First, a much more accurate
atmospheric model is implemented to compute the drag effects. The gravitational
model for the Earth, however, is significantly simplified since the inclusion of the
higher order terms doesn’t impact orbit decay estimates.

70
STK Astronautics Primer

SATELLITE ACCESS

Overview

Line of sight
♦ Why is line of sight important for satellite viewing?

Communication Architecture
♦ What are the elements that make up a space mission communication
architecture?

Communication Links
♦ What are the communication paths used by satellites and ground
stations?

Understanding Access
♦ What is meant by “satellite access?”

Describing Access
♦ How do I explain and quantify satellite access?

Line of Sight
Standing on a beach, looking out over the ocean on a clear day, you can see
right to the edge of the horizon, which is about 8 miles away. If you were to
watch a ship sailing away from you, you would notice that the hull would
disappear first, followed by the top of the mast The taller the mast, the further
the ship could be from the shore before disappearing completely from sight. An
object is in your line of sight if you can draw a straight line between yourself
and the object without any interference, such as a mountain or a bend in a
road. An object beyond the horizon is below our line of sight and, therefore, can
be difficult to communicate with. Early methods of long-distance
communications increased the effective line of sight by employing methods
such as smoke signals or other means. Because the line of sight was raised so
that others could see, or receive the message being sent, communication
among objects that didn’t really have a direct line of sight was achieved.

At the beginning of the 20th Century, radio engineers discovered that certain
frequencies could be bounced off the ionosphere, greatly extending the effective
line of sight for communications and creating a new “radio horizon” far beyond
the more limited visual horizon. Today’s communications satellites take this
basic principle to the extreme. Ground-based operators can “bounce” radio

70
STK Astronautics Primer

communications off satellites stationed in geosynchronous orbit, creating a


virtual line of sight extending half a world away. Furthermore, by bouncing
signals between satellites, this virtual line of sight can be extended to cover the
entire global community.

Satellite access is the problem of determining when, where and for how long a
satellite (or any number of objects you may be interested in) is within line of
sight of other objects.

Communications Architecture
To understand the satellite access problem more clearly, let’s begin by
reviewing the players in the access problem. Figure 58 illustrates the elements
that make up a communications architecture.

crosslink
return link

re
crosslink forward link

tu
rn
do

lin
wn

k
lin
k fo
ink

rw
ar
upl

d
lin
k

Figure 58 Communications Architecture. The communications architecture


consists of space and ground-based elements tied together by communications
paths or links.

The communications architecture has four elements:

Spacecraft

The spaceborne elements of the system.

71
STK Astronautics Primer

Ground stations

The Earth-based antennas and receivers that talk to the spacecraft. These are
typically remote tracking sites or users of mission data.

Control Center

The command center that controls the spacecraft and all other elements of the
system.

Relay satellites

Additional satellites that link the primary spacecraft with the ground stations
and control center.

Communications Links
Information moves among the elements of the communications architecture
using various communication paths or links.

Uplink
Data sent from a ground-based station to the primary satellite.

Downlink
Data sent from the primary satellite to a ground station.

Forward link
Data sent from a ground station to the primary satellite via a relay satellite.

Return link
Data sent from the primary satellite to a ground station via a relay satellite.

Crosslink
Data sent through either the forward or return link between the primary satellite
and a relay satellite.

Understanding Access
The simplest example of a satellite access problem is that between a satellite
TV dish and a direct-broadcast geosynchronous satellite. As a user, you just
want to point your dish and start watching the big game. Thus, you’re only
interested in downlink.

72
STK Astronautics Primer

Figure 59 Downlinking. As a direct TV subscriber, you would only be interested


in downlinking.

For satellites in geostationary orbit, the geometry and dynamic nature of both
uplink and downlink is very stable. It is this stability that make geostationary
satellites so useful for point-to-point message relaying. You can set up you
dish, point it at a pre-determined point in the sky and pretty much forget about
it. Because the ground track of a geosynchronous orbit is at most a tiny figure-
8 centered on the equator, a line of sight between the dish and the satellite is
almost constant (at least constant enough to ensure uninterrupted broadcast of
that big game!).

Describing Access
So how do you know where to point the satellite dish? This depends on two
important pieces of information:

♦ Your location (latitude and longitude)


♦ The satellite’s location (orbital elements)

From this information, STK can determine the azimuth and elevation settings
for your dish. These two important parameters are defined below.

♦ Azimuth - The compass direction between the ground site and the
satellite direction, e.g., due south would be 180°.
♦ Elevation - Angle measured from the local horizontal to the satellite
direction, e.g., directly overhead would be 90°.

In addition, STK also computes range—the distance between the dish and the
satellite. While range is not so important to the average satellite TV user, it
becomes very important for communications engineers who must ensure there
is sufficient transmission power to effectively carry the signal across this
distance.

These three parameters are fairly constant for a geosynchronous satellite;


however, for any other satellite, we must include an additional parameter—

73
STK Astronautics Primer

time. Imagine it is exactly 13:20 (1:20 p.m.) local time and you want to point a
radar antenna at an airplane flying over your position. The plane is initially
due south of you, flying north but out of sight below the horizon. As the plane
first comes into view, azimuth will be 180° and the elevation 0°. As it
continues to fly north, azimuth will stay constant (disregarding Earth rotation)
and the elevation angle will increase. As it flies overhead, the azimuth angle
will switch around to 0° (it is now north of your position) and the elevation
angle will gradually decrease until the plane once again drops from view below
the horizon. If we kept track of the azimuth and elevation viewing angles to the
plane at 10-minute intervals we could build a simple table, or access report, as
shown below.

Time Azimuth Elevation Comment


(deg) (deg)

13:20 180 -45 Below horizon (south of you)


13:30 180 0 Just coming into view
13:40 180 45 Well above horizon
13:50 ---- 90 Directly overhead (azimuth is
undefined)
14:00 0 45 Azimuth has switched around, plane is
now north of your position.
14:10 0 0 Plane drops below horizon, out of view

The geometry with respect to a plane flying directly overhead is relatively easy
to visualize. However, if you’re faced with the problem of a satellite in a highly
eccentric orbit flying over a position northwest of you on a descending node,
things become much more complicated. Fortunately, STK works out the
geometry for you. Figure 60 shows a relatively simple satellite access geometry.

74
STK Astronautics Primer

Figure 60 Satellite Access. Satellite access refers to the problem of


determining the geometry and timing of line-of-sight between various ground and
space-based objects.

Access reports can be easily generated using STK; the report provides azimuth,
elevation and range (AER) data for specified time intervals between whatever
objects you choose to define. Cumulative access time or duration can also be
reported.

Access information becomes even more complicated when multiple objects


must be taken into account. For example, if you are relaying information
between various ground stations and relay satellites or using an entire
constellation of satellites, the geometry can become very complex. To handle
these tasks, STK allows you to link together various objects to create a “chain”
for which access information can be determined. Figure 61 shows a more
complex series of “chained” objects. If one of these “objects” is the Sun, orbit
lighting—a critical parameter for power and thermal management—can be
computed.

Figure 61 Chained objects. A series of ground and space-based objects can


be “chained” together, allowing STK to compute access information between all
of them. In this picture, access among a facility, relay satellite and a second
facility is shown.

RECOMMENDED READING

For a more detailed explanation of the topics in this primer as well as an


introduction to the space environment, spacecraft design, rockets and systems,
we recommend:

75
STK Astronautics Primer

Understanding Space: An Introduction to Astronautics, Sellers, 1994,


McGraw-Hill.

To purchase a copy of Understanding Space, please contact McGraw-Hill at


www.mcgraw-hill.com, www.mhhe.com, or 800-338-3987 : The McGraw-Hill
Companies, >Order Services, PO Bos 545, Blacklick, Ohio. ISBN: 0-07-
057027-5.

Understanding Space is part of the Space Technology Series, a cooperative


activity of the United States Department of Defense and National Aeronautics
and Space Administration. Series editor is Dr. Wiley J. Larson. Other books in
the series include:

Fundamentals of Astrodynamics and Applications, Vallado, 1997, McGraw-


Hill.

Space Mission Analysis and Design, 2nd edition, Larson & Wertz (ed.), 1996,
Kluwer and Microcosm.

Space Propulsion Analysis and Design, Humble & Larson, 1995, McGraw-
Hill.

Reducing Space Mission Cost, Larson & Wertz (ed), 1996. Kluwer and
Microcosm.

Cost-Effective Space Mission Operations, Boden and Larson, 1996, McGraw-


Hill.

Spacecraft Structures and Mechanisms: From Concept to Launch, Sarafin and


Larson, 1995, Kluwer and Microcosm.

For more information about these books, please contact Kluwer at


www.wkap.nl, McGraw-Hill at www.mcgraw-hill.com or 800-338-3987, or
Microcosm at www.microcosm.com.

Future books in the series:

Modeling and Simulation: In Integrated Approach to Development an


Operations, Cloud and Rainey.

Human Space Mission Analysis and Design, Connally, Giffen and Larson.

Other recommended reading:

Fundamentals of Astronautics, Bate, Mueller & White, 1971, Dover.

1997 Microcosm Directory of Space Technology Data Sources, 1997, Wertz


and Dawson, Kluwer and Microcosm.

76

You might also like