You are on page 1of 56

A Little Book on

Gravity, Calculus and Planetary Motion

Richard R. Auelmann

2006
Preface

I start with the premise that youngsters with well above average aptitude and an
inquisitive mind should be exposed to applied mathematics, including the calculus and
differential equations, at an early age. How early, I don’t know. Not that they should be expected
to master these subjects, but they should acquire an appreciation and “feel” for the role applied
mathematics plays in explaining the physical world. This requires an ability to follow the
mathematical arguments and proofs, which is a lesser standard than being able to furnish the
proofs on their own. It’s somewhat analogous to being able to read a novel, rather than write one.
But even to read you must know the language, and calculus is the language of physics. How best
to motivate a youngster to study this material is the question.

My approach is to focus on the particular of problems that spurred the development of


applied mathematics in the first place: namely, the motions of the Solar system: that is, the orbits
of the Moon about the Earth and the planets about the Sun. Isaac Newton developed the calculus
primarily to explain these motions and in the process formulated his three laws of motion, which
form the basis for the branch of physics known as classical mechanics (or dynamics). This is an
example of the adage: “Necessity breeds invention”.

There were two reasons for this particular marriage (between applied mathematics and
dynamics). The first reason was that planetary motions were known through observation with
relatively great precision by the time Newton arrived on the scene. Copernicus had established
that the planets revolved around the Sun, and Kepler had established that they moved in elliptical
orbits with their speeds dependent on their distances from the Sun. So the motions were known,
but the forces governing them were unknown. The second reason was that planetary motions
were unaffected by the atmosphere, winds, friction, magnetism and other Earthly interactions. So
Newton was dealing with well-defined “clean” problems where the motions are affected only by
the force of gravity. To solve these problems three elements had to come together: (1) an
understanding of gravity at the fundamental level; (2) the laws of motion expressed as
differential equations relating the time rate of change of momentum to force, and (3) the
mathematical means (the calculus) to solve these differential equations. I assume you have no
notion of what is meant by a differential equation, much less how to solve one. Indeed Newton
had to invent the calculus in order to explain the mechanism of the planetary motions as well as
the motion of the Moon and artificial moons (satellites) about the Earth. But you will learn, as
we follow Newton and his predecessors in their quest to understand the motions of the Solar
system, that calculus and differential equations are not so mysterious after all.

There are many excellent books covering the development of dynamics and the law of
gravitation. Those aimed at the general public tend to be heavy on historical details, while being
sanitized in the sense that the key mathematical developments and proofs are omitted, or skirted
over. At the other end are the textbooks and treatises on planetary motion and dynamics, which
assume a high level of mathematical competence. This book is offered as a compromise between
these two treatments. The focus is on a single class of problems: those concerned with the
gravitational attraction of a sphere and a point mass particle --- the simplest problem in planetary
motion.

2
Chapters on mathematics are interspersed with those on gravity. A minimalist approach is
taken in that only the mathematics needed to follow the other chapters is included. Proofs are
presented with the intermediate steps included. You may choose to follow the proofs, or simply
skip to the final results. Numerous diagrams are included to aid in understanding. Your job is to
follow the reasoning and mathematical developments as they are laid out. There are no problems
to work. The journey gets progressively more difficult, but also more rewarding in terms of
insight into the interaction between applied mathematics and dynamics. It is the kind of book I
would have liked to have available in my formative years, and was spurred by the notion that my
two grandchildren might benefit.

3
Contents

1. Introduction

2. A Little Algebra, Trigonometry and Geometry

3. The Solar System


Model of Copernicus
Kepler’s Model
Kepler’s First Law
Kepler’s Second Law
Kepler’s Third Law

4. Galileo and the Birth of Dynamics


The Pendulum
Bodies in Free Fall
Inclined Plane Experiment
Rates of Change
Two-Dimensional Motion of Projectiles
Motion of the Planets
Huygens’ Experiments with Circular Motion
Descartes’ Theory of Vortices

5. Introduction to the Calculus


Differential Calculus
Integral Calculus

6. Newtonian Revolution
Newton’s Thought Experiment
Laws of Motion
Motion of the Moon About the Earth
Projectile Problem Revisited
The Two-Body Problem
The Kepler Problem
Kepler’s Second Law Revisited
Kepler’s First Law Revivited
Energy Integral and Velocity
Explicit Solution for Time

7. Gravitational Attraction of a Sphere


Dumbbell Problem
A Proof

8. The Gravitational Constant and the Mass of the Earth


Cavendish Experiment
Mass of the Earth

4
1. Introduction

What is gravity? You say it is the force that makes things fall. Correct. You may say it’s
the force that keeps the Earth in orbit about the Sun, and the Moon in orbit about the Earth.
Again correct. Is it the force of attraction between any two bodies? You may not so sure since
you haven’t seen two balls at rest on a flat table roll towards one another. If true, it must be a
very weak force. But if so weak, how does it dictate the motion of the Earth about a Sun almost
150 million kilometers away?

Just what parameters determine the force of gravity? The answer provided by Isaac
Newton is the Universal Law of Gravitation: Every two particles attract one another with a force
which is proportional to the product of their masses and inversely proportional to the square of
the distance between them, acting along the line joining the two particles. In mathematical terms,
the force of gravity is given by
Gm1m2
F=
r2

where m1 and m2 are the masses of the two particles, r is the distance between them, and G is
the universal constant of gravitation. Mass is a measure of the amount and density of material

within a particle. Suppose, like Newton, we were interested in determining the motion of the
Moon about the Earth. In this case we would€let m1 and m2 represent the masses of the Earth and
€ Moon € respectively, and let r be the distance between them. These are forces of attraction along
the line between the two bodies. And the magnitude of the force decreases as the square of the
distance between the two bodies. The force F2 on the Moon is towards the Earth and is equal in
magnitude and opposite in direction to€ € F on the Earth as depicted in Figure 1-1.
the force 1

m2 !
F2!

F1!
r!
m1 !

Figure 1-1. Mutual Gravitational Attraction of Two Masses ( F1 = −F2 )

It wasn’t until Newton in the latter half of the seventeenth century that the Law of
Gravitation and its ramifications were fully understood. Almost a century earlier, Galileo
correctly deduced that gravity caused bodies to fall to the Earth, €
but couldn’t reconcile the fact
that the Moon maintains a near fixed distance from the Earth, or the fact that the Earth maintains
a near fixed distance from the Sun, all because of gravity.

Knowing the Law of Gravitation is only a necessary first step. One must also explain how
a body responds to the force of gravity. Johannes Kepler used the astronomical measurements of
the Danish astronomer Tycho Brach to establish that the planets move in elliptical orbits about
the Sun. He knew how they moved, but he couldn’t explain why. This would take Newton’s

5
formulation of the Laws of Motion. These Laws of Motion are expressed in terms of time rates
of change of position and velocity. Time rates of change involve the mathematical concept of
limits, called derivatives, something I assume you know nothing about. Don’t fret. Prior to
Newton no one else knew what derivatives. They are what “differential calculus” is all about.
Because the equations for Laws of Motion involve derivatives they are called differential
equations. In fact most of the equations of physics are formulated as differential equations, so
they are worth learning about. Given the equation for the force of gravity and the Laws of
Motion, formulation of the differential equations of motion is generally straight forward, once
you learn about derivatives. The difficult part comes in solving these differential equations. This
involves a process called integration, requiring knowledge of integral calculus. Integral calculus
can be thought of as the inverse of a differential calculus. So this is what is in front of you.

Newton made no attempt to explain why it was in the nature of bodies to be attracted to
one another. In this sense his Law of Gravitation was no less mystical than Aristotle’s
explanation that it was the nature of bodies to fall. However, Newton was able to show that the
law of gravitation, when coupled with his three laws of motion, was in agreement with the
astronomical observations of his day. It may come as a surprise that Newton accomplished this
without knowing the masses of the Sun, Earth or Moon, or the value of the gravitational constant
G. The story of how the law of gravitation was established is as remarkable as any in the laurels
of science and mathematics. It was not an easy journey. Indeed it took Newton some twenty
years to fully reconcile his understanding of gravity and its affect on the motion of bodies.

Newton does not deserve sole credit for these developments as he himself admitted. We
have already mentioned Kepler. Galileo Galillei measured the acceleration due to gravity on
Earth, and came close to fully enunciating the first two laws of motion. Christian Huygens’
experiments with circular motion played a key role at this juncture. However, Newton was the
first to realize and prove, using his newly developed calculus, that the inverse square law of
gravitation applied not only to bodies on Earth, but governed the motion of the planets about the
Sun, and the motion of other planets about other Suns. This was a monumental breakthrough.
Working independently, Leibniz also developed the calculus and it is his notation and
vocabulary, which have prevailed. Nevertheless, Newton stands as the scientific giant of the era.
Like Albert Einstein of the twentieth century, he did his most important work before he reached
thirty. So if you have aspirations to walk in the footsteps of a giant, you better get cracking.

6
2. A Little Algebra, Trigonometry and Geometry

In one sense learning mathematics is a lot like learning a foreign language. There are
rules, terminology and special notations, so it helps if you don’t always have to fall back on a
dictionary or rulebook to follow what is being conveyed. This chapter covers the mathematics
and notation you will need to follow the material in Chapters 3 and 4. Refer back here as needed.

Notation

In technical writing it helps to know the common unwritten rules of mathematical


notation, particularly as they apply to physics. Plus (+) and minus (-) signs are used, but
multiplication ( × ) and division ( ÷ ) signs are not. Multiplication is implied by any number of
ways including the following:
cx, c ⋅ x, c(x), (c)(x)
€ €
The form c(x) can be a source of confusion because it can also mean that c is a function of x.
Herein I provide a clue to the meaning, if it is not obvious from the context. Division is denoted
a €
as or a b . A superscript denotes that a quantity is raised to a power ( a 2 = a ⋅ a and a1/ 2 = a
b

where is the square root sign). Usually letters in the front of the alphabet a, b, c, etc. denote
constants, while x, y, z and t denote variables. Herein x, y and z represent distances, and t
€denotes time. Greek letters also denote variables, often angles€or ratios. €

€ Fundamental Rules of Algebra

a+b=b+ a
a + b + c = (a + b) + c = a + (b + c)
ab = ba
abc = (ab)c = a(bc)

A Few Algebraic Relationships


€ a(x + y) = ax + ay
(a + b)( a − b) = a 2 − b 2
2
(a + b) = a 2 + 2ab + b 2
2
(a − b) = a 2 − 2ab + b 2
(x + a)(x + b) = x 2 + ( a + b) x + ab

7
Exponents
a0 = 1
a m a n = a m +n
am
n
= a m a−n = a m−n
a
m n
(a ) = a mn
n
( a) = an / 2
m
a m b m = ( ab)

Pythagorean Theorem
€ right triangles; that is, to triangles where two of the three sides
This theorem applies to
are mutually perpendicular as are sides a and b in Figure 2-1. The third side c is called the
hypotenuse. The Pythagorean theorem states that for any right triangle the square of the
hypotenuse is equal to the sum of the squares of sides a and b.

c 2 = a2 + b2

c!
b!

a!

Figure 2-1. Right Triangle

For example, if side a is 4 and side b is 3 then

c 2 = 16 + 9 = 25
c = 25 = 5

Area of a Triangle.

To follow some of our proofs, you will need to know that the area of a triangle is equal to
one half the product of its base times its height. Any side of the triangle can be taken as the base.
The corresponding height is then the perpendicular distance from the base to the maximum
distance from the base. Note that the three triangles ABC, ABC’ and ABC” depicted in Figure 2
have the same area. That is because C, C’ and C” are the same distance from their common base
line AB. In Chapter 4, we will show Newton’s clever proof for Kepler’s Second Law using this
relationship among equal area triangles.

8
C! C ! C !

A! B!

Figure 2-2. Equal Area Triangles

Circle and Trigonometric Functions

The distance from the center of a circle to any point on the circle is the radius r. The
circumference is the distance around the circle given by 2π ⋅ r where π = 3.14159 ⋅ ⋅ ⋅ . The area
within a circle is π ⋅ r 2 .

Consider the relationship between rectangular


€ (x, y) and€polar (r, θ) coordinates of point
P in Figure 2-3. When the radial coordinate r is a constant, the P traces a circle as a function of

the angle θ. We also introduce the trigonometric functions sine of θ (abbreviated as sin θ) and
cosine of θ (abbreviated as cos θ) and plot their values as a function of θ in units of either
degrees or radians. The relationship between these units is that there are 360 degrees (also
written 360°) in 2π radians. So the number of degrees per radian is 360/(2π )= 57.296… The sine
curve is shifted 90° or π/2 radians from the cosine curve (and vice versa). Also the amplitudes of
these functions are bounded by ±1.
x y
,
y! r r

P! 1!
x sin θ cosθ

r €
y €
θ €

x! 0! θ (deg)
O!
€ €

-1!

Definitions
0 90° 180°           270°           360°!
x
= cosθ x = r cosθ
r 0 π/2 π             3π/2             2π!
y y = r sin θ
= sin θ
r
Figure 2-3. Sinusoidal Functions

€ 9
Observe that
" x %2 " y %2 x 2 + y 2
$ ' +$ ' =
#r& #r& r2

The Pythagorean theorem provides x 2 + y 2 = r 2, so it follows that



sin 2 θ + cos 2 θ = 1

The tangent of θ (abbreviated as tan θ) is defined as
€ y sin θ
tan θ = =
x cosθ

It is introduced because it denotes the slope of a function, which is a crucial element of


differential calculus. Most technical hand-calculators application have sin, cos and tan functions.

Cartesian Coordinates

The French philosopher and mathematician Rene Descartes introduced Cartesian


coordinates in 1619. As a philosopher he was famous for his saying “dubito ergo gogito, cogito
ergo sum” (I doubt therefore I think, I think therefore I am). As the story goes while
daydreaming he watched a fly flit around the room and realized that at any instant he could
locate the fly using three numbers representing the distances from two of the walls and the
ceiling. This marked the birth of Cartesian coordinates. Customarily, they are labeled x, y and z.
When the motion is restricted to a plane, one customarily uses x and y. The location of P on the
circle in Figure 3 illustrated the use of Cartesian coordinates in a plane. More complex examples
follow in subsequent chapters.

With Cartesian coordinates, geometry became amendable to algebraic analysis. Their


introduction marked the birth of analytic geometry, by which a mathematical curve or
geometrical figure could be represented by an equation.

Vectors

A vector denotes both magnitude and direction. Figure 2-4 depicts a Cartesian coordinate
frame where the vectors r1 and r2 locate two points P1 and P2 with respect to the origin O of the
Cartesian frame xyz. Also shown is the vector r21 = r2 − r1 . Vectors are commonly depicted in
bold face type. The components of these two vectors can be represented as follows
€ €
rx1 rx 2 rx 2 − rx1

r1 = ry1 , r2 = ry 2 , r21 = ry 2 − ry1
rz1 rz 2 rz 2 − rz1

€ 10
The representation of the vector components within a vertical column of parallel lines is an
example of a “matrix” representation. To evaluate vectors first resolve them into a Cartesian
frame, and then treat them algebraically.

z! z!

rz

r
€ r2
y! y!
€ ry r21 = r2 − r1
rx r1

x! x!
€ €
Figure 2-4. Vectors and Cartesian Coordinates
€ €

11
3. The Solar System

The known planets in the 16th century were Mercury, Venus, Earth, Mars, Jupiter and
Saturn as depicted in Figure 3-1. The values of the orbit radius a (taken relative to orbit radius of
the Earth), the period P or time to complete one revolution (relative to the period of the Earth),
and the eccentricity e for these planets are given in Table 3.1.

Figure 3-1. Known Solar System (circa 16th Century) Represented as Circular Orbits in a Plane

TABLE 3.1. Modern Orbit Values

a P e
Mercury 0.3831 0.2408 0.2056
Venus 0.7233 0.6152 0.0068
Earth 1 1 0.0167
Mars 1.5237 1.8808 0.0934
Jupiter 5.2028 11.862 0.0482
Saturn 9.5388 29.457 0.0551

Eccentricity is a measure of how far the orbit deviates from a circle, with a value of zero
being a circle. Except for Mercury and Mars, the orbits are nearly circular. Precise definitions of
a and e will be given later. The representation of the Solar system with the planets rotating about
the Sun was introduced by Nicolaus Copernicus (1473-1543) and is known as the Copernican
system, in contrast to the Ptolemy’s system where the motions of the Sun and the other planets
were depicted relative to a stationary Earth.

12
Model of Copernicus

The model of Copernicus was more complex than depicted in Figure 3-1, because it
accounted for the fact that the orbits are not exactly circular. Copernicus accomplished this by
introducing two simultaneous circular motions as shown in Figure 2. This approach of
representing planetary motion as combinations of circular motion dates back to the Greeks, who
had no notion as to why the planets moved as they did, but believed circles were the model of
simplicity.

Referring to Figure 3-2, Point P is the position of the planet, while Point O is the position
of the Sun. Point A on the large circle with radius a rotates about its center C at the rate n
(expressed as either radians or degrees per unit time). If expressed in radians per unit time, then
2π/n corresponds to the period (the time required for A to complete one revolution). Point P,
located on the small circle with radius ae /2 , rotates about A at twice the rate that A rotates about

C. So P completes two revolutions about A in the time it takes A to complete one revolution
about C. Point O is a distance 3ae /2 from C. Thus e determines both the relative size of the two
circles but also locates the position of the Sun relative to the center of the large circle. For e <<1,

the path traced by P approaches a circle of radius a displaced a distance ae from C towards O,
and the period of the orbit is 2π/n. So besides describing the shape of the orbit and its position

relative to the Sun, the model also describes the location of the planet as a function of time.

y!

P!

A! 2n!
a

x!
€ C! O!

ae /2
3ae /2


€ n!

Figure 3-2. Copernican Model

Let xy be a nonrotating coordinate frame (relative to the star background) with origin at
the Sun O. The x-axis is aligned in the direction of the point of closest approach (the perihelion),
a distance a(1− e) from O. The point of maximum distance from O (the aphelion) lies along the
negative x-axis a distance a(1− e) form O. So the distance between the perihelion and aphelion


13

is 2a . Also the orbit is symmetric about the x-axis, but not about the y-axis. The x and y
coordinates of P are expressed as functions of time by

# e 3e &
€ x = a% cos nt + cos2nt − (
$ 2 2'
# e &
y = a% sin nt + sin2nt (
$ 2 '

It is customary in astronomy to denote the product of the mean orbit rate n and time t, by the
angle M = nt termed the mean anomaly. These equations provide a first approximation for
planetary motion at least€for small values of eccentricity.

Kepler’s Model

The Danish astronomer Tycho Brahe (1546-1601) and the German astronomer Johannes
Kepler (1571-1630) ushered in the next developments in astronomy. Tycho, as he was known,
was an experimentalist and Kepler was a theoretician. From his observatory at Uraniborg,
Denmark Tycho Brache established himself as the first truly great European observational
astronomer since the Greeks of antiquity.

It came to pass that Kepler (Figure 3-3) secured a job as Tycho’s assistant in 1601, with
the task of compiling tables of planetary positions. Kepler was obsessed with finding a model
that was in full agreement with this data. Much to Kepler’s chagrin, Tycho doled out his data
piecemeal on a more or less need-to-know basis. However, upon Tycho’s death later that year,
Kepler came into full possession of the treasure trove of Tycho’s data. The Copernican model
was good agreement with data for all the planets except Mercury and Mars, which differences
exceeded the accuracy of Tycho’s observations. In Kepler’s mind these errors could not be
ignored. By 1905, Kepler had established his first two laws of planetary motion, which were first
published in his 1609 book Astronomia Nova (New Astronomy). But it wasn’t until 1618 that he
established his third law of planetary motion, which he published in Harmonics Mundi
(Harmony of the Worlds) in 1619.

Figure 3-3. Johannes Kepler

14
After Brahe’s death, Emporer Rudolf II commissioned Kepler to produce tables, which
would allow one to determine the past, present and future positions of the planets based on
Brahe’s observations. This work was published as the Rudolphine Tables in 1627. It required the
introduction of Kepler’s equation, which was quite separate from his three laws of motion.
Kepler’s equation, together with his three laws of planetary motion, were published in a series of
seven books between 1618 and 1621 under the title Epitoma astronomiae Copernicanae
(Epitome of Copernican Astronomy).

Kepler’s First Law: The orbit of a planet about the Sun is an ellipse with the Sun at one
focus. This law describes the shape of the orbit about the Sun. You may ask: “What is an
ellipse?” Lay a piece of paper on a board, and place two pins in the paper as shown in Figure 3-4.
Now tie a piece of string between the two pins, leaving some slack. Now using the string as a
restraining guide take a pencil and trace a path on the paper without letting the string go slack.
You have just drawn an ellipse.

Figure 3-4. Drawing an Ellipse

Referring to Figure 3-5, the ellipse has two foci, at O and O", where the pins were
located. The long axis of the ellipse is in the OO"-direction. The maximum span of the ellipse is
2a. It is equal to the length of the string r + r" . The half distance a is called the semi-major axis.
The distance between the two foci is 2ae, where e defines the eccentricity of the orbit. For an

ellipse 0 < e < 1. For e = 0, the two foci merge into one point, and the ellipse becomes a circle.

Point C half way between O and O" is the center of the ellipse. The Sun is located at one of the
€ "
foci (we choose O) rather than O .
y!
€ P!
€ L!

a 1− e 2 r ! r!
a(1− e 2 )
ν
x!
€ Pmax! O ! C O! Pmin!

ae €
ae

€ €
a a

Figure 3-5. Parameters of an Ellipse


€ €
15
Point P on the ellipse can be located in polar coordinates ( r, υ ) using the equation

a(1− e 2 )
r=
1+ ecosυ€

The angle υ , measured from Pmin (the point of closest approach to O), is called the true anomaly.
The numerator a(1− e 2 ) is the distance from O to L and is called the semi-latus rectum.

€ The method Kepler used to establish this relationship was circuitous. Referring to Figure
3-6, he circumscribed a circle with origin at O and diameter equal to 2a. He then constructed a
line B€P " passing through P and perpendicular to the major axis Pmin Pmax of the orbit. Then he
introduced the angle χ = Pmin CP " (called the eccentric anomaly). By tedious trial and error,
fitting Tycho Brache’s observational data for the position of Mars, Kepler found that the data
€ satisfied the relationship


a(1− e 2 )
r = a(1− ecos χ ) =
1+ ecosυ

which has come to be known as Kepler’s equation. Kepler also recognized that this is one form
of the equation for an ellipse.

P !

P!

r!

ν
χ

Pmax! Pmin!
C B! O!
€ €
ae

€ a

Figure 3-6. Kepler’s Construction.



Now from Figure 3-6, we see that

CB CPmin − OPmin − OB a − a(1− e) − r cosυ ae − r cos υ


cos χ = = = =
CP' CP' a a

16

Substituting cos χ into Kepler’s equation, we obtain the previous expression for an ellipse in
terms of the polar coordinates r and υ . In other words

a(1− e 2 )
€ r = a(1− ecos χ ) =
€ 1+ ecosυ

Through the rather involved use of trigonometric identities, the above relationship can be
manipulated to provide a more convenient relationship between the true anomaly and eccentric
anomaly €
υ 1+ e χ
tan = ⋅ tan
2 1− e 2

Figure 3-7 compares the Copernican and Keplerian predictions for the planet Mercury,
which has an eccentricity of 0.2056. To make this comparison, we convert form polar
€ coordinates using the relations
coordinates r and υ to Cartesian

x = r cosυ , y = r sin υ

Note that the values of the x and y axes are normalized on the semi major axis a, which is the
same for both models. Observe that the Copernican model is more nearly circular than the
Keplerian model. €

Figure 3-7. Orbits for Mercury Predicted by the Copernican and Keplerian Models

17
Kepler’s Second Law: The radius vector from the Sun sweeps out equal areas in equal
periods of time. Figure 3-8 depicts this law where the shaded regions are equal in area so the
time to traverse from A to B is the same as the time to traverse from C to D. From this it is
evident that the planet moves fastest when is at the point of closest approach to the Sun (the
perihelion) and moves slowest at the point farthest approach to the Sun (the apehelion). Thus
while the first law describes the shape of the orbit, the second law provides an implicit statement
concerning the speed around the orbit.
B!
A!
C!
D!

O!

Figure 3-8. Depiction of Kepler’s Second Law

It is ironic that Kepler deduced this law by making two offsetting errors. Rather than
dwell on Kepler’s errors, we present Newton’s correct geometrical proof, which applies to any
central force (not just the inverse square law of gravity). Referring to Figure 3-9, Point O denotes
the center of force and A denotes the initial position of a body moving with velocity v. In the
absence of force the body will arrive at B in a given unit of time τ , and at C " in a second unit of
time. Suppose now that at the halfway point B an instantaneous force was applied such that, in
the absence of any previous motion, it would have moved to b in a unit of time. But with the
prior motion it will move along the diagonal of the parallelogram B b C C " to C, as a
€ €
consequence of Newton’s second law of motion (presented later). In the absence of force it
would then move to D" in the next unit of time. Suppose instead a second instantaneous force is
applied toward O such that in the absence of previous motion, it would have moved to c in a unit

of time. Then, as before, it will move along the diagonal of a parallelogram arriving at D in a unit
of time.
€ D"
D!

€ C! C"
c!


B!
b!

O!
A!
Figure 3-9. Newton’s Geometrical Proof of Kepler’s Second Law

18
This process can continue indefinitely. Then since the area of a triangle is equal to its
base times its height, we have that the areas OAB = OBC " because they have equal base lengths
AB and BC ". The areas OBC " and OBC are equal because they share the common base OB.
Thus the areas OAB = OBC. This process can be repeated with the results that areas OAB = OBC
= OCD each produced in a unit of time. I’ve gone through the trouble of presenting this labored

proof, only to emphasize the elegance and simplicity of an analytical proof of the same law (to
€ € €
be presented in Chapter 6) using the calculus.

Kepler’s Third Law: The squares of the periods of the planets are proportional to the
cubes of their semi-major axes. The period is the time to complete one revolution. So in
comparing the periods for planets 1 and 2, Kepler’s Third Law provides

P12 a13
=
P22 a23

Note that this is in agreement with the values listed in Table 1. As we see later, the third law
provides vital evidence concerning the functional dependence of gravity upon the distance
between any two bodies. Unlike the € first two laws, which deal with a single planetary orbit, the
third law establishes a relationship between different planetary orbits.

Remarkably, this law is independent of the orbit eccentricity. Hence, the orbits in Figure
3-10, having the same semi-major axis, also have the same period.

Figure 3-10. Two Orbits with the Same Period

Kepler’s laws represented the high water mark in the development of kinematics (how
things move) as it applied to astronomy. And they served as the springboard for the development
of dynamics, the branch of physics developed by Isaac Newton, that makes it possible to predict
how things move for a given set of forces and initial conditions. This step between kinematics
and dynamics is immense.

Kepler’s Equation. To produce the Rudolphine Tables, Kepler had to be able to locate a
planet as a function of time t elapsed from its position of closest approach Pmin to the Sun. Let n
represent the mean orbit rate over a complete orbit revolution, related to the orbit period P by the
equation n = 2π /P . The angle M = nt is called the mean anomaly. Kepler’s equation

M = χ − esin χ €

€ €
19

relates M to the previously defined eccentric anomaly χ used to establish Kepler’s first law. As
such it provides the required relationship between time and the eccentric anomaly χ . Figure 3-11
is a graphical representation of this law. The angle M satisfies the condition that the two shaded
areas are equal. Equating the two areas and dividing through by a 2 /2 we obtain Kepler’s

equation.

P ! P !
a2 a €
χ − ae sin χ
2 2
a2
M
2

P! P!

χ
M

C! B! O! Pmin! C! B! O! Pmin!
€ €

Figure 3-11. Kepler’s Equation Requires That the Shaded Areas be Equal.

Recall that the model of Copernicus, allows one to locate the position of a planet as an
explicit function of the t or M:

# e 3e &
x = a% cos M + cos2M − (
$ 2 2'
# e &
y = a% sin M + sin2M (
$ 2 '

Unfortunately, for the Kepler model neither the eccentric anomaly χ nor the true anomaly υ can
be expressed as an explicit function of true anomaly M. Rather one must use either iteration or a
series expansion. €

€ €
The iteration process is as follows:

1. Iterate M = χ − esin χ to obtain χ for a specified time t.

2. Compute r = a(1− ecos χ ) for a given χ


€ €
υ 1+ e χ
3. Compute υ from tan = ⋅ tan
€ 2 1−
€e 2

20


Admittedly this process is tedious, which is why astronomers developed series expansions for r
and υ in powers of e. The following series converge very rapidly for small values of e and are
convergent for e < 0.6627:

€ e2 e3
r /a = 1− ecos M − (cos2M −1) − ( 3cos 3M − 3cos M ) − ⋅ ⋅ ⋅
2 8
2 3
5e e
υ = M + 2esin M + sin2M + (13sin 3M − 3sin M ) + ⋅ ⋅ ⋅
4 112

Here M = nt is the mean anomaly. To convert to Cartesian coordinates simply compute


€ x = r cosυ , y = r sin υ

In Figure 3-7 we compared the shapes of the planet Mercury orbit predicted by the
Copernican and Keplerian models. But since there was no time correlation it was not possible to
compare points at the same€instant in time, except at the crossings of the x-axis. But now
knowing the relationship between position and time through Kepler’s equation, it is possible to
make such a time comparison between the two models. Such is provided in Figure 3-12 where
the normalized Cartesian position differences ( Δx /a and Δy /a ) between the two models are
plotted as a function of M = nt .

0.05
€ €

x/a error
0.04
y/a error
0.03

0.02
x/a, y/a Errors

0.01

0.00
0 90 180 270 360
-0.01

-0.02

-0.03

-0.04
eccentricity, e = 0.2056
-0.05
Mean Anomaly, M (deg)

Figure 3-12. Time Referenced Errors in the Copernican Model for Mercury

21
Retrospective. Our discussion commenced with Copernicus and his Heliocentric model
of the Solar system. It can be argued from a purely geometrical viewpoint, that the Heliocentric
model is no more correct than Ptolmey’s Geocentric model, only less complex. On the other
hand, it is hard to image that Kepler could have developed his planetary laws of motion had he
viewed Brache’s observations from a Geo-centric perspective.

Kepler’s accomplishments were remarkable from another standpoint. He offered no


explanation as to what force caused the planets to move as they do. There was no connection to
the force of gravity and the central role it played.

22
4. Galileo and the Birth of Dynamics

While the design and construction of static structures (buildings and bridges) to resist the
force of gravity was reasonably well understood in the sixteenth century, the dome of the
Basilica in Florence being a prime example, theories concerning bodies in motion and force were
mostly wrong. The viewpoints of the Greek Philosopher Aristotle (384 BC-322BC) concerning
motion can be summarized as follows. Stones fall because it is in their nature to do so.
Descending motion is natural while ascending motion is unnatural, and will eventually change to
natural motion as exemplified by the fact that projectiles thrown upward eventually come down.
Heavy objects fall faster than light objects. Force must be continuously applied to keep an object
moving at a constant speed, and can only be applied with direct contact. Planets move in
circular motion, because a circle is the purest form. These opinions are of interest only because,
as remarkable as it may seem today, they held sway for two thousand years until the so-called
Newtonian Revolution in the 17th century.

Such misunderstanding persisted in part because of the inability to isolate the effects of
gravity, friction and the surrounding atmosphere. So it should not be surprising that the birth of
dynamics was intertwined with the motion of the planets under the gravitational influence of the
Sun, and falling bodies on the surface of the Earth. In the first case, friction and atmosphere
effects play no role. In the second case, their effect is often secondary.

Galileo Galillei (1564 - 1642), an Italian (Figure 4-1), is probably best known for his
observations concerning the Sun and the planets using telescopes that he himself crafted. He was
also an outspoken advocate of the Copernican system and his resulting controversies with the
Catholic Church led to his house arrest. While Galileo is best known regarding the Solar system,
his greatest contributions concerned the motion of bodies on the surface of the Earth and his
findings regarding gravity. In fact, it was during his period of confinement that he conducted the
key experiments that led to his understanding of gravity and the first two laws of motion. It was
for these findings, that Galileo is often credited with having ushered in the birth of dynamics.

Figure 4-1. Galileo

23
The Pendulum. In his youth Galileo had been intrigued with the idea of using a
pendulum as a clock. Years later his interest returned to the pendulum as tool to understand the
nature of gravity. The pendulum consists of a mass (or “bob”) attached to an arm, which pivots
about the opposite end (Figure 4-2). The arm itself can be a string. The bob tends to hang straight
down under the force of gravity. Displaced to one side, it will oscillate back and forth about the
vertical. The path it traces is an arc of a circle with pivot point as the center of the arc. Over time
the amplitude of the oscillation will die down due to the retarding effect of the air.

θ" L!

m!

Figure 4-2. Pendulum

The motion of the pendulum is remarkable in many respects. Suppose you measure the
time to complete one oscillation (the bob moving from one extreme to the other and back) using
a stopwatch. This time interval is called the period. Provided the initial displacement-angle θ of
the bob were not too large (say no more than 15°), you would find that period does not
measurably change as the amplitude dies down.

Double the weight of the bob without changing the length of the string, and again
measure the period. Surprisingly, you will find the period to be unchanged. Thus you would have
confirmed that the period of the pendulum is independent of the mass m of the bob. What is the
implication of this? If you displace the pendulum to one side and hold it, you would sense that
the restraining force depends on the mass of the pendulum. Wouldn’t this mean that the force
trying to move the pendulum must also depend on the pendulum mass? So while the motion of
the pendulum is independent of its mass, the force trying to move it appears to depend on the
mass. Within this statement lies a clue concerning the First Law of Motion, which we will come
to later.

Let’s get down to specifics. If the length L of the string were one meter, you would find
that the period P of the pendulum to be almost exactly 2 seconds. If you doubled the length of
the string from one meter to two meters, you would find the period to be the √2 times longer or
2.828 seconds. And if the length of the string were again doubled to four meters, you would find
the period to be 4 seconds. If you made additional measurements, and graphed the results you
would be able to fit a smooth curve through the measurement points as depicted in Figure 4-3.

24
3.0

2.5

2.0

Period, P (seconds)
1.5

1.0

0.5

0.0
0 1 2 3 4 5 6 7 8

Length, L (meters)

Figure 4-3. Pendulum Period as a Function of Pendulum Length

In the process you would have established the relationship: P(seconds) = 2 L(meters) .
This is the equation for the curve in Figure 4-3. We say this is an empirical finding because it is
based on observation and experiment rather than on theory. In other words, it tells you how the
pendulum responds, but does not explain the underlying theory behind this response.

Bodies in Free Fall. Conventional wisdom prior to Galileo was that heavy bodies fall
faster than light bodies. Galileo’s study of the simple pendulum led him to an understanding of
the relationship between the force of gravity and motion. He reasoned that a pendulum is no
more than a body in free fall restrained by the pendulum arm, and the fact that the period is
independent of mass, implied that the motion of a body in free fall is also independent of its
mass. Galileo supposedly demonstrated this to surprised on-lookers by simultaneously dropping
light and a heavy objects from a tall building (possibly the Leaning Tower of Pisa) and showing
they hit the ground at virtually the same time.

Inclined Plane Experiment. To better understand the relationship between force and
motion, Galileo needed to establish the relationship between the distance traveled and time from
release for bodies in free fall. Because Galileo could not accurately measure this relationship
(falling objects move too fast for the measurement tools available) he slowed the process by
rolling balls down an inclined plane. He reasoned that the time scale would lengthen in direct
proportion to the slope of the inclined plane. Galileo found that the total distance traveled after 2
seconds was 4 times that traveled in the first second. After 3 seconds the total distance traveled
was 9 times that traveled in the first second. After 4 seconds the total distance traveled was 16
times that traveled in the first second, and so on. Indeed the total distance s was proportional to

25
the square of the total elapsed time t. This relationship is written s = c ⋅ t 2 where c is a constant
dependent on the slope. The force of gravity is somehow wrapped up in the constant c.

By reducing the slope of the inclined plane to zero, and giving the ball a small shove, he

undoubtedly found that the ball rolled a very long distance before measurably slowing. With a
flat plane, the force of gravity on the ball was zero. The only things slowing the ball were friction
between the ball and the surface, and drag due to the air. In their absence, Galileo reasoned that
the ball would roll at a constant speed indefinitely. This led to his formulation of the first law of
motion: In the absence of force, a body at rest tends to stay at rest, and a body in motion, tends
to stay in motion at a uniform rate. This was the first break from Aristotle’s position that force is
required to sustain motion.

Now return to the inclined plane experiment. We do not know the magnitude of the slope
Galileo used, but we could repeat Galileo’s experiment and record the slope. Then by adjusting
the time scale to account for the slope of the inclined plane, we can then determine the motion of
a body in free fall, and so determine the force of gravity. So lets get on with it. I propose we use
an inclined plane that is 1-cm drop in 100 cm (Figure 4-4). In essence we are reducing the force
of gravity (compared to free fall) by a factor of 100.

H = 1 cm!
Slope not to scale!

0 5 10 20 50 100!

L = 100 cm = 1 meter!

Figure 4-4. Inclined Plane Experiment.

Use a bubble level to be certain that the surface upon which we place the inclined plane is
horizontal. Release a small ball and measure the time at discrete distances (say at 5, 10, 20, 50
and 100 cm) marked along the way. Make multiple time measurements at each distance and take
the average to reduce the measurement error. If performed correctly, you would have obtained
the following measurements.

Distance, s (cm) time, t (s)


0 0.00
5 1.01
10 1.43
20 2.02
50 3.19
100 4.52

Note the following ratios using data from this table:

26
2 2
10 "1.43 % 20 " 2.02 %
=$ ' = 2, =$ ' =2
5 # 1.01& 10 # 1.43 &
2 2
50 " 3.19 % 100 " 4.52 %
=$ ' = 10, =$ ' = 10
5 # 1.01 & 10 # 1.43 &

This indicates that distance is proportional to the square of the time. Indeed, the data satisfies the
equation s(cm) = c ⋅ t 2 (sec) where c is a constant equal to 4.905 cm/s2. See Figure 4-5.

100

80
Distance, s (cm)

60

40

20

0
0 1 2 3 4 5
Time, t (seconds)

Figure 4-5. Measured Response for a 1:100 Slope

Increase the inclination of the plane (raise the height H, while keeping L the same), and
you will find that the time scale of the curve is compressed. The distance traveled will still vary
as the square of time, but the constant c will be larger. And in the limit when the plane is vertical
(so the ball is in free fall), the value of c would be 490.5 cm/s2. The reason is that when the slope
of the plane was 1 cm to 100 cm, the effect of gravity was reduced by 100 to 1. With the plane
removed, the full effect of gravity acts on the ball, so we are able to conclude that the distance
traveled by a ball in free fall is s(cm) = 490.5 ⋅ t 2 (sec) . Since there are 100 cm in 1 meter, we can
more conveniently express the above relationship as s(m) = 4.905 ⋅ t 2 (sec) . This relationship is
the same as in Figure 4-5 except that the vertical scale is in meters, rather than in centimeters. To
determine just how the force of gravity is wrapped up in the constant c, we must introduce the

concept of “rate of change”.

Rates of Change. Velocity is the time rate of change of distance, so it is slope of the
distance versus time curve. For the case of the ball rolling down the inclined plane, is obvious
that the slope of curve changes with time. You could obtain a crude estimate of the slope at any

27
instant in time by the following procedure. Draw a tangent to the curve at the time of interest
(say at t = 3 seconds as in Figure 4-6). The ratio of the slope height to the slope width is the
velocity.

Figure 4-6. Time Rate of Change

Knowing the equation of the distance-time curve s = c ⋅ t 2 , one can easily estimate the
slope at any time t using the following procedure. Referring to the above figure, compute the
difference Δs between the values of s at t + Δt and the value of s at t using the values of s
defined by the equation of the curve. We obtain

2
€ Δs = c ( t + Δt ) − ct 2
= c ( t 2 + 2tΔt + Δt 2 ) − ct 2
= c (2t + Δt )Δt

The ratio
Δs
€ = c (2t + Δt )
Δt

provides an estimate of the velocity at time t. The accuracy of this estimate improves as the value
of Δt decreases. In the limit, when Δt approaches zero, we obtain

28
# Δs &
Limit% ( = 2ct
$ Δt ' Δt →0

This limit is the equation for the velocity v of a body in free fall. So the velocity of a body in free
fall can be written
€ v = g ⋅ t, g = 2c

We have just established that the velocity of a body in free fall increases linearly with time. So if
velocity were plotted against time it would be a straight line with a slope equal to the constant a.
In other words, g is the rate of€change of velocity, which is called acceleration. Moreover, since
gravity is the force acting on a body in free fall, we have just determined that, for objects at or
near the surface of the Earth, the acceleration due to gravity g is 9.81 m/sec2. Rewriting the
distance-time equation in terms of g instead of c, we obtain

1
s= g ⋅ t 2, g = 9.81 m /sec 2
2

We have done a lot more than establish the motion of a body in free fall, for in those
steps computing
€ # Δs &
Limit% (
$ Δt ' Δt →0

we have introduced the underlying concept of differential calculus, as invented by Sir Isaac
Newton. We will return to this in Chapter 5.

Two-Dimensional Motion of Projectiles. Suppose instead of simply dropping a mass
from a height h, we threw it horizontally. The mass will follow an arc, which lies in the plane
defined by the vertical and the direction of initial velocity (Figure 4-7). Let x denote the distance
traveled in the horizontal direction and y denote the height. Both x and y are measured from the
point of release. Galileo presumably did just this, and measured the position as a function of
time. What he observed was that the horizontal velocity of the object was constant, which he
deduced from the fact that the horizontal distance traveled in each successive fixed interval of
time was the virtually the same. Any small discrepancies could be attributed to the retarding
influence of the air. He also found that the vertical distance traveled at different times from
release were the same as when he simply dropped the mass. So now Galileo knew that the
horizontal and vertical motions were uncoupled and were described by the two equations:

x = vx t
1
y = h − gt 2
2

where v x was the horizontal velocity imparted to the mass, and g was the previously determined
acceleration due to gravity (9.81 m/s2), and h is the initial height. If we use the first equation to
eliminate time t from the second€equation, we find

29
g 2
y=h− x
2v x2
g
where the quantity is a constant. This is the equation for a parabola, which has the general
2v x2
form y = A + Bx 2 where A and B €are constants. In the illustration below, the path of the projectile
is shown for three initial velocities and is launched from a height of 100 m.

€ 100

80
Vertical Distance, y (m)

60

40

20

0
0 10 20 30 40 50
Horizontal Distance, x (m)

Figure 4-7. Projectiles Launched Horizontally from 100 m

In all three cases the projectile hits the ground in the same elapsed time

2h
t1 = = 4.515 s
g
after release.

Galileo also established that a body set in motion, would remain in uniform velocity if
€ direction. In the process, he provided a key element of Newton’s
not acted upon by forces in that
First Law of Motion. It falls short of Newton’s definition of the First Law of Motion only in that
it makes no statement regarding the mass of the object.

Motion of the Planets. The planets revolve in near circular orbits about the Sun. What is
the nature of force that maintains these motions? It was understood from Galileo’s experiments
that in the absence of force a body tends to move in a straight line at a uniform velocity. And
Galileo correctly understood the effect of gravity on motion at the surface of the Earth. In spite
of his fascination with the Solar system, he apparently did not make the leap of faith that gravity

30
also dictated the motion of the planets. The notion of “force at a distance” was foreign, yet that is
precisely what is needed to explain planetary motion.

Huygens’ Experiments with Circular Motion. For the next step in understanding the
relationship between force and motion we turn to the Dutch physicist Christian Huygens (1629-
1695). Among Huygens’ many experiments those concerning circular motion are germane to our
quest for understanding the nature of gravity. Huygens swung objects of different mass m at the
ends of strings of different lengths r and measured the string force required to keep them moving
in circular motion. He found that the required force F is given by

mv 2
F=
r

Since the velocity v is a constant around the circle, it can be computed by dividing the
circumference of the circle by the period P (the time to complete one revolution). Thus

2πr
v=
P

So Huygens force relationship can also be written as



4 π 2 mr
F=
P2

The radial force is proportional to the mass at the end of the string and the length of the string,
and inversely proportional to the square of the time to complete a revolution. But like Galileo,
Huygens did not grasp the concept € of “force at a distance” so he did not attribute the restraining
force acting on the planets moving in their near circular orbits as being due to gravity. Rather he
adhered to the theory of vortices as advance by Descartes.

Descartes’ Theory of Vortices. The Frenchman Rene Descartes (1596-1656) was a


mathematician of the first rank, credited with having developed analytic geometry. His method
of finding tangents to a curve undoubtedly played a role in Newton’s later development of
differential calculus. However, our immediate interest in Descartes is the publication in 1664 of
his Principes de la Philosophie where he laid out his vortex theory concerning planetary motion.
He theorized that the universe is filled with a kind of ether (he used a different term), and that the
Sun, by virtue of its rotation, induced vortices within the ether that maintained the planets in their
orbits. This appears as an extension of the Aristotle’s viewpoint that a vacuum does not exist,
and that direct contact is required to induce force. Others, including Huygens, proposed
modifications to this theory to bring it into closer agreement with the astronomical observations.
But they were never completely successful. Even so it was the dominant viewpoint of scientists
until publication of Newton’s Principia.

31
5. Introduction to the Calculus

Calculus, the chief instrument of applied mathematics, was developed independently by


Isaac Newton in England and by Gottfried Leibniz in Germany. Teaching calculus can be
approached from either a mathematician’s point of view where the priority is placed on precise
definitions of limits and the like, or from a user’s point of view where the emphasis is on solving
problems. This distinction is important because in the overwhelming majority of practical
problems, the subtle concerns of the mathematicians do not come into play. Indeed the concept
of limits offered by the two inventors of the calculus, Newton and Leibniz, would hardly pass
muster in the eyes of today’s mathematicians. Moreover, the more intuitive approach of its
inventors is far easier to grasp. Calculus has two distinct elements, differential calculus and
integral calculus, which are linked by the “Fundamental Theorem of the Calculus”.

Differential Calculus

Differential calculus is all about finding the tangent at any given point on a curve. The
curve may be a graph of a position coordinate versus time, or it may be the path traced by an
object with perhaps the vertical dimension represented as a function of the horizontal dimension
as in the case of the projectile in the preceding chapter. From a mathematical standpoint it
doesn’t matter whether the independent variable is time or a position coordinate. In a more
abstract representation we have y = f (x) where f (x) denotes a function of the independent
variable x. Consider the arbitrary function in Figure 5-1. Tangents to the curve are shown for two
values of x. The tangent slopes can be computed by taking the limits shown in the figure.
€ €
y!
f (x1 + Δx) − f (x1 )
Slope = Limit of : as Δx → 0
y2! Δx

f (x 2 + Δx) − f (x 2 )
€ Slope = Limit of : as Δx → 0
Δx

y1! y = f (x)

€ x!
x1! x2!

Figure 5-1. Slopes for an Arbitrary Function of x

A derivative dy/dx is shorthand for writing such limits: that is

dy f (x + Δx) − f (x)
= Limit of : as Δx → 0
dx Δx

32

It is the fundamental operation of differential calculus. The derivative dy/dx is not to be
interpreted as a fraction, in which the d in the numerator would otherwise cancel the d in the
denominator. This notation for a derivative (and even the name “derivative”) comes from
Leibniz, and is far more widely used than Newton’s notation of employing a dot over the variable
as in y˙ . However, Newton’s notation is more compact, and is often used in the field of dynamics
when the independent variable is understood to be time.

Suppose the y-axis in Figure 1 represents the distance from some starting point O, and x
€ represents elapsed time. Then the slope at any point represents the speed. A positive slope
denotes an increase in speed and a negative slope denotes a decrease in speed.

Table 5.1 lists the derivatives for a number of simple functions y = f (x) . The derivatives
dy
are themselves functions, which we can label = g(x) .
dx

TABLE 5.1. Simple Functions and Their Derivatives

2 1 1 1
y = x x x3 xn sin x cos x
x x2 x3

dy 1 2 3
= 1 2x 3x 2 − − − nx n−1 cos x − sin x
dx x2 x3 x4

In general one will encounter more complex functions to differentiate. In those cases, the
€ following derivative relationships will prove useful. Here u and v are functions of x, while a is a
constant.
d(au) adu
=
dx dx €

d(u + v) du dv
= +
dx dx dx

d(uv) du dv
=v +u
dx dx dx

du dv
v− u
d(u /v)
= dx 2 dx
dx v

It is also common to encounter a derivative of a derivative. We call this a second


derivative written as

33
d 2 y d " dy %
= $ '
dx 2 dx # dx &

dy
If we have y = f (x) and = g(x) , it follows that
dx

d2y d " d % d
2
= $ f (x)' = g(x) = h(x)
€ dx dx # dx & dx

In the field of dynamics, velocity represents the first derivative of position with respect to
time, and acceleration€ represents the first derivative of velocity with respect to time or the second
derivative of position with respect to time.

Integral Calculus

Consider an inverse problem. We have a derivative, which is a known function g(x)

dy
= g(x)
dx

We want to determine the function, which yields this derivative. In other words, we are looking
for the “anti-derivative”, which is denoted by

y= ∫ g(x)dx = f (x)
Here ∫ g(x)dx is called the indefinite integral of g(x) . This is Leibniz’s notation for an integral
(Newton’s notation is no longer
€ used). The significance of “indefinite” will made clear shortly.

In differential calculus, we have the function f (x) and are looking for its derivative g(x) ,
€ €
while in integral calculus we have the derivative g(x) and are looking for f (x) . Finding
derivatives is straight forward, while finding an integral can be quite difficult, and in many cases
impossible. The following two examples are easy.
€ €
€ €
dy
Example 1. Suppose we have = 2x . We have previously shown that this is obtained by taking
dx
the derivative of x 2 , so we already know that ∫ 2xdx = x 2 .

dy 2
Example 2. Suppose€we have = − 3 . We have previously shown that this is obtained by
€ dx x
€ #2& 1
taking the derivative of 1/ x 2 , so we already know that − ∫ % 3 ( dx = 2 .
$x ' x


€ 34
Consider the curve in Figure 5-2 defined by the function g(x). We want to determine the
shaded area under the curve where the left boundary is x = a, and the right boundary is x = b. We
can estimate the area by dividing it into vertical strips of equal width Δx and then summing the
area of the strips. The accuracy of the estimate is improved by reducing the width of the strips. In
the limit, as the width of the strips approaches zero, we have the exact value of the shaded area.
b
It is called the definite integral between the values of x = a and b, and is written ∫ a
g(x)dx . The
limits are what distinguish a definite integral from an indefinite integral.

g(x)!

x!
a! b!
Figure 5-2. Area Under a Curve

It is one of the great mathematical discoveries that the definite integral is given by
b
∫ a
g(x)dx = f (b) − f (a)
where
f (x) = ∫ g(x)dx

is the indefinite integral. So, if the functional value f (x) can be determined, one can simply
compute the difference f (x = b) − f (x = a) to obtain the desired limit for the area under the

curve. This relationship is known as the fundamental theorem of the calculus.
€ integral cannot be found, one can attempt to
In those situations when the indefinite

simplify the problem or make some approximations. In the last resort one can employ numerical
integration, which limits the answers to specific boundary (or initial) conditions.

35
6. Newtonian Revolution

Isaac Newton (1643-1727), co-inventor of the calculus, framer of the Universal Law of
Gravitation, and formulator of the Laws of Motion, in one fell swoop, unified the theory of
motion not only on the Earth, but of the Moon about the Earth, and the planets about the Sun. He
(Figure 6-1) ushered in what has come to be known as the Newtonian Revolution.

Figure 6-1. Isaac Newton

At age nineteen, Newton entered Cambridge University, graduating with a BA degree


after three and a half years of study. Immediately thereafter the university was closed due to the
Great Bubonic Plague of 1664-65, and Newton was forced to pursue further studies on his own.
It was during this time that he created much of what is now known as the Calculus. The
remarkable thing is that Newton achieved most of his famous accomplishments when he was in
his twenties, and published very little. He was secretive and tended to keep discoveries to
himself.

Newton may not have been the first to surmise that the planets moved under the action of
force at a distance, which followed an inverse square law of attraction. Sometime after 1674,
Robert Hooke (1635-1703), the Curator of Experiments of the Royal Society, had speculated
along these lines in correspondence with Newton around 1674. And Christopher Wren, also a
member of the Royal Society, had discussed these same ideas with Newton in 1677. In 1684,
Edmund Halley (of comet fame) challenged Hooke to prove that the elliptical motion of the
planets was a unique consequence of an inverse square law of gravitation, but Hooke was not up
to the task.

36
Later, in the summer of 1684, Halley visited Newton and asked him what path a planet
would follow if governed by the inverse square law of gravitation. When Newton said it would
be an ellipse, Halley asked how he knew. Newton said he had calculated it. Halley asked to see
the calculations, but Newton had misplaced them. (It appears that Newton’s misplaced
calculations were made around 1680.) In November 1684, Newton produced a paper
documenting his proof.

This episode prompted Newton to finally document his discoveries in astronomy and
dynamics in his monumental work Philosophiae Naturalis Principia Mathematica
(Mathematical Principles of Natural Philosophy). The Principia, as it is customarily referred,
was submitted to the Royal Society in 1686 and published a year later.

Surprisingly, the Principia is devoid of the calculus that Newton used to obtain his
findings. Concerned that the reader would find the calculus too difficult, he developed
geometrical arguments to support his findings and thereby avoid the calculus. Even so, the
Principia is notoriously difficult.

However, we will not shun the calculus, and will employ it as required. As such, our
proofs are not those employed in the Principia. In what follows we will complete the
development of the Law of Gravitation initiated by Galileo, introduce the Laws of Motion,
address the orbit problem Halley posed to Newton, and finally reconcile a concern Newton had
about the gravitational attraction of a large mass upon a small mass near its surface. This latter
problem was finally solved by Newton after years of effort.

Newton’s Thought Experiment.

Neither Galileo nor Huygens appreciated the fact that the force of gravity on the surface
of the Earth was the same force that maintained the moon in orbit about the Earth and the planets
in orbit about the Sun. However, Galileo was clearly on the right track with parabolic free fall
trajectories near the surface of a flat Earth model. This is where Newton’s famous thought
experiment comes into play.

Newton envisioned a cannon on the top of a mountain firing projectiles horizontally


(Figure 6-2). If atmospheric effects were ignored, the projectile would follow the parabolic
curve derived by Galileo. Imagine that the velocity of the projectile was increased to the point
that the curvature of the Earth came into play. Not only was the ground “falling off”, but the
direction of gravity was no longer parallel to its direction at the cannon. As the initial velocity is
increased the projectile moves farther around the Earth, until at some velocity, the projectile will
move in orbit about the Earth. In this sense Newton anticipated Sputnik, almost three centuries
before that first artificial satellite was launched.

37
Figure 6-2. Newton’s Thought Experiment Envisioned Putting a Satellite Into Orbit

Newton extended this reasoning that the force, which kept the high velocity projectile in
orbit about the Earth, was the same force that kept the Moon in orbit about the Sun. And he also
reasoned that the force of gravity was not a unique property of the Earth. Gravity was also the
force that kept the planets in orbit about the Sun.

From Kepler’s Third Law Newton knew that the square of the orbital period is
proportional to the cube of the mean orbit radius ( P 2 ∝ r 3 where ∝ is the proportionality
symbol). But the force acting on the planets and the Moon must also satisfy Huygen’s
relationship for circular motion so Newton (as well as Hooke, Wren and Halley) reasoned that
gravitational force must be inversely proportional to the square of r.
€ €
And taking into account Galileo’s findings that the motion of a body in a gravitational
field is independent of its mass, Newton could conclude that the gravitational force must be
proportional to mass of the orbiting body, in order to cancel the mass on the right hand side of
Huygens’ equation.

And finally, it is reasonable to assume that the gravitational force is proportional to the
mass M of the central body. Taking these three factors into account, Newton deduced that the
gravitational force is of the form
GMm
F= 2
r

where G is some yet undetermined constant. This is the previously stated law of gravitation,
which when combined with Huygens’ force equation, re-establishes Kepler’s third law of
planetary motion: €
GM
n=
r3

where n is the orbital rate, and M is the mass of the Sun.



38
Motion of the Moon about the Earth.

In order to prove that the law of gravitation was universal, Newton had to show that it
applied to other centers of attraction besides the Sun. The motion of the Moon about the Earth
was an obvious choice. In the latter case, r represents the radius of the lunar orbit, and ME
represents the mass of the Earth. We can use it to determine the product GME:

4π 2 r 3
GM E =
P2

The period of the Moon and its mean distance from the Earth were known with considerable
accuracy in the sixteenth century. Just how these were determined is another interesting story,

which we will not go into. However, using the precise values known today

Period, P = 27.322 days r = 384,400 km

we find that GME = 3.986 x 105 km3/sec2

Galileo had shown that a body in free fall near the surface of the Earth had an
acceleration a = 9.81 m/sec2. The Greek Eralosthenes had determined the radius of the Earth to
within one percent accuracy. Taking the radius rE of the Earth to be 6374 km, we have from the
measurements of a body in free fall near the surface of the Earth:
2
GM E = arE2 = (9.81×10−3 km /s2 )(6374 km) = 3.986 ×10 5 km 3 /s2

It was some time before 1670, that Newton made this comparison between the force of
gravity at the surface of the Earth and at the Moon. So Newton had evidence that his law of

gravitation applied to the Moon orbiting the Earth as well bodies in free fall near the surface of
the Earth. With this concept of force at a distance he unraveled the mysteries of the Solar system,
and opened the door for future advances in Physics. His reasoning and the approach are the mark
of a theoretician of the highest order. And in the pecking order of those who count, the
theoretician stands tallest.

The Laws of Motion.

Newton’s three laws of motion, which are the foundation of dynamics, are as follows.

I. Every body will continue in its state of rest or of uniform motion in a straight line
except in so far as it is compelled to change that state by impressed force.

II. Rate of change of momentum is proportional to the impressed force and takes place
in the line in which the force acts.

III. Action and reaction are equal and opposite.

39
The term “uniform motion” means constant velocity. And momentum is defined as the
product of mass and velocity. The term velocity has meaning only with respect to a specified
frame of reference, which, for convenience, will be taken to be a Cartesian coordinate system
with a specified origin and orientation. When examining the motion of a planet in the Solar
system, Newton found that the equations of motion take on a simple form when the origin of the
frame of reference is at the mass center of the bodies comprising the Solar system, and the axes
do not rotate with respect to the star background. When examining the motion of the Moon under
the gravitational attraction of the Earth, a frame of reference with origin at the Earth-Moon mass
center and with axes that do not rotate with respect to the star background is suitable. When
examining the motion of a stone hurled into the air, a frame of reference fixed to the surface of
the Earth, with one axis vertical and another axis oriented parallel to the direction of motion is
suitable. Such frames of reference are often referred to as inertial frames.

Consider a Cartesian frame with xyz coordinates of a particle of mass m acted upon by a
force F. In mathematical terms, the second law of motion for the particle is

Fx = max , Fy = may , Fz = maz

Acceleration in a given direction depends only on the component of force in that direction. In
vector form this becomes simply F = ma . Note that a vector defines both the magnitude and the
direction. In order to€distinguish a vector from its scalar value it is customary to display it in
bold type.

A remarkable aspect of this law of motion is that the equations of motion have the same
form in other frames of reference that translate at a uniform velocity with respect to the primary
inertial frame of reference. This principle is sometimes referred to as Newtonian relativity.

The fact that Newton’s Second Law of Motion involves acceleration, which is the rate of
change of velocity, which in turn is the rate of change of position, means that we must deal with
differential equations, so-called because they contain derivatives. So at this point we introduce
the calculus of the previous chapter. Newton’s second law can be expressed as six first-order
differential equations of motion

dv x Fx dv y Fy dv z Fz
= , = , =
dt m dt m dt m
dx dy dz
= vx, = vy , = vz
dt dt dt

or equivalently the three second-order differential equations



d 2 x Fx d 2 y Fy d 2 z Fz
= , = , =
dt 2 m dt 2 m dt 2 m

In either case, we are dealing with a sixth-order set of differential equations.


40
Presumably the force components can be expressed as functions of the six dependent
variables (x, y, z, v x , v y , v z ) and the independent variable t. The task is to find the solutions of
the six dependent variables as functions of time for a specified set of initial values of the
dependent variables.
€ In some cases the motion lies in a single direction. In such a situation, it makes sense to
choose one of the xyz axes to lie in that direction so you need only need to determine the
velocity and position in one direction. That was the case when Galileo supposedly dropped
objects from the Leaning Tower of Pisa.

Projectile Problem Revisited.

Much more likely, the motion lies in a plane, so for this two-degree of freedom problem,
one must deal with a fourth-order set of differential equations. That was the case when Galileo
launched projectiles from a height h, and established that the path followed by the projectile was
a parabola. In that case we choose the x axis to be horizontal, and the y axis to be vertical.
Galileo arrived at his solution by experimentation. Now we will show that his results were
consistent with the Second Law of Motion. In the horizontal direction dv x /dt = 0 , so the velocity
component v x is a constant equal to its initial value. The horizontal distance is then given by the
integral
x = v xo ∫ dt + c1 = v xo t €

where we have assumed that the x = 0 at t = 0 so that the constant c1 is also zero.
€ the acceleration is given by dv /dt = −g where g is the constant
In the vertical direction y
acceleration due to gravity. The minus sign indicates that the gravity acts in the minus y

direction. The vertical velocity is given by the integral

v y = −g ∫ dt + c 2 = −gt

where the constant c 2 is the initial velocity in the x direction, taken to be zero. The vertical
position is given by the integral
€ 1
y = −g ∫ tdt + c 3 = h − gt 2
2

where the constant c 3 is equal to the initial height h. What we have just demonstrated is that the
Second Law of Motion makes it possible to predict motion for a given set of initial conditions
€ of the force is known a priori. In this case the force was along the
provided the functional form
y-axis with value –mg.

The choice of coordinates was logical because there was no force in the horizontal
direction, and we immediately deduced that the horizontal velocity was equal to its initial value.
Thus we were able to immediately reduce the order of the problem. It is often the case that the

41
choice of variables is crucial in finding a solution to a problem. However, you should realize that
there are no set rules for choosing variables. Often it is by trial and error.

The Two-Body Problem.

The classic two-body problem is one in which two bodies move under their mutual
gravitational attraction. Referring to Figure 6-3, point O is taken to be any point with zero
acceleration. The vectors r1 and r2 locate the masses m1 and m2 with respect to O. Let r denote
the vector difference
r = r2 − r1
€ magnitude
€ €
so that r is the scalar of the distance between the two bodies.
€ r
m1 !

€ r1 r2

€ €O!

Figure 6-3. Vector Location of Two Point Masses

Then according to the universal law of gravitation and Newton’s second law, the motion
of the two bodies are defined by the vector equations

Gm1m2
m1r˙˙1 = − (r1 − r2 )
r3
Gm1m2
m2r˙˙2 = − (r2 − r1 )
r3

These two vector equations can be uncoupled in the following manner. Divide the first equation
by m1 and the second equation by m2. Then subtract the first resulting equation from the second
resulting equation to obtain€

G( m1 + m2 )
r˙˙ = − r
r3

which describes the motion of body 2 with respect to body 1. Observe that the resulting equation
of motion involves the sum of the two masses. In the case when m1 is the mass of the Sun and m2

is the mass of the Earth, we have

mSun + mEarth = 1.000003 mSun

and in the case when m1 is the mass of the Earth and m2 is the mass of the Moon, we have
€ 42
mEarth + mMoon = 1.01226 mEarth

This process reduces the two-body problem to the Kepler (one-body) problem.

The Kepler Problem €

Consider now the problem of planetary motion about the Sun, the same problem Halley
asked Newton about in 1684. This is also a two degree-of-freedom problem, with the motion
lying in the plane defined by the radius vector r (along the line joining the Sun and the planet)
and the velocity vector v (along the direction of motion) of the planet and any instant in time.
Now choose a non-rotating x and y coordinate frame with origin at the Sun, and lying in the
plane of motion. According to the universal law of gravitation, the force on the planet is in the
direction of the Sun, and is inversely proportional to the square of the distance to the Sun.
Accordingly, the second law of motion provides

d2x µm x
m = Fx , Fx = − ⋅
dt 2 r2 r
d2y µm y
m 2 = Fy , Fx = − 2 ⋅
dt r r

Here m represents the mass of the planet, and the constant µ is the product G(M + m) where M
is the mass of the Sun, and G is the constant of gravitation. The factors x/r and y/r are the non-
dimensional quantities, €which resolve the components of the force vector F along the x and y
coordinates. The distance r is given by

r = x2 + y2

Because the mass of the planet cancels out, the equations of motion can be written as

d2x x € x
= −µ 3/2
x˙˙ = −µ 3/2
dt 2 (x 2 + y 2) (x 2
+ y 2)
or in Newton’s notation
d2y y y˙˙ = −µ
y
2
= −µ 3/2 3/2
dt (x + y 2)
2
(x 2
+ y 2)

Note that in Newton’s notation a single dot over a variable represents a first derivative with
respect to time, and two dots over a variable represent a second derivative with respect to time.
€ €
The above set of differential equations is fourth order with two derivatives in x and two
derivatives in y. Note also that the equations are coupled (they both depend on x and y), which
makes this a more difficult problem than Galileo’s projectile moving over a “flat” Earth.

43
Observe that if we multiply the second equation by x and the first equation by y, and
compute the difference we obtain
d
xy˙˙ − yx˙˙ = ( xy˙ − yx˙ ) = 0
dt

Because the derivative of x˙y − y˙x is zero, it follows that


€ x˙y − y˙x = h

where h is the so-called momentum constant. Note that h is the magnitude of the vector h = r × v
where v is the velocity defined by v = dr/dt. The vector h is normal to the plane of the orbit.

Returning to the original differential equations, if we multiply the first by x˙ and the

second by y˙ and add the two equations together we obtain

x˙x˙˙ + y˙ y˙˙ + µ
( xx˙ + yy˙ ) =0 €
2 3/2
€ (x 2
+y )
which is equivalent to
d #1 2 µ &
% ( x˙ + y˙ 2 ) − (=0
€ dt %$2 2 2
x + y ('

Consequently
1 2 µ

2
( x˙ + y˙ 2 ) − =E
x + y2
2

where E is the so-called energy constant. The quantity x˙ 2 + y˙ 2 is equal to the square of the
velocity. Accordingly, the first term in the energy equation is called the kinetic (or motion)

energy T, and the second term is called the potential energy V.
€ along the radial direction r, it is advantageous
Because the gravitational force is directed
to perform a coordinate transformation from the x and y rectangular coordinates to the r and υ
polar coordinates. The coordinate transformation and their derivatives are given by

x = r cosυ

y = r sin υ
x˙ = r˙ cosυ − rυ˙ sin υ
y˙ = r˙ sin υ + rυ˙ cosυ
˙x˙ = ˙r˙cosυ − 2 r˙υ˙ sin ν − rυ˙˙ sin υ − rυ˙ 2 cosυ
y˙˙ = ˙r˙sin υ + 2 r˙υ˙ cos ν + rυ˙˙ cosυ − rυ˙ 2 sin υ

It follows that

44
˙x˙ cosυ + ˙y˙ sin υ = ˙r˙ − rυ˙ 2
x˙y − y˙x = r 2υ˙

Thus, the equations of motion in polar coordinates are


€ µ
˙r˙ − rυ˙ 2 = −
r2
r 2υ˙ = h

The second equation can be used to eliminate υ˙ from the first equation, which then becomes
€ h2 µ
˙r˙ − 3
=− 2
€ r r

In polar coordinates the previously derived energy integral is written



1 2 2 2 µ 1 " 2 h2 % µ
2
(r˙ + r υ˙ ) − r = E or equivalently
2#
$ r˙ + 2 ' − = E
r & r

Kepler’s Second Law Revisited. Recall that Kepler stated that equal areas are swept out
in equal€periods of time. Newton’s geometric proof was presented in Chapter 2. Here we present
an analytic proof for the same law. Figure 6-4 shows € an arc of an ellipse with focus at O. Points
P and Q are on the ellipse. The area of the triangle QOP is given by

r( r + Δr)
ΔA = ⋅ sinΔυ
2

The change in area per unit time is given by



ΔA r( r + Δr) sinΔυ Δυ
= ⋅ ⋅
Δt 2 Δυ Δt

sinΔυ
As Δυ decreases, r + Δr approaches r, and approaches unity. Then in the limit as Δt
Δυ

approaches zero, we have
dA r 2 dυ
€ € = ⋅ €
€ dt 2 dt

But previously we showed that r 2υ˙ is the momentum constant h. Therefore the time rate of
change of area is a constant, which establishes Kepler’s second law.

45
Q!
r + Δr

Δυ P!
€ r

O!

Figure 6-4. Area Swept in a Unit of Time

Kepler’s First Law Revisited. We are now at the heart of the question posed by Halley
to Newton concerning the shape of the orbit being dictated by an inverse square law of
gravitation. The proof presented below (and the one given in most texts on planetary motion)
employs two changes of variables, one involving the independent variable t, and the other
involving the dependent variable r. Because the shape of the orbit describes r in terms of υ ,
rather than r in terms of time t, we will use the equation for υ˙ , to change the independent
variable from t to υ . Also some clever individual found that the problem could be simplified by
introducing the variable u defined as the reciprocal of r. So we will make use of the following

relationships obtained using differential calculus


1
u=
r
dr 1 du 1 du dυ du
=− 2 =− 2 = −h
dt u dt u dυ dt dυ
2 2 2
d r 1 d du d u dυ 2 2 d u
= − = −h = −h u
dt 2 u 2 dt dυ dυ 2 dt dυ 2

Using the above relationships, the equation of motion


€ h2 µ
˙r˙ − 3
=− 2
r r

is transformed into the following differential equation



d 2u
+u=C
dυ 2

where C is the constant µ /h 2 . The solution to this equation is


€ 46
u = C + Acosυ + Bsin υ

where the constants of integration A, B and C, are evaluated from initial or boundary conditions.
This is easily confirmed by taking the derivatives

du
= −Asin υ + Bcosυ

d 2u
= −Acos υ − Bsin υ
dυ 2

d 2u
and substituting the values for u and into the differential equation.
€ dυ 2

The equation for the shape of the orbit in terms of the original dependent variable r is
€ 1
r=
C + Acosυ + Bsin υ

All that remains is to evaluate the constants A, B and C. It is clear that this equation is identical
to the equation for an ellipse

a(1− e 2 )
r=
1+ ecosυ

when the constants of integration are assigned the following values:



h2 e
C= = a(1− e 2 ), A= , B=0
µ a(1− e 2 )

This constitutes the proof that an elliptical orbit is consequence of the inverse square law of
gravitation.

Energy Integral and Velocity. Previously we established the energy integral E, which
we can now evaluate in terms of the orbit parameters. At the point of closest approach, we have:
υ = 0, r = a(1− e), r˙ = 0 . Substitution into the energy integral provides:

1 " 2 h 2 % µ µ " h 2 /µ % µ µ " 1+ e % µ


E = $ r˙ + 2 ' − = $$ 2 '
2'− = $ '−
€ 2# r & r 2 # a (1− e) & a(1− e) 2 # a(1− e) & a(1− e)

which reduces to
µ
€ E =−
2a

€ 47
The energy integral can also be used to solve for the magnitude of the velocity

h2
v = r˙ 2 +
r2

at any point in the orbit. Using the value of the energy constant, we have
€ #2 1&
v = µ% − (
$ r a'

so that
€ =
µ #1+ e &
v = v max % ( at r = rmin = a(1− e)
a $ 1− e '
µ # 1− e &
v = v min = % ( at r = rmax = a(1+ e)
a $1+ e '

Deduction of Kepler’s Equation. The above solutions describe the shape of the orbit,
€ the time it takes to move between any two points in the orbit. For this we
but do not specify
return to the energy integral
1 " 2 h2 % µ µ
$ r˙ + 2 ' − = −
2# r & r 2a

which can be used to determine time t as an explicit function of r. Because the equation depends
only on r and its time derivative, it can be rewritten as the integral equation

dr
t= ∫ +τ
2 2
rmin −µ /a + 2µ /r − µa(1− e ) /r

where rmin is the minimum value of r, and the constant of integration τ is the time when r is at its
minimum value.

In Chapter 3, we showed that

a(1− e 2 )
r= = a(1− ecos χ )
1+ ecosυ

Substituting the latter expression for r into the above integral equation we obtain

a3
t −τ =
µ
∫ (1− ecos χ )dχ

€ 48
which solution is Kepler’s equation (geometrically derived in Chapter 3)

M = n ( t − τ ) = χ − esin χ

where the quantity n = µ a 3 corresponds to the mean orbit rate, and M is the mean anomaly.

Perspective. So what did Newton accomplish besides showing that Kepler’s laws of
planetary motion (derived from planetary observations) were consistent with the Law of
€ In dealing with Kepler’s problem (the motion of two point masses with respect to
Gravitation?
one another), Newton knew the solution (Kepler’s laws) before he started. But in more complex
problems one does not know the answers beforehand, and Newton’s Laws of Motion allow one
to predict the answers before observations are even made.

Unfortunately, exact analytic solutions such as Kepler’s equations to the two-body


problem are often unattainable. This leaves one with two choices: either to seek an approximate
analytic solution often using series approximations, or to obtain particular solutions using
numerical integration of the exact equations of motion. An approximate analytic solution can
provide insight into the sensitivity over all parameters and initial conditions, but may be difficult
to obtain and often requires a high degree of mathematical ingenuity. With the advent of high
speed computing the trend has been to use numerical integration to obtain particular solutions
(valid only for a specific set of parameters and initial conditions). The latter may also be used to
test the validity of approximate analytic solutions.

49
7. Gravitational Attraction of a Sphere

When considering the motion of the moon about the Earth, we can think of the Earth and
the Moon as being point masses, because the distance between the Earth and the Moon is so
much greater than the radius of either the Earth or the Moon. But this is not the case for bodies in
free fall near the surface of the Earth. How are we to sum up the forces from all portions of the
Earth? To simplify matters, assume that the Earth has a uniform mass distribution. Because of
symmetry we can logically conclude that the net sum of all the forces lies on a line from the
particle toward the center of the Earth, so we can introduce an effective gravitational center. But
where along the line joining the center of the Earth and the particle do we locate this effective
gravitational center? This is the question that vexed Newton?

Do we place the effective gravitational center at the geometric center of the sphere as
shown on the right hand side of Figure 7-1? One might question this choice because those points
making up the hemisphere closer to m would exert more force than those points making up the
other hemisphere. This would suggest that the effective gravitational distance is less than s.

m! m!

F! s F! s

R!
€ € F!

M!
4
M = ρVsphere = πR 3 ρ
M! 3
ρ = mass density
GMm F Vsphere 4R 3
F= 2 L ≡ = = 2
s Gmρ s2 3s

Figure 7-1. Gravitational Attraction of the Earth on a Near Point-Source Particle
€ €
Dumbbell Problem. Lets carry this thought further. Replace the spherical Earth by two
point masses each with half the total mass, one located at the centroid of the upper hemisphere
and the other located at the centroid of the lower hemisphere. Link them by a rod of zero mass to
keep them separated at a fixed distance. We call this a “dumbbell. Now compute net
gravitational force exerted by the dumbbell on m. The geometry is shown in Figure 2 on the left.
In the diagram on the right, we replace the dumbbell by a single point with the same total mass
M located a distance s’ from m, so selected as to provide the same force as exerted by the
dumbbell. We see that s’ is shorter than s, which shows that the effective gravitational center of a
dumbbell is not at its center.

50
m! m!
F! F!

s"
s!

M/2!

c!

M!

Equivalent!
system!
c!
3
M/2! Suppose c = s
8
GMm
F= 2
GMm GMm (s")
F= 2 + 2
2( s − c ) 2( s + c ) (1− 3/8)(1+ 3/8) s = 0.8047s
2 s'=
GMm 1+ (c /s) 1+ (3/8) 2
= 2
⋅ 2 2
s (1− c /s) (1+ c /s)
Figure 7-2. Dumbbell and Its Equivalent
€ Gravitational Center
€ This thought experiment would seem to imply that effective gravitational center of a
sphere is not at its center. However, the implication is wrong. Indeed Newton showed that for the
purpose of computing the gravitational attraction, a sphere can indeed be replaced by a point
mass as its center (as in Figure 7-1). So why did we go through the exercise with the dumbbell?
The reason was to drive home the point that the correct answer is not obvious. In fact, Newton
delayed the publication of his famous treatise on dynamics some twenty years, largely because
he could not come up with a suitable mathematical proof for this problem. In order to do so he
had to invent integral calculus, the complement to differential calculus.

A Proof. The problem posed is to prove that in applying the Law of Gravitation a sphere
can be replaced by an equivalent point mass located at its center. The proof offered below differs
from the one presented by Newton. The proof is conceptually outlined in Figure 7-3. First the
sphere is sliced into disks perpendicular to the direction of the small point mass m. Then each
disk is divided into rings, and each ring is divided into opposing segment pairs. The net force
along the line from the center of the sphere to the mass m is computed for each segment pair.

51
The force for each ring is then computed by summing up the segment pair forces. The force for
each disk is the sum of the forces produced by all the rings in a disk. Finally the force produced
by the sphere is obtained by summing the forces produced by all the disks. We then compare this
force to that which would be obtained by a point located at the center of the sphere, but having
the same mass as the sphere.

2. Divide each disk into rings!


3. Divide each ring into segment pairs!

1. Divide the sphere into thin!


disks perpendicular to the!
attracting object!

m!

m!

5. By summing segment pair F, !


compute each ring F!
!
6. By summing ring F, compute!
F!
each disk F!
!
7. By summing disk F, compute! 4. Compute net force F!
sphere F! due to segment pairs!

Figure 7-3. Outline of a Proof

The actual process requires the use of integral calculus, and a considerable number of
algebraic manipulations. The entire proof is presented in Figure 7-4. The summing process
described above is performed using integrals available in standard Tables of Integrals. My old
standby is A Short Table of Integrals, by B. O. Pierce, published by Ginn and Companyin 1929,
when Pierce was a Professor of Mathematics at Harvard.

52
Ring volume, Vring = 2πρdρdz
ρ


2
€ r 2 = ρ 2 + (s − z)

ρ a2 = R 2 − z 2
€ 2π $ s − z ' 2π ( s − z) ρdρdz
ρ Lring = & ) ρdρdz =
r2 % r (

2
[
ρ + ( s − z ) ⋅ ρ 2 + ( s − z)
2 2
]
€ ρa
€ ρdρ
Ldisk = 2π ( s − z) dz ∫
Lring r s 0
2
[
ρ + ( s − z ) ⋅ ρ 2 + ( s − z)
2 2
]
ρa
' *
1 s− z
= −2π ( s − z) dz = 2π )1− ,dz
€ ρ 2 + ( s − z)
2 ) 2 2 ,
ρ a + ( s − z) +
(
€ dz
0
€ z ' *
R = 2π )1−
s− z ,dz
) 2 2
R − z + ( s − z) +
2 ,
(
O!
ρa ' *
€ € = 2π )1−
s− z
, dz
€ ( R 2 + s2 − 2sz +



m!
R R
s− z
Lsphere = 2π ∫ dz −2π ∫ dz
F! −R −R R + s2 − 2sz
2

s % R
R 2 + s2 + sz
R (

= 2π '2R + R 2 + s2 − 2sz − R 2
+ s 2
− 2sz *
'& −R 3s2 −R *
)
%2R + R 2 + s2 − 2sR − R 2 + s2 + 2sR (
' *
= 2π ' R 2 + s2 + sR
€ z 2 2
R + s − 2sR +
2 2
R + s − sR 2 2 *
R + s + 2sR *
'&− 3s2 3s2 )
R

M! Assume s > R, Then


€ €
€ $ R 2 + s2 + sR R 2 + s2 − sR '
Lsphere = 2π &2R + ( s − R) − ( s + R) − 2 ( s − R) + ( s + R))
% 3s 3s2 (
3
4 πR V
= = sphere
3s2 s2

GmVsphere
F= QED
s2

Figure 7-4. Proof in Detail


53
8. Gravitational Constant and the Mass of the Earth

The law of gravitation depends on the gravitational constant G. However, Newton was
able to explain the motions of the Solar system without knowing the value of G. To see why this
was possible, consider the motion of the Moon about the Earth. The gravitational force acting on
the Moon depends on the mass of the Earth M and the mass of the Moon m. But since the second
law of motion also depends m, m cancels out, leaving the product GM. So one can determine the
motion of the Moon knowing only the product GM and not its components G or M. And the
product GM can be determined knowing the orbit. So Newton was not particularly concerned
with the determination of G. Moreover, he did not believe it was practical to determine G by any
experiment, because the gravitational force between any masses on the Earth would be
overwhelmed by the mass of the Earth itself. However, in 1797, the English chemist and
physicist, Henry Cavendish (1731-1810) performed an experiment capable of predicting G to
within a percent of the currently accepted value

G = 6.674 ×10−11 m3s-2kg−1

Cavendish Experiment. The Cavendish experiment was actually conceived by John


Mitchell, a friend of Cavendish. But Mitchell died before he could carry it out. Mitchell’s
experimental apparatus was € eventually passed on to Cavendish who performed a series of
experiments using this device. The apparatus is schematically represented in Figure 8-1a. A
small lead sphere was connected to each end a rod that was suspended horizontally on a very fine
wire. This “torsion bar” assembly was inside a class enclosure to protect it from air currents.
Outside and immediately adjacent to the enclosure were two large lead spheres, which were
suspended in a manner that allowed them to rotated to any fixed position around the axis of the
torsion bar wire. The idea was to displace the large spheres with respect to the null position of
the small sphere (Figure 8-1b) and measure the change in rotation of the torsion bar. At the new
equilibrium position of torsion bar the gravitational torque on the torsion bar will exactly cancel
the restraining torque of the wire:
TGravity = Twire = kθ

So if the torsion constant k of the wire were known, and the rotation angle θ from the torsion bar
null were measured, it would be possible to determine the torque due to gravity.

Before the large lead spheres were in place, the torsion constant k was determined by
rotating the wire connection point so as to induce an oscillation of the torsion bar. The frequency
f of the oscillation was then measured. Then knowing the moment of inertia I of the torsion rod,
the constant k could be computed using the relationship

k
f = 2π
I

So this is how the torsion bar spring constant is calibrated.


54
Returning to full experiment with the large spheres in place, the gravitational torque can
be computed from the measured value of θ and the computed value of k. Then knowing the
values for the masses m and M and their separation distances at the offset equilibrium position,
the gravitational constant G can be computed. Today versions of this experiment are routinely
performed as part the curriculum of many introductory physics classes.

θ"
Tread!

Glass!
enclosure!

M!
m! m!

a) Side View before Displacement! b) Top View after Displacement!

Figure 8-1. Cavendish/Miller Torsion Bar Experiment

Mass of the Earth. Because of his experiment, Cavendish is credited as the man who
weighed the Earth This is because the product GME was known ( 3.986 ×10 5 km3 /s 2 ). Dividing
by the now known value of G, the mass of the Earth ME can be determined ( 5.97 ×10 24 kg ).


55
Mach, Ernst, The Science of Mechanics, The Open Court Publishing Company, 1960.

Hoyle, Fred, Astronomy, Cresent Books, Inc. (no date)

Gamow, George, Gravity, Dover Publications, Inc., Mineola, N.Y. 2002

Gribbon, John, The Scientists, Random House, New York, 2004.

Moulton, Forest, R., An Introduction to Celestial Mechanics, MacMillan Company, New York,
1914.

Plummer, H. C., An Introductory Treatise on Dynamical Astronomy, Dover Publications, Inc.


New York, 1960. (Originally published in 1918)

56

You might also like