Second Ed
Second Ed
DYNAMICS
Second Edition
Patrick Hamill
Contents
Chapter 2. Kinematics 29
2.1. Galileo Galilei (Optional Historical Note) 30
2.2. The Principle of Inertia 31
2.3. Basic Concepts in Kinematics 32
2.4. The Position of a Particle on a Plane 43
2.5. Unit Vectors 44
2.6. Kinematics in Two Dimensions 47
2.7. Kinematics in Three Dimensions 51
2.8. Summary 60
2.9. Problems 61
3.8. Summary 98
3.9. Problems 99
1
2 PREFACE
Acknowledgements
I thank my colleagues at San Jose State University and NASA Ames
Research Center, particularly Drs. Alejandro Garcia and Michael Kauf-
man. I am especially indebted to the many students in my mechanics
courses whose influence on this book cannot be overestimated.
Part 1
Principles of Mechanics
Chapter 1
A Brief Review of
Introductory Concepts
1.1. Kinematics
Kinematics is defined as the study of motion. Essentially, kine-
matics involves determing the relationships between position, velocity,
acceleration, and time.
9
10 1. A BRIEF REVIEW OF INTRODUCTORY CONCEPTS
and Z Z
dr = vdt.
Exercise 1.1. You were driving your new Ferrari at 62 mph (= 100
km/hr) when you spotted a police car. Naturally, you hit the brakes.
You slowed to 31 mph, covering a distance of 50 m. (a) What is your
constant acceleration? (b) How much time did it take to slow to 31
mph? Answers: (a) -5.79 m/s2 (b) 2.4 sec.
friction is 0.4. a) Draw the free body diagram. What forces act on the
block? b) What is the tension in the string? c) If the string is cut,
what is the acceleration of the block? Answers (b) T = 37.63 N (c)
a = 1.51 m/s2 .
30o
dW F · dx
P = = .
dt dt
Since dx/dt ≡ v, another expression for power is P = F · v.
1.4. Momentum
A moving particle is characterized by having a particular momen-
tum. When we use the term “momentum” we usually are referring to
the linear momentum, not to be confused with the angular momentum,
which we will define in a few moments.
1.5. ROTATIONAL MOTION 15
For a fixed, stationary axis, the center of the wheel is at rest. All
other points are moving in circles around it. If you looked straight down
the axis you would see the circle shown on the right side of Figure 1.4.
Point P is on the rim of the wheel. The angular position of P is given
by the angle between some fixed line and the radius vector to P. If the
wheel is turning, after a time dt point P will have moved a distance
ds to P 0 . Recall from geometry that ds = rdθ where dθ (in radians of
course!) is the angle subtended by the arc P P 0 = ds. Point P moves
with speed
ds rdθ
v= = = rω.
dt dt
The speed of point P is called the “tangential speed” because instan-
taneously P is moving tangent to the rim of the wheel. It is sometimes
convenient to write the tangential speed as vT . Then
vT = rω. (1.12)
Taking the time derivative of vT yields the tangential acceleration
aT ,
d2 θ
dv d dθ
aT = = r = r 2,
dt dt dt dt
dω
where we used the fact that r is constant. But dt
= α, so
aT = rα. (1.13)
Axis of
rotation
r
F
N ≡ r × F. (1.14)
In Equation (1.14) r is the vector from the axis of rotation to the
point of application of the force. The direction of the torque is perpen-
dicular to the plane defined by r and F and it is usually represented
as a vector along the axis of rotation.
Exercise 1.8. Show that a given force applied at any point along
the line of action yields the same torque.
Exercise 1.9. An athlete is holding a 2.5 meter pole by one end.
The pole makes an angle of 60◦ with the horizontal. The mass of the
pole is 4 kg. Determine the torque exerted by the pole on the athlete’s
hand. (The mass of the pole can be assumed to be concentrated at the
center of mass.) Answer: Torque = 24.5 N m.
1.5. ROTATIONAL MOTION 19
1
T = Iω 2 . (1.17)
2
Note the similarity with the translational kinetic energy T = 21 mv 2 .
l = Iω, (1.18)
2
where I = mr .
This relation can also be expressed as l = mr2 ω or as l = mvr.
But angular momentum is a vector, so
l = Iω
where ω is the angular velocity vector, directed along the axis of rota-
tion.
For an extended rigid body the total angular momentum is usually
denoted by L and, as you might expect, is given by
L = Iω
2
P
where now I = i mi ri⊥ .
22 1. A BRIEF REVIEW OF INTRODUCTORY CONCEPTS
1.6. Summary
The position of a particle is given by the vector r. Velocity is
defined by
dr
v= ,
dt
and the acceleration is
dv
a= .
dt
Newton’s second law for a particle with constant mass is
F = ma. (1.20)
Work is defined by the integral of the dot product of force and
displacement, Z
W = F · dr. (1.21)
1.7. PROBLEMS 23
1.7. Problems
Problem 1.1. Two ships are sailing in a thick fog. Initially, ship
A is 10 miles north of ship B. Ship A sails directly east at 30 miles
per hour. Ship B sails due east at constant speed vB then turns and
sails due north at the same speed. After two hours, the ships collide.
Determine vB .
24 1. A BRIEF REVIEW OF INTRODUCTORY CONCEPTS
car starts immediately and accelerates at 8 miles per hour per second.
At that same moment the teenager steps on the gas, but his car only
accelerates at 2 miles per hour per second. (a) How far from the starting
point does the police car overtake the speeder? (b) How fast are they
going at that time? (c) Why is the speed you calculated for the police
car unrealistic?
Problem 1.8. A brick is on a wooden plank that is resting on a
table. One end of the plank is slowly raised so that it forms an angle
θ with the horizontal table top. When θ = 60◦ the brick starts to slide
down the plane. (a) Draw the free body diagram. (b) Determine the
coefficient of static friction between brick and plank. (c) If the coef-
ficient of sliding friction is one half of the coefficient of static friction,
determine the acceleration of the block.
Problem 1.9. A block of mass M is on an inclined plane with a
coefficient of sliding friction µ. At a given instant of time the block is
at some point P on the plane and is moving up the plane with a speed
v0 . (a) Obtain the time for the block to reach its highest point relative
to P. (b) Obtain the time for the block to slide back down to point P.
(c) Obtain an expression for the velocity of the block at the time it
returns to P.
Problem 1.10. A toy rocket burns fuel for 1.5 seconds. During
that time, it accelerates upwards at 60 m/s2 . It then coasts upwards
to some maximum altitude before falling back down. Determine the
maximum altitude reached.
Problem 1.11. Atwood’s machine consists of two weights (M1 and
M2 ) suspended at the ends of a string that passes over a pulley. Assume
massless, inextensible strings and a frictionless pulley. Let M1 =6 kg
and M2 = 5.5 kg. The masses are released from rest. Determine the
distance descended by the 6 kg mass when its velocity reaches 0.5 m/s.
Problem 1.12. (a) Determine the rotational kinetic energy of a
wheel of your bicycle when your linear speed is 20 km/hour. You
may assume the wheel is a hoop of mass 1.5 kg and radius 30 cm. (b)
Compare your result with the translational kinetic energy of the wheel.
(c) Give a theoretical explanation for the relationship found in parts
(a) and (b).
Problem 1.13. The frictional force between water and seabed in
shallow seas causes an increase in the day by about 1 ms/century.
Determine the torque that causes this change. Assume the Earth is a
uniform sphere.
26 1. A BRIEF REVIEW OF INTRODUCTORY CONCEPTS
reasonable. Obviously, the policeman will catch the teenager, but your
answer is not reasonable if the distance required is hundreds of miles!)
Chapter 2
Kinematics
29
30 2. KINEMATICS
that Galileo changed the world and our understanding of it. Some
people condemn the Church authorities for trying to stop scientific
progress; others are more understanding of the authorities who saw
their entire world view threatened by a revolutionary iconoclast.
completely specify the position of the bead, no matter how the wire
may be twisted or curled. However, the linear distance may not be the
most convenient parameter to use. For example, the position of a bead
sliding on a circular hoop might best be described in terms of an angle.
The simplest example of one-dimensional motion is a particle mov-
ing along a straight line, such as the x-axis. The acceleration is
dv
a= .
dt
Here I wrote a and v rather than a and v because if the motion is one-
dimensional there is no need to use vectors. Multiplying both sides by
dt gives the following very simple differential equation:
dv = adt.
To solve, you integrate both sides:
Z vf Z tf
dv = adt,
v0 t0
where the limits indicate that the time runs from t0 to tf (that is,
“initial” time to “final” time), and the initial and final values of the
velocity are v0 and vf .
The quantity t0 represents the time when the particle was at the
initial position x0 . Since the starting time for any process is arbitrary
(it just depends on when you push the button on the stopwatch) it is
nearly always set equal to zero (t0 = 0). R
Now you can go no further. You cannot evaluate the integral adt
because you do not have an expression for the acceleration as a function
of time. But if you do know the acceleration as a function of time,
a = a(t), then you can carry out the integration. When you do so,
you will obtain an expression for the velocity as a function of time,3
v = v(t). Once you have determined the velocity, you can continue the
analysis and write
dx = v(t)dt,
which follows immediately from the definition of velocity.
Integrating once again, you will obtain an expression for the posi-
tion as a function of time,
x = x(t),
or, more explicitly,
x = x(t, v0 , x0 ).
3To be precise, this will give you an expression for the velocity in terms of the
time, the initial time and the initial velocity, v = v(t, t0 , v0 ).
34 2. KINEMATICS
the time for the projectile to reach the top of its trajectory. Since
vy = v0y + ay t we have
0 = v0y − gttop
so
ttop = v0y /g.
Worked Example 2.1. Consider the projectile motion of a
cannon ball. Assume the initial speed is v0 and the cannon is aimed
at an angle θ above the horizontal. Determine the range.
Solution: The range (R) is the horizontal distance traveled
by the projectile along a flat horizontal plane; it is the value of
x when y = 0. The initial velocity components are v0x = v0 cos θ
and v0y = v0 sin θ. Since the total time of flight is twice the time
required to reach the top of the trajectory, the range is given by
x(t = 2ttop ). It is easy to obtain a formula for the range in terms
of the initial velocity as follows:
vy = v0y − gt.
But vy = 0 when t = ttop so 0 = v0y − gttop = v0 sin θ − gttop .
Consequently,
ttop = (v0 sin θ) /g.
Therefore,
R = v0x (2ttop ) = 2v0 cos θ (v0 sin θ) /g,
v02
R=2 sin θ cos θ. (2.5)
g
and consequently
g (x − x0 )2 x − x0
y=− 2
+ v0y + y0 ,
2 v0x v0x
g v0y v0y
y = − 2 (x2 − 2xx0 + x20 ) + x− x0 ,
2v0x v0x v0x
g g v0y gx2 v0y
y = − 2 x2 + [( 2x0 + )]x + (− 20 − x0 + y0 ).
2v0x 2v0x v0x 2v0x v0x
This has the required form as can more easily be appreciated if we
set x0 = 0 and y0 = 0 to obtain
g 2 v0y
y=− x + x.
2v0x v0x
vT = ω × r. (2.7)
v(t+Δt)
Δv v(t)
Δv
v(t) v(t+Δt) Δθ
Δθ
By definition4,
dv v(t + ∆t) − v(t) ∆v
a= = lim = lim .
dt ∆t→0 ∆t ∆t→0 ∆t
Tip-to-tail addition of the vector v(t + ∆t) and the vector −v(t)
shows that ∆v is a vector pointing towards the center of the circle.
To help you see this, on the left side of Figure 2.2, I drew the vector
∆v at a midpoint between the vectors v(t) and v(t + ∆t). You can
appreciate that the vector ∆v is pointing approximately towards the
center of the circle. As ∆t → 0, the vector ∆v points more and more
closely towards the center.
The same concept is shown in a different way on the right hand side
of Figure 2.2. If the angular velocity is constant then the speed (that
is, the magnitude of the velocity vector) is a constant equal to v, but
the direction of the vector changes with time. The tip of the velocity
vector traces out a circle, as indicated on the right hand side of Figure
2.2. (A plot showing the path of the tip of the velocity vector is called a
hodograph.) For circular motion at constant speed, the hodograph is a
circle of radius v. The vector ∆v is a chord of this circle. Transferring
the vector ∆v back to the left hand side of the figure shows that it
points toward the center of the circle.
Consequently, a particle moving at a constant speed in a circular
path is accelerating towards the center of the circle. We can determine
the magnitude of this acceleration by considering the hodograph again.
In the limit of infinitesimally small times, ∆θ → dθ, and ∆v → dv,
4In the rest of this section, the velocity v should be expressed as vT but I left
off the subscript to keep the notation simple. This is common practice and one
must distinguish between tangential and linear velocity from the context
42 2. KINEMATICS
and the chord approaches the subtended arc. From the basic relation
ds = rdθ, but applied to the hodograph, we see that in the limit, the
magnitude of ∆v is dv = vdθ. Therefore,
dv vdθ
a= = .
dt dt
But dθ = ds/r, so
v ds v2
a= = .
r dt r
That is, if a particle is moving in a circle at constant speed, it is
accelerated towards the center with a “centripetal” acceleration given
by
v2
ac = .
r
Since v = rω, this can be written in the equivalent form
ac = ω 2 r.
Note that a point in a rotating body can experience three different
types of acceleration:
(1) linear acceleration (if the body as a whole is accelerating),
(2) tangential acceleration (aT = rα; if the angular velocity of
the body is changing),
(3) centripetal acceleration (ac = v 2 /r = ω 2 r; due to the rota-
tion of the body).
y P
5
3 r
2
1
q
x
1 2 3 4 5
a sinh η a sin ζ
x= and y = .
cosh η − cos ζ cosh η − cos ζ
You will be happy to know that we will not be using this particular
set of coordinates! Fortunately, most coordinate transformations are
much simpler than this one.
z
^k
^i y
^j
x
ı̂ · ı̂ = 1, ̂ · ̂ = 1, k̂ · k̂ = 1,
ı̂ · ̂ = 0, ı̂ · k̂ = 0, ̂ · k̂ = 0.
These relations describe the orthogonality of the unit vectors. Similarly,
from the definition of the cross product, Equation (2.8), we find that
ı̂ × ı̂ = 0, ̂ × ̂ = 0, k̂ × k̂ = 0,
46 2. KINEMATICS
and
ı̂ × ̂ = k̂, ̂ × ı̂= −k̂
̂ × k̂ = ı̂, k̂ × ̂= −ı̂
k̂ × ı̂ = ̂, ı̂ × k̂= −̂
The last six relations are easily remembered with the help of the
mnemonic of Figure 2.5.
^ ^
i i
^
+ ^
j
^
k
- ^
j
k
Exercise 2.9. Two vectors are given by A = 2ı̂ +3̂ +4k̂ and
B = 3ı̂ − 2 ̂ Determine A + B, A − B, A · B and A × B.
Exercise 2.10. Calculate the angle between the vector V=2ı̂+4̂+6k̂
and the vector W=2ı̂+2̂+2k̂. Answer: 22.21◦
Exercise 2.11. The direction angles of a vector are the angles
between the vector and the x, y, and z axes. Determine the direction
angles of the vector r = 6ı̂+5̂−2k̂. Answer: α = 41.0◦ , β = 51.67◦ ,
γ = 104.3◦ .
Exercise 2.12. A vector forms the diagonal of a cube of side a.
Express it in terms of the unit vectors.
Exercise 2.13. Show that the cross product of two vectors A and
B can
be written as
the following determinant:
ı̂ ̂ k̂
Ax Ay Az
Bx By Bz
Exercise 2.14. Let A and B be the sides of a parallelogram. Show
that the area of the parallelogram is A×B.
2.6. KINEMATICS IN TWO DIMENSIONS 47
y ^ ^r
q
P
r
^
j q
^ x
i
The unit vectors r̂ and θ̂ are not constant because they are not
associated with a set of fixed axes. Rather, they are associated with
a vector r that can change with time. If r changes in magnitude, this
does not affect r̂ or θ̂ but if r changes in direction, then the direction
of the unit vectors r̂ and θ̂ will change. This means that r̂ and θ̂ are
functions of θ. In mathematical form we express this as
r̂ = r̂(θ),
and
θ̂ = θ̂(θ).
The position of a point in plane polar coordinates is given quite
simply by
r = rr̂.
^
j
^ ^
q r
q ^
i
Figure 2.7. The unit vectors r̂ and θ̂ and the unit cir-
cle. All the unit vectors originate at the center of the
circle and all have magnitude unity, so they end on the
circumference of the circle.
2.6. KINEMATICS IN TWO DIMENSIONS 49
If you take the derivative of this equation with respect to time, you
obtain an expression for the velocity in polar coordinates.
dr d dr dr̂
v= = (rr̂) = r̂ + r .
dt dt dt dt
The last term comes from the product rule of differentiation and reflects
the fact that the unit vector r̂ may be changing with time. Although
the magnitude of r̂ is always unity, it may change in direction. Since r̂
is a function of θ, from the calculus the differential of r̂ is
dr̂
dr̂ = dθ.
dθ
So,
dr̂ dr̂ dθ dr̂
= = θ̇.
dt dθ dt dθ
But what is dr̂/dθ? Looking at Figure 2.7 you see that r̂ and θ̂ can be
resolved into Cartesian components as follows:
r̂ = 1 cos θı̂ + 1 sin θ̂,
where I explicitly included the number 1 to emphasize that the radius of
the unit circle is 1. You obtain a similar expression for θ̂ by inspection
of Figure 2.7. Dropping the “1” we write:
r̂ = cos θı̂ + sin θ̂, (2.10)
θ̂ = − sin θı̂ + cos θ̂. (2.11)
Therefore,
dr̂ d
= (cos θı̂ + sin θ̂) = − sin θı̂ + cos θ̂ = θ̂.
dθ dθ
Furthermore,
dθ̂
= − cos θı̂ − sin θ̂ = −r̂.
dθ
You may wish to memorize the following relations, as you will be using
them frequently in this and other courses.
dr̂ dθ̂
= θ̂ and = −r̂. (2.12)
dθ dθ
Having determined dr̂/dθ and dθ̂/dθ we can obtain dr̂/dt and dθ̂/dt
as follows:
dr̂ dr̂ dθ̂
= = θ̇θ̂,
dt dθ dt
and
dθ̂ dθ̂ dθ
= = −θ̇r̂.
dt dθ dt
50 2. KINEMATICS
z
P
^
zk
r
y
x φ ρ
The next task is to determine unit vectors for the cylindrical coordi-
nate system. This is quite simple because the new coordinate system is
made up of old coordinates whose properties you have already learned.
You can appreciate from Figure 2.9 that ρ̂ and φ̂ are the same as the
plane polar unit vectors r̂ and θ̂, and the unit vector in the z direction
is just k̂ as in the Cartesian coordinate system.
Although k̂ is constant in magnitude and direction, ρ̂ and φ̂ will
change in direction as the angle φ varies. That is, ρ̂ and φ̂ are functions
of φ. We write:
ρ̂ = ρ̂(φ),
and
φ̂ = φ̂(φ).
The set ρ̂, φ̂, k̂ form an orthogonal set. The orthogonality of these
unit vectors leads the following set of easily proved properties:
2.7. KINEMATICS IN THREE DIMENSIONS 53
ρ̂ · ρ̂ = 1, φ̂ · φ̂ = 1, k̂ · k̂ = 1
ρ̂ · φ̂ = 0, ρ̂ · k̂ = 0, φ̂ · k̂ =0
ρ̂ × φ̂ = k̂, φ̂ × k̂ = ρ̂, k̂ × ρ̂ = φ̂
The relationships between cylindrical coordinates and Cartesian co-
ordinates (the transformation equations) are:
x = ρ cos φ y = ρ sin φ z = z.
The inverse relationships are:
−1 y
p
ρ= x +y 2 2 φ = tan z = z.
x
The unit vectors ρ̂ and φ̂ can be expressed in terms of ı̂ and ̂ as
follows:
dρ̂ dφ
v = ṙ = d
dt
(ρρ̂ + z k̂) = ρ̇ρ̂ + ρ dρ̂
dt
+ ż k̂ =ρ̇ρ̂ + ρ( dφ dt
) + ż k̂,
or
v =ρ̇ρ̂ + ρφ̇φ̂ + ż k̂. (2.18)
Similarly, the acceleration is given by
a = r̈ =(ρ̈ − ρφ̇2 )ρ̂ + (ρφ̈ + 2ρ̇φ̇)φ̂ + z̈ k̂. (2.19)
r
y
^ ^
ρ k
x φ
φ
^
ρ
= −aω 2 ρ̂ + 2bk̂
z
^
r
z =r cos θ ^
φ
θr
^ ^
θ
k y =r sinθ sinφ
x = r sinθ cosϕ ^
ρ y
φ
ρ=rsinθ
x
Figure 2.10. Spherical coordinates and the unit vec-
tors r̂, θ̂, φ̂. Note that the Cartesian components of the
vector r are x = r sin θ cos φ, y = r sin θ sin φ and
z = r cos θ. The unit vectors r̂, θ̂, φ̂ point in the direc-
tion of increasing r,θ, φ, respectively. Also shown are the
cylindrical unit vectors ρ̂ and k̂.
What are the appropriate unit vectors? The unit vector r̂ is in the
direction of increasing r. The unit vector θ̂ is in the direction that the
tip of r will move if θ increases. In Figure 2.11 I sketched a quarter
circle to illustrate the motion of r as θ changes. The unit vector θ̂ is
tangent to this circle and has the direction of increasing θ. The unit
vector φ̂ lies in the x-y plane and gives the direction that the tip of
ρ̂ will move if φ increases. By sliding φ̂ down to the x-y plane, you
can see that φ̂ in spherical coordinates is the same as the φ̂ of the
cylindrical coordinates.
Although it may not be immediately obvious, the spherical unit
vectors r̂, θ̂, φ̂ form an orthogonal set. Figure 2.10 shows these three
mutually perpendicular unit vectors at the tip of the vector r. This is
56 2. KINEMATICS
φ
ρ
Recalling that ρ and φ are the same as the plane polar coordinates
(formerly denoted r and θ), and using Equation (2.25),
ρ̂ = cos φı̂ + sin φ̂.
Consequently,
r̂ = sin θ cos φı̂ + sin θ sin φ̂ + cos θk̂. (2.22)
2.7. KINEMATICS IN THREE DIMENSIONS 57
Figure 2.12. The unit vectors r̂, θ̂, φ̂ for spherical co-
ordinates and the unit circle. Also shown are k̂ and ρ̂.
In the sketch on the right, θ̂ has been translated from
the tip of r̂ to the origin.
dr̂
= θ̇θ̂ + sin θφ̇φ̂.
dt
2.7. KINEMATICS IN THREE DIMENSIONS 59
2.8. Summary
In this chapter you learned how to express the position, velocity,
and acceleration of a particle in two dimensions in Cartesian coordi-
nates and plane polar coordinates, and in three dimensions in Cartesian
coordinates, cylindrical coordinates and spherical coordinates.
In three dimensions the position, velocity, and acceleration in Carte-
sian coordinates are given by the simple relations:
r = xı̂ + ŷ + z k̂,
v = ẋı̂ + ẏ̂ + ż k̂,
a = ẍı̂ + ÿ̂ + z̈ k̂.
2.9. PROBLEMS 61
2.9. Problems
Problem 2.1. A particle initially at rest undergoes an acceleration
given by a = 3e−0.5t m/s2 . Determine the terminal velocity.
Problem 2.2. The nitrogen atom in an ammonia molecule can be
assumed to oscillate in simple harmonic motion, so that its position
at any time t is given by z = A cos ωt where A and ω are constants.
Obtain expressions for its velocity and acceleration as functions of time.
Problem 2.3. This is the story of Freddy Physics who got a ticket
for going through a red light. He explained to the judge that if he
slammed on the brakes when he saw the signal light turn yellow, he
would not be able to stop before the intersection, and if he continued
at a constant speed the light would turn red before he reached the
intersection. In this problem you will determine how far from the light
Freddy was when the light turned yellow. Suppose he is traveling at
the speed limit, v0 , that the light remains yellow for a time t0 and
that Freddy’s reaction time is τ . His car decelerates at a rate a. (a)
Determine the distance d1 he will travel in time t0 if he continues at the
speed limit (b) Determine the distance d2 required to stop. (c) Show
that he is bound to go through the red light if t0 < τ + v0 /2|a|
62 2. KINEMATICS
2.5 km 1.5 km
If the orbital radius increases by a small amount ∆r, the period will
increase by ∆τ. Show that
∆τ 3 ∆r
= .
τ 2 r
Problem 2.24. Prove the triangle inequality which states that for
two vectors A and B, |A + B| ≤ |A| + |B| .
Problem 2.25. Show that A · (B × C) is the volume of a paral-
lelepiped whose edges can be represented by the vectors A, B, C.
Problem 2.26. Evaluate (A × B) · C if A = 3ı̂+2̂+4k̂, B = 3ı̂+
2k̂ and C = 4k̂. Also show that you get the same result by evaluating
A · (B × C).
Problem 2.27. Show that
Ax Ay Az
(A × B) · C = Bx By Bz .
Cx Cy Cz
Problem 2.28. Prove the “BAC minus CAB” rule:
A × (B × C) = B(A · C) − C(A · B).
Problem 2.29. Derive the law of cosines for triangles, i.e., show
that if the sides of a triangle have lengths a, b, c, and the angle between
a and b is θ, then
c2 = a2 + b2 − 2ab cos θ.
Problem 2.30. The position of a particle is given by r = 2t (me-
ters), θ = 5t (radians). (a) Plot the path of the particle. (b) Obtain
an expression for the speed as a function of time.
Problem 2.31. Suppose the position of a particle as a function of
time is given by r = 1 + sin t and θ = 1 − e−t . Obtain v and a in terms
of r̂ and θ̂.
Problem 2.32. A particle moves in a plane such that r = 1 − cos θ
m/s and θ̇ = 4 rad/s. Determine v and a.
Problem 2.33. A particle moves in a plane in such a way√ that
r = 2 + sin t (m) and the magnitude of the velocity is v = 2 cos t
(m/s). Find a formula for θ = θ(t).
Problem 2.34. For plane polar coordinates, determine
d3 r
.
dt3
66 2. KINEMATICS
This chapter (and much of the rest of this book) deals with dy-
namics, that is, the relation between the forces acting on a body and
its motion. A force is an interaction between a body (or particle) and
its environment, usually described as a push or a pull in a specified
direction. In this book you will encounter a number of familiar forces,
such as the gravitational force and forces exerted by springs as well as
a few less familiar ones.
On a very basic level, all forces are manifestations of the “funda-
mental forces,” namely, the gravitational force, the nuclear (or strong)
force, and the electroweak force. The electroweak force is often thought
of as two different forces, the electromagnetic force and the weak nu-
clear force. The electromagnetic force, in turn, is often thought of as
the electric force and the magnetic force.1 All the known forces in
nature are ultimately related to these fundamental forces. For exam-
ple, the force exerted by your muscles can be traced back to electrical
forces.
Dynamics is neatly summarized by Newton’s three laws. You stud-
ied Newton’s laws in your introductory mechanics course. By this time,
you surely know the laws by heart and you know how to use them to
solve reasonably complicated physics problems. However, it is impor-
tant for you as a physicist to have a thorough understanding of these
fundamental statements about the nature of the physical universe. In
1To learn more about the fundamental forces you might read the interesting
article by Charles Seife, “Can the Laws of Physics be Unified?” Science, 309, 82
(2005).
69
70 3. NEWTON’S LAWS: DETERMINING THE MOTION
this chapter I will discuss some aspects of Newton’s Laws that you
may not have considered before. I hope this will give you a greater
appreciation for the scope and significance of the laws.
According to legend, the professor quit his job and left the position
to Newton.
Newton did not formulate his laws out of the blue. He was very fa-
miliar with the work of his predecessors, especially Galileo and Kepler.
(As he stated, “If I have seen further than others, it is by standing on
the shoulders of giants.”) In fact, Newton’s first law, the law of iner-
tia, was formulated and demonstrated by Galileo. It is possible that
Galileo got the idea from Rene Descartes. However, the second and
third laws were first formulated by Isaac Newton.3
Although Newton discovered the law of universal gravitation when
he was very young, he did not publish it until many years later because
his early calculations did not give the correct value for the period of
the Moon. Urged by his friend, the architect Sir Christopher Wren,
Newton carried out the calculations more carefully, obtained the right
answer, and convinced himself that his law was correct.
Newton invented the reflecting telescope and carried out a large
number of experimental investigations, particularly in optics. He was
also interested in alchemy and theology. But his greatest contribution
was in formulating physics as an exact mathematical science. His ge-
nius was immediately recognized and he soon became one of the most
famous men in Europe. After Newton, people realized that natural
processes take place in a manner that can be analyzed and predicted.
People began to think of the universe as a “clockwork” mechanism in
which the future development of any system could (in principle) be
determined from a knowledge of its present state. The poet Alexander
Pope, eulogized Newton with the couplet,
Nature and Nature’s Laws lay hid in Night,
God said, Let Newton be! and All was Light.
Personally, Newton was not easy to deal with. It was said of him
that, “He suffered people poorly, and fools not at all.” Since his position
at Cambridge was a semi-clerical post, he was not allowed to marry. It
appears that the life of a bachelor suited him well. He died in 1727 at
the ripe old age of eighty four.4
if you are riding in a bus and someone places a tennis ball in the aisle,
it will remain at rest in the aisle as long as the bus is moving at a con-
stant velocity. But if the bus accelerates the ball will roll. If the bus
slows down, the ball rolls forward. If the bus turns left, the ball rolls
to the right. If I were somebody who did not know very much physics,
I might conclude that the ball does not obey Newton’s first law. You
realize, of course, that I am describing the motion of the ball relative
to the bus and the bus is, as I said, accelerating. So you would refute
my statement by pointing out that relative to the ground the ball did
obey the first law. It did maintain a state of uniform motion in the
reference frame of the Earth. I might then point out that the Earth is
also an accelerating reference frame because it is rotating. How would
you respond?
If we follow such arguments to their logical conclusion we will decide
that the first law is only valid in a non-accelerating reference frame.
(This is also true of Newton’s other two laws.) A non-accelerating
frame is called an inertial reference frame. Does such a reference frame
actually exist? Newton stated that his laws were valid in a reference
frame that was “at rest with respect to the fixed stars.” But we know
the stars are all in motion! It is probably best to treat the concept
of an inertial reference frame as a useful idealization. For many prob-
lems, the Earth can be treated as if it were at rest. This approximation
breaks down when the rotation of the Earth must be considered. Then
we usually assume that the inertial reference frame is a nonrotating
reference frame with origin at the center of the Earth. That approx-
imation breaks down if we need to include the orbital motion of the
Earth around the Sun. If necessary, we can take the origin of coordi-
nates to be at the center of the Sun, or even the center of the galaxy.
dp
= F, (3.1)
dt
where F is the net or total force acting on the body. (You may prefer
to write it as Fnet or as ΣF.) Now momentum is defined as p =mv, so
if the mass is constant, Newton’s second law takes on the familiar form
F = ma.
Keep in mind that Newton’s second law is applicable only in an inertial
reference frame.
Newton’s second law expressed in the form a = F/m is sometimes
referred to as the “equation of motion.” The reason is that if the forces
are known, we can determine the acceleration of the body, and once
the acceleration has been determined we can integrate to obtain the
“motion,” that is, the velocity and position as functions of time. Ac-
cording to classical mechanics, a knowledge of the position and velocity
of all the particles in a system allow one to determine the forces and
hence to predict the future development of the system.6 Therefore, an
6This is a wonderful fact of nature. As stated by Landau and Lifshitz on the
very first page of their excellent advanced mechanics book, “If all the coordinates
and velocities are simultaneously specified, it is known from experience that the
state of the system is completely determined and that its subsequent motion can,
in principle, be calculated.” (Emphasis added.) They are suggesting that there is
no fundamental reason why the motion of a system depends only on position, ve-
locity and time; we must accept it as an experimental fact. (Reference: Mechanics,
Course in Theoretical Physics, Volume 1, L. D. Landau and E. M. Lifshitz, Perg-
amon Press, New York, 1976. Page 1.) Interestingly, in electrodynamics there is a
quantity called the “Abraham-Lorentz force” that is usually expressed as a function
of acceleration. This force leads to some strange behavior that has been referred to
as “philosophically repugnant” by David Griffiths (Introduction to Electrodynamics,
3rd Ed, Prentice Hall, 1999, page 467).
3.3. NEWTON’S SECOND LAW AND THE EQUATION OF MOTION 75
that mG and mI are equal to within one part in 20 million. This leads
one to conclude that gravitational mass and inertial mass are indeed
the same thing. The fact that inertial mass and gravitational mass are
equal is called “The Principle of Equivalence.” Albert Einstein used
this equivalence as a basic postulate of his theory of General Relativity
(1915).7
Newton’s second law is a cornerstone of classical physics because
it can be used to calculate the acceleration of a body, given the force
acting on it. Once the acceleration of a body is known, the laws of
kinematics determine its velocity and position at any later time. This
means that if you know the net force acting on a body, you can calculate
its position at any future time. The ability to predict the motion is the
power of the second law.
Recall again that dynamics is the study of how a force affects the
motion of a body. As you might suspect, dynamics usually involves
accelerating bodies. However, in some situations there may be forces
acting on a body but nevertheless the acceleration is zero. Consider,
for example, a body acted upon by two equal and opposing forces. The
effects of these forces cancel out and the body does not accelerate. Zero
acceleration and zero velocity is an important special case in dynamics
and is called statics. Statics is of particular interest to civil engineers
who want to make sure the structures they design, such as bridges and
skyscrapers, will have zero acceleration. Statics was treated briefly in
Section 1.5.2 and will be dealt with in greater detail in Chapter 14.
The principle of superposition states that if two or more forces act
on a particle, the net effect is that due to a single force equal to the
vector sum of all the forces. You will be exposed to the principle of
superposition in other areas of physics. For example, the net electric
field at a point is the vector sum of all the electric fields acting at that
point.
In our discussion of Newton’s laws we considered interactions be-
tween particles. When we apply Newton’s second law to an extended
body the acceleration a refers to the acceleration of the center of mass
of the body. A net external force F applied at any point of an extended
rigid body will cause the center of mass of the body to accelerate ac-
cording to a = F/m. This is often not at all intuitive. For example,
consider a tricycle with the pedals in a vertical position as shown in
Figure 3.1. If you pull forward on the top pedal, the bicycle will ob-
viously move forward. But what happens if you pull forward on the
bottom pedal? It might seem that this would propel the tricycle back-
ward, but Newton’s second law tells us that if we pull forward, the
tricycle moves forward. (If you do not believe me, try it! You will be
surprised by the motion of the pedal itself. This is not at all the same
as what would happen if you were riding the tricycle. In that case, the
force your foot exerts on the pedal is an internal force.)
Therefore,
d
(p1 + p2 ) = 0.
dt
That is, the total momentum of an isolated system is constant.
Is the third law always obeyed? This is a question that physicists
have debated for many years. We can state unequivocally that the third
law is always obeyed in purely mechanical systems, but the situation
is not quite as clear when we consider the electromagnetic interaction
between charged particles, as illustrated by the following example.
conclude that the third law has the same general range of validity
as the first two.b
aThis relation is based on the Biot-Savart law which, strictly speaking, is not
valid for point charges. However, the correct expressions reduce to the given
formula as long as the velocity of the particle is much less than the speed of
light. (D. J. Griffiths, Introduction to Electrodynamics, 3rd Ed., Prentice Hall,
1999, p. 439.
bRoald K. Wangsness, Electromagnetic Fields, 2nd Ed., Wiley and Sons, New
York, 1986, pages 219 and 359. For a different point of view, see “Classical
Dynamics” by Jerry Marion and Stephen Thornton, 3rd Ed., Harcourt, Brace,
Jovanovich, 1988, page 45.
q1 q2
r12
+ +
v1
v
2
but they act on different bodies. Similarly, for the cart and donkey
problem, the donkey exerts a force on the cart and the cart exerts an
equal and opposite force on the donkey. Considering only the cart, the
forces on the cart are the force exerted by the donkey and frictional
forces (the road on the wheels, etc.). If the force exerted by the donkey
is greater than the frictional force, the cart accelerates forward. You
should also think about the forces acting on the donkey and explain
why the donkey is accelerating forward and not backward. Note that
the only forces that can make an object accelerate are external forces.
You cannot pick yourself up by pulling on your shoestrings!
In determining action-reaction pairs, it is useful to remember that
they are always the same kind of force. For example, a book on a
tabletop is pulled downwards by the gravitational attraction of the
Earth. It is easy to make the error that the reaction is the upward
normal force of the table. But this is not the same kind of force. The
reaction is the upward gravitational pull of the book on the Earth.
What is the reaction force to the normal force exerted by the table?
Now in the initial and final stages, the water was not moving relative
to the bucket, but the surface was flat in the initial stage and concave
in the final stage. Therefore, the behavior of the surface was not due to
the motion of the water relative to the bucket. Also, when the bucket
was rotating but the water was not rotating, the surface remained
flat. Newton concluded that the concavity of the surface was due to
the absolute rotation of the water and not its motion relative to the
bucket.
If you placed the bucket of water in a closed box and set the box
rotating, a TV camera inside the box would show the surface of the
water to be concave. Thus there is a physical phenomenon that can be
used to determine whether or not a reference frame is rotating. There
is no other reference frame (Newton decided) in which the surface of
the water is flat, so rotation is not relative.8
Ernst Mach was fascinated by this simple experiment. However,
he believed that all motion was relative, including rotational motion.
Mach claimed that rotation was relative to all the mass in the universe.
He reasoned that a bucket of water at rest would have a flat surface,
just as Newton observed. But, he asked, would the surface be curved
if the bucket were spun in a totally empty universe? In other words,
what would the bucket be spinning relative to? From Mach’s point
of view, it is the rest of the universe that causes the surface to be
curved, and the same effect would be obtained by spinning the entire
universe around a stationary bucket. You cannot determine whether
the water is rotating and the universe is at rest, or the water is at
rest and the whole universe is rotating around it! Einstein’s theory of
general relativity does not include this “total relativity” of Mach, but
later in his life Einstein tried to analyze the consequences of imposing
it on his theory.
Although the question of whether or not rotation is relative or abso-
lute is still being debated,9 we shall go along with Newton and assume
that rotation is absolute. This point of view is supported by a re-
cent in-depth study10 of Mach’s principle that shows there are many
effects that cannot be explained by a rotating universe and a stationary
8Newton stated that centrifugal force is the feature that distinguishes absolute
motion from relative motion.
9An interesting article on this subject is “Total Relativity: Mach 2004” by
Frank Wilczek, Physics Today, April 2004, page 10. Newton’s bucket is also used
as a starting point for the very readable book, The Fabric of the Cosmos: Space,
Time and the Texture of Reality by Brian Greene, Knopf, New York, 2004.
10H. Hartman and C. Nissim-Sabat, “On Mach’s critique of Newton and Coper-
nicus,” Am. J. Physics, 71, 1163 (2003).
3.6. DETERMINING THE MOTION 83
bucket. For example, Mach’s principle cannot explain how two buck-
ets, rotating in opposite directions, can both have concave surfaces.
Finally, we might note that in Quantum Mechanics we find that rota-
tional motion is quantized whereas linear motion is not. A body can
have any linear velocity whatever, but a body can only have certain
discrete values for rotational motion.
equations are valid only for a constant force. The rest of this chapter
deals with forces that are not constant.
F (t)
a= .
m
or
dv 1
= F (t).
dt m
Separating variables and integrating:
Z v
1 t
Z
dv = F (t)dt.
v0 m 0
In writing this last equation I was careless and used the same symbol (v)
for a variable of integration and a limit of integration. (This is the kind
of thing that drives mathematicians crazy!) You can usually get away
with this sloppy notation, but sometimes it can get you into trouble.
So I will rewrite the equation above and use double primes for the
variables of integration and single primes for the limits of integration.
Z v0 =v(t0 ) Z 0
00 1 t
dv = F (t00 )dt00 .
v0 m 0
So,
Z t "Z t0
#
1
x(t) = x0 + v0 t + F (t00 )dt00 dt0 .
m 0 0
Although this expression looks rather forbidding, the procedure for
1
R
finding x = x(t) is not at all difficult: you simply integrate Rm F (t)dt
to get the velocity as a function of time and then integrate v(t)dt to
obtain the position as a function of time (and don’t worry about all
those primes and double primes).
with zero velocity. Determine its position at time t=5 s. Assume the
mass of the particle is 0.1 kg. Answer: 742.5 m.
Exercise 3.6. A certain electromagnetic wave has its electric field
oriented along the z-axis. The magnitude of the field varies in time
according to E = E0 cos ωt. Initially, an electron is at rest at the
origin. Determine the motion of the electron. (Recall that the force
exerted on a charged body in an electric field is given by F = qE.)
Answer: x = −(qE0 /ω 2 m)(cos ωt − 1).
and
v(t) bt
ln v|v(t)
v0 = ln =− ,
v0 m
or
v(t) = v0 e−bt/m .
Thus we have obtained an expression for the velocity at any given
time; half of our job is finished. Integrating again we will obtain
the position as a function of time. Starting with the definition of
velocity, v = dx
dt
, we write
dx
= v(t) = v0 e−bt/m .
dt
Consequently,
Z x(t) Z t
dx = v0 e−bt/m dt.
x0 0
Integrating,
h vm it v0 m −bt/m
0 −bt/m
x(t) − x0 = − e =− (e − e0 ),
b 0 b
or
v0 m
x(t) = x0 + (1 − e−bt/m ).
b
It is interesting to consider the limiting value of our result by letting
t → ∞.13 This leads to the rather un-intuitive result for the velocity of
the canoe because it indicates that the velocity of the canoe [v0 e−bt/m ]
does not go to zero until t → ∞. The canoe is in motion for an infinite
time! How far do you suppose it will travel in this infinite amount
of time? To answer this question consider the expression for position
and set t = ∞ to obtain x(t = ∞) = x0 + v0 b/m. Note that this is
a finite distance. So, even though it takes an infinitely long time for
the canoe to stop, it only goes a finite distance. Does this result make
sense? Yes, it does. The retarding force gets smaller and smaller as
the velocity decreases. Perhaps you might complain that the problem
is poorly posed and it should ask for the time required for the velocity
to fall below some particular value. For example, once the canoe is
moving at some very small velocity, such as one inch per hour, then for
all intents and purposes it is stopped.
If an object is moving in air at less than about 20 m/s (≈ 45 mph),
it is safe to assume the retarding force is proportional to the first power
of the speed. For objects moving at higher speeds (but less than the
13It
is a good idea to subject your answers to a “sanity check” by seeing how
they behave in limiting cases such as t = 0 and t = ∞.
3.6. DETERMINING THE MOTION 89
or
dv b gm b
= ( − v 2 ) = (vT2 − v 2 )
dt m b m
Therefore,
Z v(t)
b t
Z
dv
2 2
= dt
v0 =0 vT − v m 0
which yields
1 v b
tanh−1 = t
vT vT m
and consequently,
r
vT bt bg
v(t) = vT tanh = vT tanh t.
m m
The position is given by
Z t Z t r
bg
x(t) − x0 = − vdt = vT tanh tdt,
0 0 m
where the minus sign is due to the decrease in height of the object
with time. A udu substitution leads to
Z √ gb
m u= m t
x(t) − x0 = − tanh udu
b u=0
and finally
m gb
x(t) = x0 − log cosh t.
b m
Exercise 3.7. A body is dropped from rest from a hot air balloon.
Make the unrealistic assumption that the resistive force is proportional
at all times to the velocity, i.e., fR = −bv. (a) What is its terminal
velocity vT ? (b)How much time is required for the body to reach a
velocity of 0.9 vT ? Answer: (b) 2.3m/b.
Exercise 3.8. A body is falling under the effects of gravity and a
retarding force proportional to the
p square of the speed. Determine its
terminal velocity. Answer: vT = gm/b
k
m
or r
x
k t
Z Z
dx
p = dt.
x0 mC/k − x2 m 0
The integral on the left can be found in any table of integrals,
yielding x r t
x k
sin−1 p = t .
mC/k m
x0 0
Therefore
r
−1 x −1 x0 k
sin p − sin p = t.
mC/k mC/k m
The second term on the left is just a constant; call it β. Then
r
−1 x k
sin p = t + β,
mC/k m
or r r !
mC k
x= sin t+β .
k m
This is usually written in the easily remembered form
x = A sin (ωt + β) . (3.9)
The constant r r
mC mv02 + kx20
A= =
k k
is the amplitude. The quantity
p
ω = k/m
is the (angular) frequency of the oscillation. The constant β is
called the “phase constant” and is related to the position of the
oscillator at time t = 0. By adding π/2 to β one can express x(t)
in terms of the cosine rather than the sine.
v = v(t, v0 , x0 ),
x = x(t, v0 , x0 ).
These give the position and velocity as functions of time and the ini-
tial conditions, v0 and x0 . If you are lucky, these expressions are in
closed form, that is, in terms of familiar functions such as logarithms,
exponentials, trigonometric functions, etc. (You are comfortable with
these so-called “elementary functions,” but if you were confronted with
an expression involving the gamma function or an elliptic integral, you
might not feel so happy.)
If v(t) and x(t) are given in closed form, you have a great deal
of knowledge about the motion. First of all, you can determine the
velocity and position at any time by simply plugging the time into
the equation. Furthermore, you can simply look at the equation and
get a very good idea about the motion. For example, if you obtain
the solution x = A cos ωt you immediately know that the motion is
oscillatory with amplitude A and angular frequency ω. Generally, it is
easy to manipulate the solution to obtain more information about the
system. For example, squaring the expression for the velocity to get v 2
immediately leads to an equation for the kinetic energy as a function
of time. If the motion is two-dimensional and you obtain solutions x(t)
and y(t) you can easily combine them to get an equation for the path
(or orbit) of the body.
96 3. NEWTON’S LAWS: DETERMINING THE MOTION
If you know x and y and can calculate f (x, y) you can use this rela-
tionship to determine x for a slightly larger value of y, namely, y + ∆y.
Note that the relationship is not exact. However, the approximation
becomes better and better as ∆y becomes smaller and smaller.
Let us apply this technique to the equation of motion. We want
to solve for x(t) numerically, given an initial value x0 and an explicit
expression for the acceleration, a = a(x, v, t). The procedure involves
carrying out the following sequence of operations:
1. Select ∆t to be a small time step.
2. Set x = x0 and t = t0 .
3. Determine a = a(x, v, t).
4. Obtain new time from t = t + ∆t.
5. Obtain new velocity from v = v + a · ∆t.
6. Obtain new position from x = x + v · ∆t.
7. Go to Step 3.
You go through this procedure repeatedly. At each step you obtain
new values for x, v, and t.
Consider a simple but important numerical technique called the
Euler method. This is a technique for obtaining the numerical solution
of an equation of the form
d2 x
= f (x, v, t).
dt2
If the function on the right-hand side is the force divided by the mass,
this is just Newton’s second law.
Write the second-order differential equation above as two first-order
equations. These are,
dx
= v(x, t),
dt
dv
= f (x, v, t).
dt
98 3. NEWTON’S LAWS: DETERMINING THE MOTION
Solve as follows: let x1 and v1 be the initial values of x and v. Let the
initial time be t1 and let τ be a small time step. Then
x2 = x1 + v1 τ
v2 = v1 + f (x1 , v1 , t1 )τ
The next step is to replace x1 by x2 and v1 by v2 and let t1 be replaced
by t2 = t1 + τ. Then evaluate
x3 = x2 + v2 τ,
v3 = v2 + f (x2 , v2 , t2 )τ.
You can appreciate that, in general, the technique consists in replacing
xn by xn+1 and vn by vn+1 and repeating as many times as desired.
This is the essence of the Euler method.
Some years ago, Alan Cromer16 noted that the Euler method, as
expressed above, is unstable. That is, the answer diverges further and
further from the correct solution. However, Cromer discovered that the
equations could be made stable by a simple but crucial change, namely
by calculating v first and using the new value of v to determine x. The
so-called “Euler-Cromer” algorithm is
v2 = v1 + f (x1 , v1 , t1 ) · τ,
x2 = x1 + v2 τ.
This simple scheme allows you to numerically determine the velocity
and position of a particle at any future time if you know its acceleration
f (x, v, t) at a prior time.
When carrying out a numerical integration, you should check fre-
quently to determine if the solution is stable. For example, you might
check at each time step to make sure the energy or angular momentum
is conserved.
3.8. Summary
A number of important concepts were presented in this chapter. Ba-
sically, you learned (or reviewed) Newton’s three laws and you learned
how to determine the motion for a mechanical system (assuming you
know the force).
Newton’s three laws are:
(1) The law of Inertia. (A body tends to preserve its state of
motion.)
16Alan Cromer, “Stable solutions using the Euler approximation,” American
Journal of Physics, 49, 455 (1981). In this paper Cromer explains why one algo-
rithm is stable and the other is not.
3.9. PROBLEMS 99
3.9. Problems
Problem 3.1. A particle of mass 3 kg is acted upon by the two
forces (in newtons)
F1 = 6ı̂,
and
F2 = 3ı̂ + 3̂.
Determine the magnitude and direction of the acceleration.
Problem 3.2. A pail is filled with oil to a depth of 10 cm. A
steel marble of mass 0.2 kg is released from rest 50 cm above the top
surface. Assuming the oil exerts a constant resistive force of 2.4 N on
the marble, determine the speed with which it reaches the bottom of
the pail. Answer: 3.06 m/s
100 3. NEWTON’S LAWS: DETERMINING THE MOTION
these three cases. (b) What is the force that is causing the rocket to
accelerate? (c) What is the force on the donkey that keeps it in motion?
Problem 3.15. The Aristotelians noticed that if you push a box
across the room, it moves with a constant speed, and if you push it
harder the speed is greater, but as soon as you quit pushing, the box
stops. This suggests that the equation of motion should be F = mv.
Furthermore, the Aristotelians claimed that heavier objects fall to the
ground faster than light objects. (Thus, an object four times heavier
than a light object will fall to the ground in one fourth the time.)
(a) Show that these assumptions lead to the absurd conclusion that
the Earth exerts the same gravitational force on all bodies, regardless
of their mass. (b) Describe a simple experiment to show that the
Earth does not exert the same gravitational attraction on all bodies,
regardless of their mass.
Problem 3.16. Two carts of masses M and 2M have springs at-
tached to either end as shown in Figure 3.5. Cart 1 has mass M and
cart 2 has mass 2M. They are on a frictionless segment of track of
length L with barriers at the ends. The two carts are brought together
at the center of the track, the springs in contact are compressed, and
the carts are released from rest. Cart 1 moves left, hits the barrier at
x = 0 and bounces back. Cart 2 moves right, hits the barrier at x = L
and bounces back. The collisions are completely elastic. Where do the
carts meet? The dimensions of the carts are much smaller than the
length of track, so the carts can be treated as particles. (b) Where do
the carts meet if cart 2 is four times more massive than cart 1?
1 2
0 L/2 L
where b and v0 are constants. (a) Find the force acting on the particle,
F (x). (b) Determine x(t). (c) Obtain an expression for the force as a
function of time, F (t).
Problem 3.18. A certain object is losing mass at a rate κ (kg/sec).
It is acted upon by a constant force F. (a) Determine its acceleration
and its position as a function of time. Assume that at t = 0, the mass
is m0 , the velocity is zero and the object is at x = 0.
(b)
Show that
1 F
as κ → 0, the expression for x(t) reduces to x = 2 m0 t2 , that is,
x = 21 at2.
Problem 3.19. An object of mass 2 kg is thrown upwards with an
initial velocity of 4 m/s. It is acted upon by the constant gravitational
force of the Earth and a mysterious force of 2 newtons that always
opposes the motion of the object. (This is like the force of air resistance,
except we are assuming the force is constant.) Determine the position
of the object at time t = 0.5 s. (Be careful; by this time the object is
falling downwards and the mysterious force is acting upwards!)
Problem 3.20. A builder plans to use a chain of length 1.5 meters
and linear density 7 kg/m to suspend a 220 kg mass from an overhead
beam. Determine the forces at either end and at the middle of the
chain. Obtain an equation for the force on a link at any point of the
chain. If the maximum tension that can be supported by a link is 2200
N, where will the chain break?
Problem 3.21. A bathroom scale is placed in an elevator. A man
weighing 180 lbs stands on the scale and observes that as the elevator
ascends, the scale reads 200 lbs while the elevator is accelerating and
140 lbs when it is slowing to a stop. What are the acceleration and the
deceleration of the elevator? (g=32.2 ft/s2 )
Problem 3.22. A bullet initially traveling at 400 m/s passes through
a 2 cm thick slab of wood and emerges at speed vf . When fired into a
thick block of the same wood, the bullet is found to penetrate 10 cm.
Determine vf .
Problem 3.23. A physics student designed an apparatus for her
experimental physics class. The apparatus consisted of a block m =
0.25 kg mass attached to a compressed spring of constant k mounted on
a cart of mass M = 2 kg, as shown in Figure 3.6. The car is pulled by a
string that passes over a pulley to a 10 kg hanging weight. The student
used photogate timers to determine the acceleration of the mass on the
spring when it was 15 cm from its equilibrium position. She found this
104 3. NEWTON’S LAWS: DETERMINING THE MOTION
m
M=2kg
10
zero, how much time does it take for the skydiver to reach a speed of
90% of the terminal velocity?
Problem 3.34. A bullet of mass m is fired vertically upward with
an initial velocity v0 . Assume the resistive force of the air is given by
mkv where m is the mass, v is the velocity and k is a constant. How
much time is required for the bullet to reach the top of its path?
Problem 3.35. A physics student on the roof of a tall building
drops a ball of diameter 10 cm and mass 0.3 kg. The air density is 1.2
kg/m3 . (a) What is the terminal speed? (b) Obtain an expression for
the speed of the ball as a function of how far it has fallen. (c) If the
roof is 100 meters above the ground, what is the speed of the ball when
it hits the sidewalk?
Problem 3.36. An object is dropped from an airplane that is flying
at an altitude of 6000 m at a speed of 1000 km/hr. Determine the
horizontal distance traveled by the object by the time it hits the ground.
For the sake of making the calculations somewhat simpler, you may
assume that the resistive force of the air is proportional to the first
power of the velocity and that the terminal velocity of the object is 98
m/s. (Justify neglecting small terms.)
Problem 3.37. An object of mass m in a uniform gravitational
field is dropped from some initial height. Assume that the force of air
resistance is proportional to the first power of the velocity (∝ bv). Show
that the distance y it has fallen in time t is
mg m −bt/m
y= t+ e −1 .
b b
Problem 3.38. Assume the force of air resistance is given by Fair =
2
Dv where v is the velocity and D is a constant equal to 0.01 kg/m.
An object of mass 2 kg initially at rest, is dropped from a height of
1000 m. The gravitational force can be assumed constant. Determine
how far the object has fallen in 20 s.
Problem 3.39. An object of mass 10 kg is dropped from a high
place. Assume air resistance is proportional to the velocity squared
(= Dv 2 ) where D = 0.01 Ns2 /m2 . (a) Evaluate the terminal velocity.
(b) Determine the time required for the ball to reach 0.9vT .
Problem 3.40. A bead on a long straight wire is repelled from the
origin x = 0 by a force proportional to its distance to that point. The
bead is initially at rest at position xo . Determine its motion.
Problem 3.41. A particle of mass m kg is at rest at x0 . It is
subjected to a force of magnitude F = k/x where k = 4 Nm. the
3.9. PROBLEMS 107
force is directed along the positive x-axis. Determine the speed of the
particle when it passes through the point x = 2x0 .
Problem 3.42. A block of mass m on a frictionless horizontal
surface is attached to two springs in series. The springs have force
constants k1 and k2 . Determine
p pthe angular frequency of oscillation of
the block. Answer: ω = 1/m (k1 k2 )/(k1 + k2 ).
Problem 3.43. An asteroid, initially at rest, falls from a great
distance (x0 ) towards a star. You are asked to determine the time
required for it to hit the star. The star can be assumed to be at rest
at the origin at all times. Evaluate the time required for the asteroid
to reach x = 0, that is, assume the star is a point object.
Problem 3.44. A mass m is attached to a spring of constant k.
Determine the motion if it is initially at the unstretched
R position and
is given an impulse J. (Impulse is defined by J = Fdt.)
Problem 3.45. (a) Obtain a general relation for the motion of a
particle of mass m acted upon by a repulsive force F = +kx. Assume
the particle is initially at the position x = x0 . Note that the solution
will have the form x = Aeβt + Be−βt . (b) It may surprise you to know
that for this force there is a possible solution in which the particle
moves to the origin and remains at rest there. In this case, what is the
initial value of the velocity?
Problem 3.46. An asteroid of mass m is initially very far from
Earth and has zero velocity relative to Earth. It falls under the action
of the gravitational force, F = −GmM/z 2 where M is the mass of
Earth. Determine the speed of the asteroid as a function of z, its
distance from the center of Earth.
Problem 3.47. An enterprising scientist drills a hole straight through
the Earth, from North Pole to South Pole, and drops an object of mass
m in the hole at the North Pole. How long does it take for the object
to emerge from the hole at the South Pole? You may assume that there
is no air in the hole. Hint: The gravitational force inside a uniform
sphere is given by F = −GM 0 m/r2 where M 0 is the mass contained
within a sphere of radius r and r is the distance from the center of the
sphere to the object m.
Computational Projects
Computational Project 3.1. Write a program to plot the ve-
locity and position of the canoe of Worked Example 3.3. Assume the
mass of the canoe plus person is 200 kg. Use b = 150 kg/s as a first
108 3. NEWTON’S LAWS: DETERMINING THE MOTION
estimate and get plots for various values of b. Use a reasonable choice
for the initial velocity.
Computational Project 3.2. Demonstrate that the Euler-Cromer
method is stable but the Euler method is not. Do this by assuming
a particle moves in one dimension (x) under the influence of a force
F = −kx. This will lead to simple harmonic motion. Compare the
total energy as a function of time as obtained using the Euler method
and the Euler-Cromer method. Also, on the same plot, present the to-
tal energy as a function of time as obtained from the exact (analytical)
expression. You will appreciate that the Euler method does not con-
serve energy, the total energy obtained by the Euler-Cromer method
oscillates around a constant value, and the analytical expression gives
a constant value for the energy at all times.
Computational Project 3.3. A projectile is fired perpendicu-
lar to the Earth’s surface with an initial velocity of 600 m/s. Assume a
constant gravitational force acts on the projectile so its acceleration is
9.8 m/s2 downwards. Write a program to determine the position of the
particle as a function of time. Next, consider the same problem, but
now include the effect of air resistance, assuming it is a retarding force
that can be expressed as fair = 2.5 × 10−4 v 2 where v is the velocity of
the projectile.
Chapter 4
The Lagrangian Method
In this chapter you will learn a different approach for solving physics
problems. This approach is associated with Lagrange,1 although many
other physicists and mathematicians contributed to it, including Leib-
niz and Euler.
We shall first introduce a function called the Lagrangian, and then
show how it can be used to determine the equations of motion.
A mechanical system is fully described by the Lagrangian. This
means that if you know the Lagrangian for a system, you can deter-
mine the equation of motion, the momentum, and all other relevant
mechanical quantities.
So far, we have been using Newton’s second law to determine the
equation of motion. The Lagrangian method is a very useful alterna-
tive technique for determing this equation, but before discussing the
Lagrangian, it is helpful to review how Newton’s second law leads to
an expression for the equation of motion.
l
θ
T
s mgcosθ
mg
mgsinθ
Figure 4.1. Forces acting on a simple pendulum. The
length of the string is l.
Exercise 4.1. Draw a force diagram for each of the masses of the
double planar pendulum. Consider the motion of mass m2 . What is
wrong with writing m2 s̈2 = −m2 g sin θ2 ?
y
x
l
1
θ1
m1
l2
θ m2
2
equations are obtained using the calculus of variations. You will learn
that Lagrangian dynamics is based on a profound statement about
the nature of the physical world called Hamilton’s principle. Later (in
Chapter 8) you will see how the Lagrangian is related to symmetries
in physical systems, and to the conservation laws of physics.
The Lagrangian method, as described here, is applicable to systems
in which all the forces are conservative. Consequently, we will not
consider systems with frictional forces. Although there are ways of
incorporating dissipative forces into the Lagrangian formulation, they
involve advanced methods and we do not consider them in this book.
So if friction is present, you must go back to Newton’s second law.
The study of Lagrangian dynamics begins with the definition of a
physical quantity called the Lagrangian. It is denoted by L and is de-
fined as the difference between the kinetic energy (T ) and the potential
energy (V ):
L = T − V. (4.2)
1
L = T − V = m(ẋ2 + ẏ 2 ) − mgy. (4.3)
2
x = x(q1 , q2 , q3 , t)
y = y(q1 , q2 , q3 , t)
z = z(q1 , q2 , q3 , t).
4.2. THE LAGRANGIAN 113
The q’s are called “generalized coordinates.” In practice they will often
be Cartesian, cylindrical, or spherical coordinates, but they are not
limited to these familiar coordinates.2
In the example of the pendulum, the transformation equations from
x and y to θ are
x = l sin θ,
y = −l cos θ,
where the coordinate origin is at the point of suspension and θ is mea-
sured from the negative y-axis (this is not the way we usually define
θ!)
Differentiating these equations with respect to time,
ẋ = lθ̇ cos θ,
ẏ = lθ̇ sin θ.
Plugging these expressions into the Lagrangian for the pendulum (Equa-
tion 4.3) leads to
1 1
L = m(lθ̇ cos θ)2 + m(lθ̇ sin θ)2 + mgl cos θ
2 2
1 2 2
= ml θ̇ (cos2 θ + sin2 θ) + mgl cos θ
2
1 2 2
= ml θ̇ + mgl cos θ. (4.4)
2
Equation (4.4) gives the same information as Equation (4.3) that was
previously obtained for the Lagrangian in terms of x and y, except that
now L depends only on θ.
All I have done so far is to write the Lagrangian in Cartesian co-
ordinates and then transform to polar coordinates. In doing so, the
Lagrangian went from being a function of two variables to being a
function of one variable. This procedure illustrates a very important
point: the Lagrangian should be expressed in terms of the least possible
number of coordinates.
Let me give you a simple “cookbook” procedure for obtaining the
Lagrangian in terms of a set of generalized coordinates q1 , q2 , q3 .
1. If at all possible, write the Lagrangian L = T −V in terms
of Cartesian coordinates. In Cartesian coordinates the translational
2The generalized coordinates are normally a “minimal set” of coordinates. For
example, in Cartesian coordinates the simple pendulum requires two coordinates
(x and y), but in polar coordinates only one coordinate (θ) is required. So θ is the
appropriate generalized coordinate for the pendulum problem.
114 4. THE LAGRANGIAN METHOD
kinetic energy is
1
T = m ẋ2 + ẏ 2 + ż 2 .
2
2. Write the transformation equations:
x = x(q1 , q2 , q3 , t),
y = y(q1 , q2 , q3 , t),
z = z(q1 , q2 , q3 , t).
3. Take derivatives to obtain expressions for ẋ, ẏ, ż in terms
of q’s, q̇’s and t.
4. Write L = T − V in terms of the q’s, q̇’s and t.
You might wonder why I am emphasizing that you should first write
the Lagrangian in Cartesian coordinates. The answer is simple. The
translational kinetic energy in Cartesian coordinates always has the
simple form T = 21 m (ẋ2 + ẏ 2 + ż 2 ) , whereas it can be very compli-
cated in other coordinate systems. Furthermore, the potential energy
is often (but not always) easier to express in Cartesian coordinates.
A very simple (but important) system is a block of mass m on a
frictionless surface and connected to a spring of constant k. See Figure
4.3. It is easy to appreciate that the Lagrangian for this system is
1 1
L = mẋ2 − kx2 . (4.5)
2 2
k
m
y
x
s
here:
1 1
L = (m1 + m2 )l12 θ̇12 + m2 [l22 θ̇22 + 2l1 l2 θ̇1 θ̇2 cos(θ1 − θ2 )]
2 2
+m1 gl1 cos θ1 + m2 g(l1 cos θ1 + l2 cos θ2 ).
This problem involves two coordinates, θ1 and θ2 , and therefore
there are two Lagrange equations, namely:
d ∂L ∂L
− = 0, (4.10)
dt ∂ θ̇1 ∂θ1
d ∂L ∂L
− = 0.
dt ∂ θ̇2 ∂θ2
To facilitate writing them, first evaluate the derivatives:
∂L
= −m2 l1 l2 θ̇1 θ̇2 sin(θ1 − θ2 ) − m1 gl1 sin θ1 − m2 gl1 sin θ1 ,
∂θ1
∂L
= m2 l1 l2 θ̇1 θ̇2 sin(θ1 − θ2 ) − m2 gl2 sin θ2 ,
∂θ2
∂L
= (m1 + m2 )l12 θ̇1 + m2 l1 l2 θ̇2 cos(θ1 − θ2 ),
∂ θ̇1
∂L
= m2 l22 θ̇2 + m2 l1 l2 θ̇1 cos(θ1 − θ2 ).
∂ θ̇2
Plugging into Equations (4.10) yields two coupled equations of mo-
tion:
d h 2 2
i
m1 l1 θ̇1 + m2 l1 θ̇1 + m2 l1 l2 θ̇2 cos(θ1 − θ2 )
dt
+m2 l1 l2 θ̇1 θ̇2 sin(θ1 − θ2 ) + (m1 + m2 )gl1 sin θ1 = 0,
and
d 2
m2 l2 θ̇2 + m2 l1 l2 θ̇1 cos(θ1 − θ2 )
dt
−m2 l1 l2 θ̇1 θ̇2 sin(θ1 − θ2 ) + m2 gl2 sin θ2 = 0.
These equations can be further simplified and written in the
form
h i
2 2
(m1 + m2 )l1 θ̈1 + m2 l1 l2 θ̈2 cos(θ1 − θ2 ) + θ̇2 sin(θ1 − θ2 )
+(m1 + m2 )gl1 sin θ1 = 0,
and
h i
m2 l22 θ̈2 +m2 l1 l2 θ̈1 cos(θ1 − θ2 ) − θ̇12 sin(θ1 − θ2 ) +m2 gl2 sin θ2 = 0.
4.4. DEGREES OF FREEDOM 121
(I’m sure you now appreciate why I did not attempt to solve
the double pendulum problem by using F = ma!)
Exercise 4.4. Find the equation of motion for a disk rolling down
a perfectly rough inclined plane of angle α. (The Lagrangian was ob-
tained in Worked Example 4.2.) Answer: s̈ = 32 g sin α.
Exercise 4.5. Find the equation of motion for a sphere rolling
down a perfectly rough inclined plane of angle α. (Use the Lagrangian
technique.) The moment of inertia of a sphere is 52 mR2 . Answer: s̈ =
5
7
g sin α.
Exercise 4.6. Using the Lagrangian technique, determine the equa-
tion of motion for a body of mass m falling in a constant gravitational
field.
Exercise 4.7. Find the Lagrangian and the equation of motion for
two astronomical bodies on a collision course that are attracting one
another according to Newton’s Law of Universal Gravitation. (Place
the origin of coordinates at the point where the bodies will collide.)
m1 m2
Answer: m1 r¨1 = −G (1+m 2 (where r1 is the distance from the
1 /m2 )r1
center of mass to body m1 ).
But mẋ is just the linear momentum! Therefore, for this system, the
linear momentum is related to the Lagrangian by
∂L
px = .
∂ ẋ
Next, consider the problem of a simple pendulum consisting of a
mass m hanging from a string of length l. According to Equation (4.4),
the Lagrangian for this system is
L = 12 ml2 θ̇2 + mgl cos θ.
Taking the derivative of the Lagrangian with respect to θ̇ yields
∂L
= ml2 θ̇.
∂ θ̇
2
But ml θ̇ is the angular momentum of the pendulum! In this case the
derivative of the Lagrangian with respect to the angular velocity θ̇ is
the angular momentum.
In the first case, the Lagrangian was expressed in terms of x and
ẋ. That is, L = L(x, ẋ). The generalized coordinate was x and the
generalized velocity was ẋ. In the second case, the Lagrangian was a
function of θ and θ̇, that is, L = L(θ, θ̇). Here the generalized coordinate
was θ and the generalized velocity was θ̇. These two cases illustrate the
5I must warn you that physicists are somewhat careless in their usage of the
term “generalized coordinate.” You will often hear it applied in situations in which
the qi are not all independent.
124 4. THE LAGRANGIAN METHOD
Exercise 4.8. Determine the generalized momenta pθ1 and pθ2 for
the double planar pendulum. Is either one of them a constant? Answer:
pθ2 = m2 l22 θ̇2 + m2 l1 l2 θ̇1 cos(θ1 − θ2 ). No.
Exercise 4.9. (a) Determine the generalized momentum for a sphere
of mass m and radius R rolling down an inclined plane of angle α. Let
126 4. THE LAGRANGIAN METHOD
Note that the dt has disappeared because time is frozen during a virtual
displacement. Consequently, the virtual work can be expressed as
!
X X ∂xi X X ∂xi
δW = Fi δqj = Fi δqj .
i j
∂qj j i
∂qj
4.6. GENERALIZED FORCE (OPTIONAL) 127
and it has the same form as our usual expression for work, as a force
times a displacement.
Under certain circumstances, the generalized force can be obtained
from a scalar potential function V as
∂V
Qj = − .
∂qj
This relation is analogous to the well known relation between real
forces and the potential energy, F = −∇V.
y 2
ds
dy
dx
1
x
Figure 4.5. The quantity ds is an element of the path
from 1 to 2. Note that ds2 = dx2 + dy 2 .
4.7. THE CALCULUS OF VARIATIONS (OPTIONAL) 129
The solution of the problem is the function y = y(x) that makes the
integral an extremum.7 Note that Φ is a function of a function be-
cause Φ is a function of y, and y itself is a function of x. Φ is called a
“functional” to distinguish it from an ordinary function.
Keep in mind that the integral I is a line integral between limits
x1 and x2 . It is the distance from point 1 to point 2 along a particular
path y = y(x). Obviously, if you use a different function y1 = y1 (x),
you will integrate along a different path and get a different value for
the integral.
Figure 4.6 illustrates the problem of the shortest distance between
two points in a plane. The three paths y1 (x), y2 (x), and y3 (x) have
different path lengths. The problem asks us to find the shortest of all
the possible paths between the fixed end points. Let us assume that
y0 (x) is the shortest path, and let us further assume that y1 (x) is a
path that only differs infinitesimally from y0 (x). That is, at each point
7By “extremum” we mean a maximum or a minimum, that is, values of a
function where the derivative is zero. The derivative of a function is also zero at an
inflection point, but in the calculus of variations we are always interested in either
the maximum or the minimum of the integral of a functional.
130 4. THE LAGRANGIAN METHOD
y y1(x)
y (x)
2
y3(x)
x x x
1 2
Figure 4.6. Three possible paths between two fixed end points.
on the path (except at the end points) y1 (x) is slightly different from
y0 (x). We can write the relationship between y1 (x) and y0 (x) in the
form
y1 (x) = y0 (x) + 1 η(x)
where 1 is a small quantity and η(x) is a function of x. Note that η(x)
is completely arbitrary except that it must be zero at the endpoints.
That is,
η(x1 ) = η(x2 ) = 0,
because all the paths between x1 and x2 must meet at the endpoints.
Remember that the problem is to find the minimal path y0 (x). I
will denote this minimal path simply by y(x) from now on. In general,
a nearby path is a member of a family of paths that are described by
the relationship
Y = Y (x, ) = y(x) + η (x) , (4.14)
where y(x) is the minimal path. Here is taken as a continuous vari-
able. Different paths correspond to different values of . The minimal
path is obtained when = 0.
The path length along any particular member of this family of
curves is
Z x2 Z x2
dY
I = I() = Φ x, Y, dx = Φ (x, Y, Y 0 ) dx.
x1 dx x1
Here the integral I is explicitly expressed as a function of . The desired
curve has = 0; it is the curve that minimizes I. Therefore, the
condition on I is
dI
= 0.
d =0
Observe that
d x2
Z x2
∂Φ ∂Y 0
Z
dI 0 ∂Φ ∂Y
= Φ(x, Y, Y )dx = + dx.
d d x1 x1 ∂Y ∂ ∂Y 0 ∂
4.7. THE CALCULUS OF VARIATIONS (OPTIONAL) 131
y0
p = const = C1 .
1 + y 02
Squaring both sides and rearranging,
y 02 = (C1 )2 (1 + y 02 ). (4.17)
02
Solving for y ,
y 02 (1 − C12 ) = C12
s
C12
y0 = = constant = m.
1 − C12
But y 0 = dy/dx, so
dy
= m.
dx
A final integration yields
y = mx + b,
the equation for a straight line! Thus we have proved that a straight
line is the shortest distance between two points. (Whew!)
4.7.1. Hamilton’s Principle. By this time you are surely won-
dering what any of this has to do with physics. The answer lies in
Hamilton’s principle. Hamilton’s principle states that the behavior
of any physical system will minimize the time integral of the
Lagrangian. In other words, a physical system that evolves over time
from t1 to t2 will follow a “path” q = q(t) such that the integral
Z t2
I= L(q, q̇, t)dt
t1
q’s for y’s and t for x and changing the order of the terms, the Euler-
Lagrange equation (Equation 4.16) becomes
d ∂L ∂L
− = 0,
dt ∂ q̇ ∂q
which you recognize R as Lagrange’s equation!
The integral Ldt is called the “action.” Using that terminology,
Hamilton’s principle is:
The time development of a dynamical system will minimize the action.
Hamilton’s principle is sometimes referred to as, “The fundamental
principle of mechanics.”9
To make this a bit more explicit, consider a simple physical process,
such as the motion of a falling rock. This physical system can be
considered to evolve from some initial situation at time t1 to a different
final situation at time t2 . Initially the rock is at height h and has zero
velocity. Therefore the initial conditions are (q, q̇) = (h, 0). The√final
conditions (just before it hits the ground) are q = 0 and q̇ = 2gh.
Hamilton’s principle states that the behavior of the rock (its position
and
R t2 velocity at any instant of time) is such as to minimize the quantity
t1
L(q, q̇, t)dt. This leads to the variant of the Euler-Lagrange equation
called the Lagrange equation. Finally, the Lagrange equation generates
the equation of motion, which for the falling rock, is simply d2 q/dt2 =
−g.
4.7.2. Relation to Newton’s Second Law. As mentioned above,
Newton’s second law can be derived from the Lagrange equation. As
an illustration, consider the one-dimensional motion of a particle of
mass m. Assume the particle is acted upon by a conservative force F .
Recall that in one dimension, if F is conservative it can be obtained
from the potential energy V = V (x) by F = −dV /dx. The Lagrangian
is
1
L = T − V = mẋ2 − V (x),
2
and Lagrange’s equation is
d ∂L ∂L d ∂ 1 2 ∂ 1 2
0 = − = mẋ − V (x) − mẋ − V (x)
dt ∂ ẋ ∂x dt ∂ ẋ 2 ∂x 2
d ∂V (x)
= (mẋ) + = mẍ − F.
dt ∂x
9You may have studied Fermat’s principle in optics which states that the path
of a ray of light from one point to another is such as to minimize the time of flight.
This is a special case of Hamilton’s principle.
134 4. THE LAGRANGIAN METHOD
So
F = mẍ
as expected.
Thus, Newton’s second law is a consequence of Lagrange’s equations
and Lagrange’s equations are a consequence of Hamilton’s principle.10
I suppose you might conclude that Hamilton’s principle is more basic
and important than Newton’s laws. Nevertheless, it is true that New-
ton’s laws are the basic relations from which all of physics was derived.
Hamilton’s principle is a beautiful, sophisticated, and elegant way of
describing the way nature behaves, but it is not very easily understood.
Newton’s laws, on the other hand, can be grasped by anyone with an
elementary knowledge of mathematics.
3N
X
H(p1 , p2 , · · · , p3N ; q1 , q2 , · · · q3N ; t) = pi q̇i −L(q1 , · · · q3N ; q̇1 , · · · q̇3N ; t).
i=1
10You might ask: “Why does nature always minimize the time integral of the
Lagrangian?” I do not know the answer to this question. I know that Hamilton’s
principle is an expression of the way the physical universe is put together. I know
that is how nature behaves, but I cannot tell you why it behaves that way. (Physics
tells you how, not why.)
4.8. HAMILTON’S EQUATIONS (OPTIONAL) 135
Having obtained the Hamiltonian you can use it to obtain the equa-
tions of motion of the system. Recall that the equation of motion as
expressed by Newton’s second law as well as by Lagrange’s equation
is a second-order differential equation. On the other hand the Hamil-
tonian formulation yields two first-order equations for the momentum
and position. They are:
∂H ∂H
ṗi = − and q̇i = + . (4.20)
∂qi ∂pi
These are Hamilton’s equations of motion and they are completely
equivalent to Newton’s second law and to the Lagrange equations of
motion.
There are a number of different ways of deriving Hamilton’s equa-
tions. One way is to start with Equation (4.19) and write the differen-
tial of H as
X ∂L ∂L
∂L
dH = pi dq̇i + q̇i dpi − dqi − dq̇i − dt.
i
∂qi ∂ q̇i ∂t
Using the definition of generalized momentum to cancel the first and
last term in parenthesis, and using Lagrange’s equation to write
d ∂L d ∂L
= pi = ṗi =
dt ∂ q̇i dt ∂qi
we obtain
X ∂L
dH = (q̇i dpi − ṗi dqi ) − dt.
i
∂t
But the differential of H as obtained from H = H(pi , qi , t) is
X ∂H ∂H
∂H
dH = dpi + dqi + dt.
i
∂pi ∂qi ∂t
A term-by-term comparison of the two expressions for dH yields Equa-
tions (4.20) as well as
∂H ∂L
=− .
∂t ∂t
For the sake of a specific example, consider the Hamiltonian of a
particle of mass m moving vertically in a uniform gravitational field
g. (For example, you might want to determine the Hamiltonian and
Hamilton’s equations for a ball of mass m thrown vertically upward.)
Begin with the Lagrangian. It is
1
L = T − V = mż 2 − mgz.
2
136 4. THE LAGRANGIAN METHOD
∂ p2
∂H
ṗ = − =− + mgz = −mg
∂z ∂z 2m
∂ p2
∂H p
q̇ = ż = + = + mgz = .
∂p ∂p 2m m
The second equation reads ż = p/m or p = mż which is just the
definition of momentum. The first equation states that ṗ = −mg or
d d
p = mż = mz̈ = −mg.
dt dt
That is, z̈ = −g, as expected.
Now
p2 m2 v 2 1
= = mv 2 = T,
2m 2m 2
and
V = mgz,
so
H = T + V = E = total energy.
Therefore, in this case, H = E. In fact, the Hamiltonian is nearly
always the total energy expressed in terms of generalized momentum
and position. Students frequently assume that H is always equal to
E. This is not true. The following conditions determine when H is
constant and when it is equal to E.
4.8. HAMILTON’S EQUATIONS (OPTIONAL) 137
4.9. Summary
To determine the equation(s) of motion do one of the following:
(1) Draw the free body diagram and apply F = ma. (That is,
determine the equation of motion by inspection.)
(2) Write the Lagrangian and apply Lagrange’s equations.
4.9. SUMMARY 139
4.10. Problems
Note: Problems 4.15 through 4.20 are based on optional section 4.7.
Problem 4.1. Determine the Lagrangian for a particle sliding on
the inner surface of a hemisphere of radius a (the “bar of soap in a
sink” problem). See Figure 4.7.
θ
ϕ y
x m
m
Figure 4.8. The ring connected to the spring is free to
slide on the inclined rod.
θ l
m
wedge. Since all surfaces are frictionless, the box will slide down the
wedge and the wedge will slide in the opposite direction on the plane.
Determine the accelerations of the two bodies.
Problem 4.8. A system of two masses and three springs is illus-
trated in Figure 4.8. Write the Lagrangian. Determine the generalized
momenta for this system. (Hint: Measure the positions of the masses
from their equilibrium points.)
k k k
1 2 3
M M
1 2
the wedge. All surfaces are frictionless and the spring is massless. (a)
Write the Lagrangian for the system. (b) Obtain expressions for the
generalized momenta. (c) Obtain the equations of motion. See Figure
4.12.
X y
s m
M α
Figure 4.12. A block and spring on a sliding wedge.
The dot in the middle of the spring represents the equi-
librium position of the block, and is a distance X from
some fixed point. The stretch of the spring is measured
from the equilibrium position. The position of the block
is given by x and y where x = X + s sin α.
k
m
Problem 4.17. Determine the relation y=y(x) such that the fol-
lowing integral is an extremum. What is the shape of the curve?
Z x2
1/2
(x)(1 + y 02 )
dx.
x1
axis
ds Area of shaded
Curve of
region is
length s
2πxds
Figure 4.14. A curve of length s is revolved about an
axis to generate a surface of revolution.
11This problem was apparently first devised by one the Bernoulli brothers who
sent it out as a “challenge problem” to the scientists of that day. It is said that
Newton solved the problem in a few hours. (The calculus of variations had not
been invented yet!)
146 4. THE LAGRANGIAN METHOD
Computational Projects
Computational Techniques: To solve the first computational prob-
lem below you will have to use a Computer Algebra System (CAS) such
as Mathematica or Maple. We do not have the space nor the time to
get into the details of these excellent programs, but you should learn
some of the basics of one such system. They are easily available. For
example, Matlab incorporates symbolic computation using the Maple
kernel. (Maple is a CAS that was developed primarily at the University
of Waterloo in Canada and more recently at ETH in Zurich - Einstein’s
Alma Mater.) In Matlab’s version of Maple, if you want to obtain, say,
an expression for y = cos2 x + sin2 x you simply declare x, y and z as
symbolic variables, using the command
syms x y z
then write
y = cos(x)ˆ2+sin(x)ˆ2
Then write
simplify(y)
You will obtain (of course) y=1.
Another example, is to write
z = sin(x)/cos(x)
then write
simple(z)
4.10. PROBLEMS 147
to get
z = tan(x)
Of course, you will have to consult the “help” files to get further infor-
mation, but let me assure you that using Maple is not difficult.
Computational Project 4.1. Use a symbolic manipulator such
as Maple to obtain the Hamiltonian for the double planar pendulum.
Computational Project 4.2. Solve for the motion of the double
pendulum. Assume the two bobs have the same mass and that the two
strings are of length 0.1 m. Compute examples of the motion for several
initial conditions.
Computational Project 4.3. A 2 kg object is launched hori-
zontally from a 5 m cliff with an initial velocity of 20 m/s. As you
know, it will follow a parabolic path. It hits the ground a distance 20.2
m horizontally from the launch point and will have been in the air for
1.01 s. The purpose of this project is to evaluate
R the action for this
process. The action is defined by the integral Ldt where L = T − V.
After you have written a computer program to evaluate the action for
the actual path of the projectile, evaluate the action for the following
hypothetical path between the same two end points and taking the
same amount of time. The hypothetical path is a straight line from
(0,5) to (20.2,0) at a constant speed of 20.6 m/s. You will find that the
action for the hypothetical path is greater than the action for the actual
path since the actual path for any dynamical system always minimizes
the action. You may wish to construct other paths between the end
points and show that their action is greater than that for the actual
path. (Suggestion: Use small time steps [about 1/1000 of a second],
evaluate L at each time step and estimate the integral by summing the
quantities L(t)∆t.)
Chapter 5
The Conservation of Energy
This definition may cause some confusion because dr represents the displacement
of the body, but it is not necessarily in the r̂ direction. It is wrong to write dr =drr̂
unless the displacement happens to be in the r̂ direction. For example, in spherical
coordinates,
dr = drr̂+rdθθ̂ + r sin θdφφ̂.
149
150 5. THE CONSERVATION OF ENERGY
This theorem refers to the total or net work done by all the forces
acting on the particle.
The proof of this theorem is very simple. Let F be the net force
acting on a particle of mass m. Then using the definition v =ds/dt
and Newton’s second law in the form F = m dv dt
, we can write
Z r2 Z t2 Z t2
dv
W = F·ds = F · v dt = m · v dt.
r1 t1 t1 dt
Now
d 2 d dv dv dv
(v ) = (v · v) = · v + v· =2 · v.
dt dt dt dt dt
Therefore
dv 1 d 2 d 1 2
m · v = m (v ) = mv .
dt 2 dt dt 2
So
Z r2 Z t2 Z T2
d 1 2
W = F·ds = mv dt = d(T ) = T2 − T1 = ∆T,
r1 t1 dt 2 T1
The second approach is, in principle, the same as the first approach,
but it may be easier to apply in certain situations. This approach is
based on the fact that for motion along a smooth curve, the position
can be specified by a single variable, say λ. This single independent
variable can be the distance along the curve from some starting point,
or the time, or some other parameter (as the angle θ in the exampleR
below). Then writing both F and ds in terms of λ, the integral F·ds
reduces to a single integral over one variable. Thus, if F = F(λ) and
s = s(λ), then
Z Z
ds
W = F·ds = F(λ)· dλ.
C C dλ
The last integral can be expressed in a variety of ways. For example,
if F is given in Cartesian coordinates, we write
Z λ2
dx dy dz
W = Fx + Fy + Fz dλ. (5.3)
λ1 dλ dλ dλ
This expression shows the close relationship to the first method. Nev-
ertheless, the second method is often more convenient, as illustrated
by the following example.
5.2. WORK ALONG A PATH. THE LINE INTEGRAL 153
Y
+1
X
-1 +1
changing. We are not considering the work done by the force that
is dragging the particle through this field.
Solution: Treating t as a parameter, we write F and ds in
terms of t. Since s=xı̂+ŷ we appreciate that x = 3t and y = 2t2 .
Furthermore,
ds
ds = dt = (3ı̂ + 4t̂)dt
dt
Then the force can be expressed in terms of t as
F = (3t)(2t2 )ı̂ − (3t)2 ̂ = 6t3 ı̂ − 9t2 ̂.
Consequently,
Z Z t=1
3 2
W = F·ds = (6t ı̂ − 9t ̂) · (3ı̂ + 4t̂) = (18t3 − 36t3 )dt
t=0
18 4 1 9
=−t 0 = − J.
4 2
Now let us solve the problem using equation (5.2). Note that
we need to have expressions for dx and for dy. Since x = 3t and
y = 2t2 we can solve for t and obtain y as a function of x; that is,
t = x/3, so y = 2(x/3)2 = (2/9)x2 . Then
Z 3 Z 2 Z 3 Z 2 Z 3 Z 2
2 2 2 9
W = Fx dx + Fy dy = xydx − x dy = x x dx − y dy
x=0 y=0 0 0 0 9 0 2
3 2
2 x4 9 y2 81 36 9
= − = − =− J
9 4 0 2 2 0 18 4 2
Obviously, the two methods yield the same answer.
Force Field
y y
(a) (b)
18
2
x x
0 1 4 0 3
Note that the gradient will generate a vector, the divergence (be-
ing a dot product) will generate a scalar, and the curl (being a cross
product) will generate a vector.
3In general an isopleth is a line drawn on a map through all points hav-
ing the same numerical value. You are familiar with the temperature isopleths
(“isotherms”) and pressure isopleths (“isobars”) drawn on weather maps.
158 5. THE CONSERVATION OF ENERGY
A
x
F is conservative iff ∇ × F = 0.
This statement is easy to prove. If F is conservative, then it can
be represented by the negative gradient of the potential energy, i.e.,
F = −∇V. Consequently,
∇ × F = ∇ × (−∇V ) = −∇ × (∇V ) = −curl ( grad V ) .
But the curl of the gradient of any function is zero. Therefore, if we
can write F as the gradient of V, then ∇ × F = 0.
For the sake of completeness and because the steps in the proof are
important to understand, I will now prove that curl grad V = 0 for
any function V . Using Cartesian coordinates,
∂ ∂ ∂ ∂V ∂V ∂V
∇ × ∇V = ı̂ + ̂ + k̂ × ı̂ + ̂ + k̂
∂x ∂y ∂z ∂x ∂y ∂z
ı̂ ̂ k̂
= ∂/∂x ∂/∂y ∂/∂z
∂V /∂x ∂V /∂y ∂V /∂z
2
∂ 2V
2
∂ 2V
2
∂ 2V
∂ V ∂ V ∂ V
= ı̂ − − ̂ − + k̂ − .
∂y∂z ∂z∂y ∂x∂z ∂z∂x ∂x∂y ∂y∂x
But
∂ 2V ∂ 2V
= ,
∂y∂z ∂z∂y
because the order of taking partial derivatives does not affect the an-
swer. Therefore, each term in parenthesis is zero and
∇ × ∇V = 0.
Thus, we have proved that
curl grad (any scalar function) = 0.
5.3.4. Del in Other Representations. So far we have been us-
ing Cartesian coordinates, but it is often useful to work in some other
set of coordinates, such as cylindrical coordinates or spherical coordi-
nates. I am going to show you how to obtain representations for ∇ in
these coordinate systems. However, before doing so we must discuss
coordinate transformations.5
Coordinate Transformations: Suppose you want to transform
from the Cartesian coordinates x, y, z to a different coordinate system
5If you would like to have more information on this subject, an excellent refer-
ence is Chapter 2 of Mathematical Methods for Physicists, 5th Ed.” Arfken G, and
Weber H, Academic Press, NY, 2001.
162 5. THE CONSERVATION OF ENERGY
z
y=const q 3=const
x=const q2=const
y t
z=const q 1=cons
x
(a) (b)
Figure 5.5. (a) A Cartesian coordinate system. (b) A
general coordinate system.
Just as ı̂,̂, k̂ are defined as the unit vectors of the Cartesian system, so
too you can define the quantities ê1 , ê2 , ê3 as the unit vectors of the new
coordinate system. These are defined so that ê1 points in the direction
of increasing q1 and is perpendicular to the surface q1 = constant, and
similarly for ê2 and ê3 .
The distance between two nearby points is ds. In the Cartesian sys-
tem ds is related to dx, dy, dz by the generalized Pythagorean relation
ds2 = dx2 + dy 2 + dz 2 .
If you express this distance in terms of the new coordinates you will
get a fairly complicated expression that reduces to the form
X
ds2 = h2ij dqi dqj i, j = 1, 2, 3. (5.8)
ij
∂ 1 ∂ ∂
∇ = ρ̂ + φ̂ + k̂ .
∂ρ ρ ∂φ ∂z
The element of volume in cylindrical coordinates is obtained immedi-
ately from Equation (5.10) as
dτ = ρdρdφdz.
166 5. THE CONSERVATION OF ENERGY
z
dz
dρ ρdϕ
y
r
x ρdϕ
dρ
In Exercise 5.10 you will show that these relations lead to:
ds2 = dx2 + dy 2 + dz 2
= dr2 + r2 dθ2 + r2 sin2 θdφ2 . (5.13)
By inspection, the elements of the metric (the scale factors) for spher-
ical coordinates are
h21 = 1, h22 = r2 , h23 = r2 sin2 θ.
The unit vectors are r̂, θ̂, φ̂ so using Equation (5.11) the gradient is
expressed as:
∂ 1 ∂ 1 ∂
∇ = r̂ + θ̂ + φ̂ .
∂r r ∂θ r sin θ ∂φ
From Equation (5.10), the volume element (illustrated in Figure 5.7)
is
dτ = h1 h2 h3 dq1 dq2 dq3 = r2 sin θdrdθdφ.
dr
z dϕ
rs in θ rs in θ d ϕ
r rd θ
θ dθ
x ϕ
dϕ rs in θd ϕ
Exercise 5.10. Fill in the missing steps between the two equations
in (5.13).
168 5. THE CONSERVATION OF ENERGY
Fus = −Fgrav .
Therefore, Z x2 Z x2
Fus dx = − Fgrav dx,
x1 x1
where the gravitational force is mg, directed downward.
When we raised the rock, the change in its potential energy was
Z x2
V2 − V1 = − Fgrav dx
x1
= negative of work done by the gravitational force
= (positive) work done by us.
Thus, the increase in potential energy is equal to the work done by
us, just as you would expect.
The work done by the nonconservative force (us) plus the work
done by the conservative force (gravity) is zero and since there is no
6Tobe precise, we need to exert a force slightly greater than the force of gravity
for an instant to get the rock moving, and when it reaches the top we exert a force
slightly less than the force of gravity so the rock will slow down and stop. Overall,
the average force we exert on the rock is equal to the gravitational force.
5.4. FORCE, WORK, AND POTENTIAL ENERGY 169
Note the use of the vector definition of the differential (see Equation
5.6). Furthermore, ∆V ≡ V (r2 ) − V (r1 ), so
W = −∆V.
That is, the work done by the conservative force will cause the potential
energy to decrease by the same amount. This might be the opposite
of what you expected. You have known for a long time that doing
work on an object will increase its potential energy. But please note
carefully that I said the work done by the conservative force causes an
equal decrease in the potential energy. When you raise a body from
the ground to a height h the conservative force (gravity) does negative
work and the potential energy increases. When the body falls back to
the ground, the conservative force does positive work and the potential
energy decreases.
You might wonder where the zero point of potential energy should
be situated. Recall that the change in potential energy is equal to the
negative of the work done by the conservative forces. Thus, you are
dealing only with the changes in V and the actual numerical values of
V at the end points of the process make no difference. The choice of
the zero level of V is completely arbitrary (but remember that once
you have made the choice you must stick with it).
For example Figure 5.8 shows a mass m at three different locations:
at the bottom of a well of depth d, on level ground, and on top of a
hill of height h.
If you chose the zero level of potential energy to be the level ground,
then V2 = 0; the particle in the well has negative potential energy
V1 = −mgd, and the particle on the hill has positive potential energy,
170 5. THE CONSERVATION OF ENERGY
3
h
2
dd
11
If V (r0 ) = 0, then Z r
V (r) = − F·ds.
r0
(The reference position is often taken to be at infinity.)
the effect of any other bodies in the universe.) Answer: 6.24 × 107
joules.
book was zero because it had zero kinetic energy to begin with (and
the floor is the zero point of potential energy). When the book is on
the table, once again it has zero kinetic energy and its total energy is
E = V = mgh. Where did this energy come from? Clearly you are
the ultimate source of the energy of the book. The net work done on
the book (by you and gravity) was zero, so the book’s kinetic energy
did not increase. The increase in potential energy was equal to the
negative of the work done by the gravitational force, but this energy
can be recovered (if the book falls off the table, for example). The
work done by you is not recoverable because the force you exerted was
nonconservative. If you could evaluate the chemical energy supplied
to your muscles, do you think it would be equal to mgh? (Actually it
would not, because energy is also internally converted into heat in a
working person’s body.)
Mechanics primarily deals with problems involving conservative
forces, so the conservation of energy principle is nearly always written in
the simple form E = T +V = constant, or equivalently, Ti +Vi = Tf +Vf
where subscript i represents some initial state of the system and sub-
script f represents some final state. If you wish to use the energy
principle when nonconservative forces are acting you can include them
as a work term. That is, if Wnc is the work done on the system by
nonconservative forces as the system evolves from initial state i to final
state f , then
Tf + Vf = Ti + Vi + Wnc
Z f
= Ti + Vi + Fnc · ds.
i
In solving problems, make sure you have the right sign on Wnc , de-
pending on whether the nonconservative work increases or decreases
the total energy of the system. This is usually obvious from the state-
ment of the problem or from a consideration of the behavior of the
system.
escapes, it reaches infinity with zero velocity. Ignore any other bodies
in the universe. Answer: 11.2 km/sec.
Exercise 5.20. Evaluate the velocity required for a body at 1 AU
from the Sun to escape from the solar system. Assume the solar system
consists only of the Sun (a good approximation).
A B C D E x
V(x)
E
x
A B C D E
V(x)
forbidden forbidden
region region
-x 1 x=0 +x
1
Equilibrium point
between the points labelled ±x1 . These two points are called “turning
points.” Even though the actual physical system is a block of mass m
sliding back and forth horizontally, you can represent it by the round
dot sliding up and down in a “potential well,” as shown in the up-
per part of the figure. If the dot starts from rest at a turning point
(say −x1 ) it will slide down to x = 0, picking up speed and reaching
maximum velocity at the bottom, then slowing down as it climbs the
potential hill and coming to a stop at the other turning point (at +x1 ).
The physical motion of the block is given by the projection of the dot
onto the horizontal (x) axis. At the turning points the velocity of the
block is zero. The regions beyond the turning points are labeled “for-
bidden region.” A particle in the forbidden region would have V (x) > E
(because the turning point is the location where V (x) = E). The al-
lowed region is where V (x) < E. It is obvious that V (x) > E is
impossible because T = E − V, and T would be negative if V > E.
Since T = 12 mv 2 , a negative value of the kinetic energy requires either a
negative mass or an imaginary value for the velocity. There is no such
thing as a negative mass, and an imaginary value for velocity is not
allowed because velocity is an observable quantity and all observable
physical quantities are real.8
8Physicists
frequently use complex quantities to make the mathematics of a
problem easier, but only the real part has physical significance. Keeping this simple
fact in mind will help you to avoid confusion, especially in your study of electro-
magnetic waves where the electric and magnetic fields are expressed as complex
5.7. SOLVING FOR THE MOTION: THE ENERGY INTEGRAL 179
The last expression came from evaluating the standard form inte-
gral
2ax − b
Z
dx 1
√ = −√ sin−1 √ .
ax2 + bx + c −a b2 − 4ac
5.8. THE KINETIC ENERGY OF A SYSTEM OF PARTICLES 181
Therefore,
r
2 1 −1 1 − 3x
t = − √ sin ±
m 3 2
1 − 3x √
± = − sin 60t
2
1 2 √
x = ± sin 60t.
3 3
The particle oscillates sinusoidally in a potential well between the
turning points at x = − 31 and x = +1.
Before going any further, you should refresh your memory on the
definition of center of mass. In particular, recall that for a collection
of N particles located at ri , the position of the center of mass, denoted
rc , is PN
mi ri
rc = Pi=1N
. (5.16)
i=1 m i
So
XN
M rc = mi ri , (5.17)
i=1
P
where M = mi is the total mass.
mi
z i
r'
i
r cm
i
rc
y
5.10. Summary
In this chapter you have been exposed to a number of mathematical
concepts and a number of physical concepts. In this summary the two
sets of concepts are listed separately, first the math, then the physics.
The scale factors hij depend on the geometrical properties of the coor-
dinate space and collectively are called the metric.
Del and Volume Element in Other Representations. Using
the transformation properties of coordinates we derived expressions for
the volume element and del in terms of generalized coordinates as
W = −∆V = −Vf + Vi .
Tf + Vf = Ti + Vi .
5.11. Problems
Problem 5.1. The force R on a particle is given by F =3xı̂+2ŷ N.
Determine the line integral F·ds for the straight line path that starts
at the origin and ends at the point (3,6) m. Answer: 49.5 J.
Problem 5.2. In this problem we assume that somehow an elec-
tron is dragged past a proton in a straight line path. The electron is
subject to the electrostatic force and to whatever external force keeps
it moving in a straight line at constant speed. (a) Evaluate the work
done by the electrostatic force as the electron is taken from x1 to x2
along the trajectory of Figure 5.14. (b) What is the work done by the
external force required to keep the electron on the straight line path?
(c) Evaluate the work done by the electrostatic force if the electron is
taken around a semicircular path with the proton at the center.
x ^r
x1 x2
0
q
- q
r ds = dx
d
F
+
u, v, φ :
x = uv cos φ,
y = uv sin φ,
1 2
u − v2 .
z =
2
Problem 5.7. The “elliptic cylindrical” coordinates, u, v, z, are
defined by
x = a cosh u cos v,
y = a sinh u sin v,
z = z.
Determine the metric. Write an expression for ds.
Problem 5.8. Consider the u, v, z coordinates defined by
1 2
u − v2 ,
x =
2
y = uv,
z = z.
(a) Determine the metric for these coordinates. (b) Evaluate ∇ in these
coordinates.
Problem 5.9. The prolate spheroidal coordinates η, θ, φ are related
to the Cartesian coordinates by the following transformation equations
x = a sinh η sin θ cos φ,
y = a sinh η sin θ sin φ,
z = a cosh η cos θ.
Determine the expression for ∇ using these coordinates.
Problem 5.10. (a) Using the expression for ∇ given by Equation
(5.11), obtain ∇ · V in cylindrical coordinates. Note that the deriva-
tives also act on the unit vectors, as in going from Equation (2.8) to
∂V
Equation (2.9). Answer: ∇ · V = ρ1 ∂ρ ∂
(rVρ ) + ρ1 ∂φφ + ∂V∂z
z
. (b) Show that
the same result is obtained
h from i
∇ · V = h1 h12 h3 ∂q∂1 (h2 h3 V1 ) + ∂q∂2 (h1 h3 V2 ) + ∂q∂3 (h1 h2 V3 ) .
Problem 5.12. Near the surface of the Earth the potential energy
is mgh. Use this fact to obtain an equation for the position of a falling
particle as a function of time. Show that the acceleration of the particle
is g.
Problem 5.13. Upon finding that the top of a certain mountain
is a perfect cone, a surveyor describes it mathematically in cylindrical
coordinates by the formula z = h0 − ρ. (a) Sketch the contour lines.
(b) Show that the direction of steepest ascent everywhere points toward
the summit.
Problem 5.14. A mountain rises above a flat plain. The height of
the mountain above the plane is described by the relation
z(x, y) = 2000 exp − x2 + 2y 2 /8000
meters.
(a)How high is the mountain? (b) If you are standing at x = 20 m, y =
10 m, how high above the plain are you? (c) At that location, what is
the direction of quickest descent? (d) What is the maximum rate of
descent at this point? (That is, how many meters do you descend for
every meter of horizontal displacement?) Answer: (d) -13.12 m/m.
Problem 5.15. The temperature in a certain location is given by
T = T0 − A(2x2 + y 2 + z 2 ) (in kelvin). Distances are in meters and
the constant A has the value 0.5 K/m2 . The value of T0 is 300 K. (a)
At point (1,2,2) what is the direction and rate (in K/m) of maximum
temperature increase? (b) If you move one meter in that direction
what is the temperature at the new position? (c) The temperature
at the new position as calculated by the formula is not equal to the
value obtained by multiplying the rate of temperature decrease by the
displacement. Explain this discrepancy.
Problem 5.16. (a) Determine the center of mass of the system
composed of three particles of masses 1, 2, 3 kg located at (0,1), (1,0)
and (1,1) respectively. (b) Assume the 1 kg particle has a velocity
of 1̂ m/s, the 2 kg particle has a velocity of 2ı̂ m/s and the 3 kg
particle has a velocity of 3(ı̂+̂) m/s. Determine the total kinetic energy
of the system. (c) What is the velocity of the center of mass? (d)
Determine the kinetic energy of the particles relative to the center of
mass. Answers: (a) (5/6)ı̂+(4/6)̂, (b) 31.5J, (c) 136
ı̂+ 10
6
̂ m/s, (d) 9.08
J.
Problem 5.17. Two positive charges (Q1 and Q2 ) are located on
the x-axis at positions x = ±a. A negative charge (−q) of mass m is
constrained to move along the y-axis. The electrostatic potential of the
5.11. PROBLEMS 191
system is
qQ1 qQ2
V =− −
4π0 r1 4π0 r2
where r1 and r2 are the distances from the negative charge to the
positive charges. The negative charge is released from rest at y = b.
Obtain an expression for the maximum speed of the negative charge.
Where does this occur?
Problem 5.18. A football player of mass M decides to go bungee
jumping. He suits up in a halter with a long elastic cord of unstretched
length b. After tying the free end of the rope to a high bridge he leaps
out into space. A few moments later he finds himself suspended far
above the ground. He climbs the bungee cord back up to the bridge.
Determine the ratio of the work done climbing the elastic rope to the
work done in climbing an inelastic rope of length b. You may assume
the elastic rope obeys Hooke’s law and has a force constant k.
Problem 5.19. A girl on a swing is pushed harder and harder by
her older brother. Eventually, she is swinging in an arc whose highest
point is 1.5 m above its lowest point. Assume the swing has massless
ropes 2.8 m long and that the mass of the girl (plus swing seat) is 30
kg. What is the tension in the ropes when the girl passes through the
lowest point?
Problem 5.20. A block of mass 5 kg is on a horizontal surface.
The coefficient of sliding friction is 0.5. The mass has a speed of 10 m/s
and is moving towards a stationary spring that is 3 meters away. The
spring constant is 4000 N/m. The block strikes the spring, compresses
it, and bounces back. How far from the spring does the block come to
rest?
Problem 5.21. A radioactive nucleus decays by emitting an alpha
particle with energy 6.0 MeV. The total energy released in the disinte-
gration is 6.2 MeV. Determine the mass of the recoiling nucleus (ignore
relativistic effects).
Problem 5.22. A particle of mass m and total energy E is moving
in a one-dimensional potential given by V (x) = −bx. Determine the
motion.
Problem 5.23. A rail gun is a device that accelerates a projectile
to extremely high speeds using magnetic forces. Suppose a particle of
mass m is fired vertically from the surface of the Earth. Ignoring air
resistance, show that the maximum height reached by the particle is
192 5. THE CONSERVATION OF ENERGY
given by
GM m
−R
|E|
where R is the radius of the Earth and M is its mass. (Assume that
its total energy E is negative.)
Problem 5.24. Prove that for a particle the rate of change of
kinetic energy dT
dt
is equal to the dot product of the force acting on it
and the instantaneous velocity of the particle, that is, F · v.
Problem 5.25. Let h = r − R be the position of a particle above
the surface of Earth where R is the radius of the Earth. Prove that in
the limit
hR
the gravitational potential reduces to mgh.
Problem 5.26. A particle of mass m is in a one-dimensional po-
tential energy field given by
2
V (x) = −Ae−ax ,
where A and α are constants. (a) Plot the energy diagram. (b) Deter-
mine the turning points if the particle has a total energy E = −0.5A.
(c) Determine the turning points if the particle has a total energy
E = −Ae−1 . (d) Assume the particle has zero energy and is located at
x = −∞. It is given an infinitesimal shove towards the origin. What is
its velocity when it passes through the origin?
Problem 5.27. In a certain region of space the potential energy
can be expressed as
A
V = −p ,
x2 + y 2 + z 2
where A is a constant, and the origin is excluded. (a) Obtain an expres-
sion for the force. (b) Determine the work required to take a particle
from (x1 , y1 , z1 ) to ∞. (c) Express the force in spherical coordinates.
Problem 5.28. The potential energy of a vibrating diatomic mol-
ecule as a function of the separation (s) between the two atoms is
approximately given by the “Morse Function,”
V (s) = V0 (1 − e−(s−s0 )/δ )2 − V0 ,
where s0 , δ, V0 are constant parameters. (a) Obtain an expression for
the force on the atoms. (b) Determine the separation between the
atoms when the potential energy is a minimum. (c) What is the
5.11. PROBLEMS 193
We now turn our attention to problems that can be solved using the
law of conservation of linear momentum. Examples of such problems
include the motion of a rocket, collisions in one and two dimensions,
and the behavior of a system when an impulsive force acts on it.
Next we carry out the indicated multiplications and discard all second
order differentials. (The term dM · dV is a second order differential.
The product of one infinitesimal quantity with another infinitesimal
quantity generates a very small quantity indeed!) This procedure yields
(and you should verify this result yourself):
M dV = −udM.
Dividing both sides by dt gives
dV dM
M = −u . (6.2)
dt dt
Consider the left hand side of (6.2). Since dV/dt is the acceleration of
the rocket and M is its mass, the left hand side is mass times acceler-
ation. Therefore, the right hand side looks like the force on the rocket,
usually called the thrust. The thrust depends on the rate at which fuel
is burned (dM/dt) and on the velocity with which the burned fuel is
ejected, u. It is convenient to express Equation (6.2) in scalar form as
dV dM
M = −u .
dt dt
The right hand side appears to have the wrong sign, but recall that
dM/dt is negative. Since M is changing, this equation is not of the
form F = ma. In fact, when solving problems, it is usually safer to
write
dM
dV = −u .
M
(What would you write if the rocket is slowing down by firing its retro-
rockets?)
To generate a large thrust, a rocket motor is designed to burn fuel
very rapidly (large dM/dt) and to expel it at as high a speed as possible
(large u). The whole purpose of burning the fuel is to convert it to a
198 6. CONSERVATION OF LINEAR MOMENTUM
Exercise 6.3. Joe and Bill are railroad men who are riding side
by side on flatcars rolling on parallel tracks. The tracks are straight
and perfectly horizontal. There are absolutely no frictional forces or
air resistance. It starts to snow. Joe sweeps the snow off his flatcar as
soon as it lands, sweeping it off the side, perpendicular to the direc-
tion of motion of the flatcar. Bill, the lazy one, simply lets the snow
accumulate on his flatcar. Who travels further in the same interval
of time? Answer this question conceptually and also mathematically.
(The snow falls perfectly vertically.)
200 6. CONSERVATION OF LINEAR MOMENTUM
6.3. Collisions
Another application of the principle of conservation of linear mo-
mentum is in the analysis of collisions. Generating and studying col-
lisions between elementary particles are a principal way physicists ex-
plore the underlying properties of nature. A collision between two
bodies often involves a strong, short range interactive force between
the bodies. When two extended bodies (such as automobiles or bil-
liard balls) come in contact, we idealize the collision and assume no
forces act except during the instant of contact. Taking the two bod-
ies as the entire system, these forces are internal forces. During an
ideal collision, no external forces are acting. Consequently, the total
momentum of the system is constant.
In a contact collision, there is a very short range repulsive force that
acts while the surfaces of the two bodies are touching.2 A glancing
collision, such as illustrated in Figure 6.2, occurs when the velocity
vectors of the two bodies are not aligned along the line of centers.
The distance b between the initial velocity vectors is called the impact
parameter.
A collision may not involve the actual physical contact of the two
bodies. The force exerted by one body on the other may be a long range
force such as the force of electric repulsion between an alpha particle
and the nucleus as in Rutherford’s experiment, or the force of gravity
when a comet in a hyperbolic orbit approaches from infinity, swings
2Theorigin of these forces is the repulsion between the electrons in one body
and the electrons in the other. When the two “electron clouds” start to overlap,
there is a repulsive Coulomb force between them. Fortunately, we do not need
detailed information about the force between the bodies because we can solve the
problem using the law of conservation of momentum in which internal forces play
no part.
6.3. COLLISIONS 201
about the Sun, and travels back out to infinity. The same physics
applies regardless of the range of the forces.3
As a basic collision problem, consider the situation illustrated in
Figure 6.3 where body M1 with velocity V1 makes a glancing collision
with body M2 that is initially at rest. (We can always find a coordinate
system in which one body is initially at rest.) The two bodies move
off with velocities V10 and V20 at angles θ1 and θ2 relative to V1 , as
shown in the figure. To keep things simple, assume the bodies are not
rotating.
There are no external forces acting on the system so the law of
momentum conservation states that
Pi = Pf . (6.3)
That is, the initial momentum and the final momentum are equal. This
is a vector equation so it is equivalent to the three scalar equations
Pxi = Pxf , Pyi = Pyf , Pzi = Pzf ,
(if two vectors are equal, their components must be equal). It is con-
venient to place the origin of coordinates at the original position of
body M2 and to let the x-axis be defined by the direction of V1 . Select
the z-axis perpendicular to the plane of the motion, i.e., perpendicular
to the plane containing V10 and V20 . Then Pzf = 0. By momentum
conservation, Pzi = 0. Therefore the problem is two-dimensional; the
motion takes place entirely in the xy-plane.
3We often refer to colliding bodies as particles, even though they may be as-
tronomical objects. It is appropriate to use the term particle when we are dealing
with long range forces. In a glancing collision, the surfaces of two extended bodies
come in contact and they should not be called particles. Nevertheless, physicists
are a bit careless in the usage of this term, and in dealing with collisions the word
particle is often used when, strictly speaking, it should not be.
202 6. CONSERVATION OF LINEAR MOMENTUM
Going back to the general case, I will now show you how to manip-
ulate momentum conservation and kinetic energy conservation, Equa-
tions (6.4), (6.5), and (6.6), to solve for the unknown quantities in
terms of the known quantities. Let me warn you that the algebra is a
bit tedious so you might want to get up and get a cup of coffee before
we start. Also, I am not going to give you all the intermediate steps,
so while you are up you had better get your pencil because you will
not understand the final result unless you can work out all the steps.
To begin, rewrite Equations (6.4) and (6.5), placing all the terms
with subscript 1 on one side. Thus:
This equality corresponds to our experience that the struck object flies
off with the velocity the incoming particle had before the collision.
Subcase 1.2 M1 6= M2
Now allow the two colliding bodies to have different masses, but
still assume a head-on collision so θ1 = θ2 = 0. Again, Equation (6.11)
reduces to Equation (6.12). Selecting the plus sign gives V10 = V1 ,
implying no collision at all. This case is of no interest, so select the
minus sign. That is,
M1 − M2
V10 = V1 . (6.13)
M1 + M2
Therefore, if M1 > M2 , the velocity V1 0 is positive. Plugging Equation
(6.13) into (6.7) and keeping in mind that θ1 = θ2 = 0, you get the
velocity acquired by M2 :
M1
V20 = 2V1 . (6.14)
M1 + M2
Equations (6.13) and (6.14) agree with your experience and intuition.
If a heavy (massive) object collides with a light (less massive) object,
the heavy object will keep on moving in its original direction. If a
billiard ball strikes a ping-pong ball, M1 >> M2 , and
. M1
V10 = V1 = V1 .
M1
That is, the billiard ball keeps on moving at essentially its original
velocity. What happens to the ping-pong ball? According to Equation
(6.14), if M1 >> M2 , then V20 = 2V1 . That is, the light object moves
off with a speed of twice the speed of the incoming heavy object.
It is easy to appreciate that if M1 < M2 , the situation is simply
reversed. If a light object hits a heavy object then M1 < M2 and V10 is
negative. That is, the light object bounces back. For example, if you
throw a ball against a wall, then M1 is the mass of the ball and M2
is the mass of the Earth (assuming the wall is attached to the Earth).
So M1 << M2 and according to Equation (6.13), V10 , the final velocity
of the ball, will be V10 = −V1 . The ball bounces back with its initial
speed. Similarly, V20 = 0. The Earth stands still.
π
θ1 + θ2 = .
2
Exercise 6.8. Show that the null solution for Equation (6.15) im-
plies a head-on collision.
Exercise 6.9. Equation 6.16 leads to the two solutions V20 = ±V1 sin θ1 .
Using the facts that θ1 + θ2 = π/2 and V10 = V1 cos θ1 , show that only
the positive solution is obtained. (The negative solution corresponds
to θ1 = 0 and V10 = V1 , that is, a miss.)
In the limit M2 /M1 → 1 you obtain θmax = cos−1 (0) = π/2. In the
limit M2 /M1 → 0 you obtain θmax = cos−1 (1) = 0. Therefore,
π
0 < θ1 < .
2
This tells you that if the incoming object is the more massive body, it
will be scattered through an angle θ1 smaller than π/2.
Subcase 2.3: M1 < M2
The case of a glancing collision in which M1 < M2 can also be
analyzed with Equation (6.11). It is left as an exercise to show that if
the target M2 is much more massive than M1 , then the incoming body
will bounce off with its original speed (but with a change in direction).
Note that this case reduces to Subcase 2.2 if you interchange the two
bodies.
In conclusion, you have seen that an elastic collision between two
bodies can be analyzed using the conservation of linear momentum
and the conservation of kinetic energy. The equations obtained are
amazingly complex for such a simple problem. An important benefit
you should get from the analysis of collisions is an appreciation for how
to extract physical meaning from mathematical relations. Be aware
that in a collision the conservation of momentum always holds, but
conservation of kinetic energy may not.
6.5. Impulse
When a bat hits a baseball, a large force acts for a short period of
time. Such a blow gives rise to an impulse. By definition, an impulse
6Thecoefficient of restitution actually depends on various other factors, such
as the medium in which the collision occurs, but Newton’s formula is a good
approximation.
212 6. CONSERVATION OF LINEAR MOMENTUM
(e)
where Fi is the external force acting on i. Note that the summation
over the internal forces is subject to the condition j 6= i because a
particle cannot exert a force on itself.
There is one such equation for each particle. Adding all N equations
yields
N N N N
X dpi dPtot X (e) X X
= = Fi + Fij ,
i
dt dt i i j6=i
P
where Ptot = pi is the total momentum of the system. The last term
(the double sum) is zero because by Newton’s third law, Fij = −Fji .
By writing out a few terms you will see that the double sum consists
of pairs of terms that cancel each other out. Therefore,
dPtot X (e) (e)
= Fi = Ftot (6.17)
dt i
(e)
where Ftot is the total (or net, or resultant) external force acting on the
system. Thus, we have shown that the internal forces have no effect on
the total momentum.
This result has another important consequence. If the masses of
the particles are constant, we can write
dPtot X
= mi r̈i ,
dt i
m1
r2-r1
r1
m2
r2
O
and
m2 r̈2 = +F.
Figure 6.5 shows two particles, m1 and m2 , at positions r1 and r2
with respect to the inertial origin O. The “relative” vector r gives the
position of m2 with respect to m1 and (by tip-to-tail addition) is given
by
r = r2 − r1 .
Differentiating the relative coordinate twice with respect to time yields
r̈ = r̈2 − r̈1 .
Substituting for r̈2 and r̈1 from the equations of motion gives
F F m1 + m2
r̈ = + = F
m2 m1 m1 m2
or
m1 m2
r̈ = F.
m1 + m2
The quantity mm11+m
m2
2
is called the “reduced mass.” It is denoted by µ.
So the equation of motion for the relative coordinate is
µr̈ = F. (6.20)
As an example, consider a system composed of a star and a planet.
The star is much more massive than the planet and for all intents
and purposes the star remains at rest and the planet orbits around it.
If the mass of the star is m1 and the mass of the planet is m2 and
if m1 >> m2 , then µ ∼ = m2 . On the other hand, for a binary star
system both masses may be approximately equal and the two stars
orbit around their common center of mass. If both stars have the same
mass, say m, the reduced mass is µ = 21 m.
Consider the Sun-Earth system. The relative coordinate gives the
distance from the Sun to the Earth. F is the force the Sun exerts on the
Earth. In an analysis of this system you should write µr̈ = F and not
m2 r̈ = F. The reason is that the Sun is accelerating, so a coordinate
system with origin at the Sun is not an inertial coordinate system
and Newton’s second law does not hold. (Actually for the Sun and
Earth, m2 and µ are so nearly equal that the error in using m2 r̈ = F
is negligible.)
m1 v'1cm
v1cm ϕ v2cm
m1 ϕ m2
m2
v'2cm
Figure 6.6. A collision as seen in the center-of-mass
coordinate system.
in Figure 6.6. Note that there is now a single angle, so we have already
achieved some simplification of the problem.
Figure 6.7 shows the positions of the two particles and their center
of mass relative to the origin O of an arbitrary inertial frame. The
center of mass (indicated by the small open circle) lies on the line
joining the two particles and is at rc relative to O. The coordinates
r1cm and r2cm give the positions of the particles relative to the center
of mass. Note that
r1cm = r1 −rc
r2cm = r2 −rc
where the factor Q represents any energy gained or lost during the
collision. For the rest of this section I will assume elastic collisions so
Q = 0.
Let us go back to the basic collision of Section 6.3 in which a particle
of mass m1 and speed V1 strikes a particle of mass m2 at rest, but now
we analyze the problem in the center of mass coordinate system. I will
use capital letters (such as V1lab ) for speeds in the laboratory frame,
and small letters (such as v1cm ) for speeds in the center of mass frame.
The speed of the center of mass in the lab frame is denoted Vc and
is obtained by taking the time derivative of the definition of center of
mass,
m1 r1 + m2 r2
rc = ,
M
m1 ṙ1 + m2 ṙ2
∴ Vc = .
M
Since ṙ1 = V1lab and ṙ2 = V2lab = 0 the velocity of the center of mass
is
m1
Vc = V1lab . (6.24)
M
In the laboratory system the origin of coordinates is located at
the initial position of the particle at rest. After the collision the two
particles move off at angles θ1 and θ2 . In other words, the laboratory
frame is described by Figure 6.3.
The transformation between laboratory frame velocities and center
of mass velocities is illustrated in Figure 6.8, which shows the rela-
0 0
tionship between v1cm and V1lab , as seen in the laboratory frame of
reference. In constructing the figure I used the vector relationship
This equation tells you that to convert velocities from the cm frame to
the lab frame you simply add the velocity of the center of mass.
We now obtain some useful relationships between the velocities in
the two coordinate systems. The velocity of the center of mass in the
lab system is given by Equation (6.24). Then, according to Figure 6.8
6.8. COLLISIONS IN CENTER OF MASS COORDINATES (OPTIONAL) 219
Vc
v'1cm
V'1lab
ϕ
θ1
direction of V1lab
Figure 6.8. Relationship between velocities in the cen-
ter of mass system and the laboratory system for the
final velocities of particle number 1.
θ2 + θ1 = π/2.
This result indicates that if the two masses are equal, the angle
between the two outgoing particles in the lab system is a right
angle.
6.9. Summary
The law of conservation of linear momentum is applicable to nu-
merous problems. In this chapter we have been particularly interested
in two problems: the motion of a rocket and the collision of two masses.
Conservation of momentum is based on Newton’s second law. Since
F =dp/dt, if the net external force is zero, the momentum is constant.
A system such as a rocket (or a conveyor belt) whose mass is chang-
ing with time but which is not acted upon by external forces, is ana-
lyzed by requiring that the initial and final momenta be equal. For the
rocket, this leads to
dV dM
M = −u ,
dt dt
222 6. CONSERVATION OF LINEAR MOMENTUM
or,
dM
dV = −u ,
M
where u is the speed of the ejected gases relative to the rocket.
The analysis of a glancing collision of two objects is also based on
the conservation of linear momentum. If the masses are given, the
problem can be expressed in terms of the parameters V1 , V10 , V20 , θ1 , θ2 .
Momentum conservation yields two equations. Thus you need to be
given three of these parameters. In the case of elastic collisions, con-
servation of kinetic energy gives you an additional equation and you
only need to be given two of the parameters.
Assuming an elastic collision and that V1 and θ1 are given, the
conservation laws lead to an equation for V10 , the final velocity of M1 .
s
0
2
V1 M1 cos θ1 ± cos2 θ1 − 1 + M2 .
=
V1 (M1 + M2 ) M1
elastic collisions, the scattering angles in the lab and cm systems are
related by
sin φ
tan θ1 = 0
,
cos φ + (Vc /v1cm )
sin φ
tan θ2 = ,
1 − cos φ
where θ1 and θ2 are the scattering angles in the laboratory coordinate
system and φ is the (single) scattering angle in the center-of-mass sys-
tem.
6.10. Problems
Problem 6.1. A 90 kg railroad worker is on a handcar of mass 200
kg. The handcar is moving at 5 m/s when it passes under a tree. (a)
The railroad worker leaps upwards, grabs a limb and hangs on. Does
the speed of the handcar change? If so, determine its final velocity. (b)
Now consider the converse problem. The empty handcar is moving at
5 m/s when it passes under a tree and a 90 kg railroad worker drops
out of the tree onto the handcar. In this case, does the speed of the
handcar change? If so, determine its final velocity.
Problem 6.2. Prove that in a one dimensional elastic collision
between two bodies, the relative velocity between the bodies has the
same magnitude before and after the collision, but the opposite sign.
Problem 6.3. (The ballistic pendulum.) A ballistic pendulum
can be used to determine the speed of the bullet fired from a rifle by
determining its effect when it hits and is embedded in a pendulum.
Consider a ballistic pendulum that consists of a suspended block of
wood of mass M . A bullet of mass m and initial velocity v is fired
into and becomes embedded in the block. To block swings upward a
height h. (See Figure 6.9) Derive an equation for v in terms of the given
quantities.
out of a hole in the box at a constant rate. The slope of the hill is α.
(a) Show that the equation of motion of the box (plus sled) is just
dv
m = mg sin α.
dt
(b) Now assume that the sand is somehow thrown out of the box with
a velocity −v, that is a velocity in the direction opposite to the motion
of the box but with the same speed as the box. Show that the equation
of motion in this case is
dv dm
m =v + mg sin α.
dt dt
Problem 6.6. A jet boat (or jet ski) operates on the following
principle: Water is drawn through an inlet into a turbine and pumped
out at high speed through a smaller opening. The manufacturer of a
jet boat states that the turbine draws 50 gallons of water per second
and expels it at a pressure of 80 psi. The manufacturer states that the
thrust developed is over 2000 lbs. Determine whether or not this is a
realistic value. (Hint: Look up Bernoulli’s equation.)
Problem 6.7. A spherical asteroid of mass m0 is moving freely
in interstellar space with velocity v0 . It runs into a dust cloud whose
uniform density is ρd . Assume that every particle of dust that hits the
asteroid sticks to it. (a) Obtain an expression for the velocity of the
asteroid as a function of time. (b) Obtain an expression for the force
exerted on the asteroid by the dust as a function of time. Answer (a)
h i−3/4
4/3
v = v0 m0 m0 + 4kv 3K
0 m0
2/3 t where k = ρd π and K = m/r3 .
of a scale. The student lets go of the chain and observes the reading
on the scale while the chain is dropping. The student claims that
the reading on the scale is three times the weight of the length of
the chain on the scale. The Laboratory Instructor doubts the result
obtained by the student. Show that the student’s observation is correct.
(The reading of the scale is the force exerted by the scale to stop the
downward motion of the chain.)
Problem 6.14. Many years ago on a cold winter morning in Chicago,
Bonnie and Clyde stole an armored truck full of money. (The mass of
the truck was 2000 kg and its top speed was 240 km/hr or 66.6 m/s.)
Officer Dick Tracy and his driver spotted them and gave chase. (The
police car had a mass of 1500 kg and its top speed was also 240 km/hr.)
As luck would have it, the two vehicles ran off the bank onto Lake
Michigan which was covered with perfectly frictionless ice, so the two
vehicles continued to move at constant speed and maintained a con-
stant separation. Dick Tracy grew impatient, so he opened the moon
roof, stood up, and started to shoot at the armored truck with his
machine gun. The machine gun fired 120 bullets per minute with a
muzzle speed of 1000 m/s. Each bullet had a mass of 0.05 kg. All of
the bullets hit and were embedded into the armored truck. After one
minute of this, what was the speed of the police car and what was the
speed of the truck? Answer: 62.6 m/s, 69.4 m/s.
Problem 6.15. An Atwood’s machine uses two containers filled
with water on either side of the ideal, frictionless pulley. Initially,
both buckets contain the same amount of water and weigh the same.
However, one of the containers has a small hole in it, and water is
leaking out at a rate k (kg/s). The leaking container will, therefore,
move upward. Obtain an expression for the velocity of this container.
Answer: v = −gt+(2mg/k) ln [(2m)/(2m − kt)] , where m is the initial
mass of the buckets plus water.
Problem 6.16. A rocket of mass 40,000 kg is in empty space.
Determine its velocity increase after burning all its fuel if the mass of
the fuel is 90% of the mass of the rocket. The rate of fuel burn is
constant. The speed of the exhaust gas relative to the rocket is 3000
m/s.
Problem 6.17. A rocket of mass 1500 kg is designed to expel
exhaust gas at 1000 m/s. Determine the minimum required burn rate
if the rocket is to rise from the surface of the Earth. What burn rate
is required for it to have an initial acceleration of 1 m/s2 ?
6.10. PROBLEMS 227
Hint: Express the time as an infinite series. Note that the sum of a
geometric series of the form a, ar, ar2 , ar3 , · · · is S = a/(1 − r).
Problems 6.30 to 6.32 are based on Optional Section 6.8
Problem 6.30. Show that in the center of mass system, the con-
servation of kinetic energy (equation 6.23) reduces to
2 0
P1cm P2
= 1cm
2µ 2µ
if Q = 0.
Problem 6.31. Show that in an elastic collision between two par-
ticles the relative velocity does not change.
Problem 6.32. Show that the scattering angle φ in center of mass
coordinates is related to the angle θ2 in the lab coordinates by
sin φ
tan θ2 = .
1 − cos φ
(In other words, derive Equation 6.29.)
Computational Projects
Computational Techniques: The Runge-Kutta Method
In Section 3.7 I described the Euler Cromer algorithm for solving
ordinary differential equations (ODE’s). Here I will describe the Runge-
Kutta method which is more accurate and perhaps more elegant, but
less transparent. It is based on the truncated Taylor Series expansion
for a function g(t):
dg
g(t + τ ) = g(t) + τ
dt ξ
l ≡ r × p. (7.1)
See Figure 7.1. According to the definition of cross product, the angular
momentum is perpendicular to the plane defined by r and p. By the
right hand rule, the angular momentum of the particle illustrated in
Figure 7.1 points into the page.
The magnitude of the angular momentum is given by
l = mvr sin θ.
From the figure it is easy to appreciate that r sin θ is a constant; call it
b. Then l = mvb.
233
234 7. CONSERVATION OF ANGULAR MOMENTUM
dl
if N = 0 then dt
= 0 and l = constant.
7.3. ANGULAR MOMENTUM OF A SYSTEM OF PARTICLES 237
You will not be surprised to find that the internal torques these
particles exert on each other do not affect the total angular momentum.
That is, just as internal forces do not affect the total linear momentum,
so too, internal torques do not affect the total angular momentum.
Suppose the N particles have masses m1 , m2 , · · · , mN and at some
instant of time, these particles are located at positions r1 , r2 , · · · , rN .
All of the particles exert (internal) forces on each other. The force on
particle i, due to particle j, will be denoted Fij . Particle i may also
(e)
be subjected to an external force (denoted Fi ). Consequently, the
equation of motion for particle i is,
N
X
(e)
mi r̈i = Fi + Fij .
j=1
j6=i
Now note that the rate of change of the angular momentum vector of
particle i is
dli d
= (ri × mi ṙi )
dt dt
= mi (ṙi × ṙi ) + mi (ri × r̈i )
= ri × mi r̈i .
Therefore, the left hand side of Equation 7.4 is just the rate of change
of the angular momentum, so
dli (e)
X
= ri × Fi + ri × Fij .
dt j6=i
There is one such equation for each particle. Adding all these equations
gives !
N N
X dli X (e)
X X
= ri × Fi + ri × Fij . (7.5)
i
dt i=1 i j6=i
But
X dli d X d
= li = L,
i
dt dt i dt
where L is the total angular momentum of the system, defined as the
vector sum ofthe angular momenta of all the particles. Furthermore,
P (e) (e)
i ri × Fi is the total torque Ntot on the system due to external
forces. Consequently, Equation (7.5) can be written as
dL (e)
XX
= Ntot + ri × Fij .
dt i j6=i
between the particles is directed along rij and so rij and Fij are paral-
lel. Therefore, the cross product rij × Fij is zero, and all the terms in
the double sum cancel out pairwise, leaving
dL (e)
= Ntot . (7.6)
dt
That is, the rate of change of the total angular momentum is equal
to the net external torque on the system. In particular, the law of
conservation of angular momentum can be expressed as follows:
If the net external torque acting on a system of
particles (or an extended body) is zero, the total
angular momentum of the system will remain constant.
This is the law of conservation of angular momentum. It is one of
the basic physical principles governing material objects and, as far as
we know, it is never violated. When you study quantum mechanics
and are introduced to the concepts of the orbital and the spin angular
momentum of electrons, you will begin to appreciate the power of this
conservation law.
Furthermore, the law of conservation of angular momentum is a
very useful tool for solving physics problems. There is a wide variety
of problems in which no external torque acts on a particle or system
of particles. Then the angular momentum is constant and the angular
momentum before some event is equal to the angular momentum after
the event.
For example, suppose a comet is attracted to some star, approaches
it, swings about the star in a hyperbolic path, and then travels back out
to “infinity.” The force exerted by the star on the comet is along the
line joining them, so the torque is zero. Therefore, the initial angular
momentum is equal to the final angular momentum, where “initial”
and “final” can refer to any two points along the trajectory.
A common problem found in introductory physics textbooks has
a running child leap onto a playground merry-go-round. This type of
problem can be solved by setting the total initial angular momentum
(the value for child plus merry-go-round before the child jumps aboard)
equal to the total final angular momentum (the value after the child
has jumped on).
Exercise 7.3. A little girl (mass = 20 kg) runs at 3 m/s and jumps
onto the rim of a playground merry-go-round that consists of a 50 kg
240 7. CONSERVATION OF ANGULAR MOMENTUM
rate of the Earth. As the Earth rotates more and more slowly, the day
grows longer. In several billion years the day will be so long that it will
be equal to the orbital period of the Moon. That is, the day will be as
long as the month.
If we consider Earth-Moon to be an isolated system, there are no ex-
ternal torques on it, and the total angular momentum must be constant.
We have seen that the angular momentum of the Earth is decreasing,
so how can the total angular momentum remain constant? The answer
is that the angular momentum of the Moon is increasing. This requires
an increase in the distance between the Earth and the Moon. That is,
the Earth is slowing down, losing rotational angular momentum, and
the Moon is receding from the Earth, gaining orbital angular momen-
tum in such a way as to conserve the total angular momentum of the
system.3
3The tidal interaction between Earth and Moon has many fascinating aspects.
It has been used to explain the eccentricity of the Moon’s orbit, but recent studies
seem to indicate that the gravitational force of Jupiter may be more important.
If you are interested, you might look up the paper by Matija Cuk, “Excitation of
lunar eccentricity by planetary resonances,” Science, 318, 244 (2007),
242 7. CONSERVATION OF ANGULAR MOMENTUM
mi r0i = 0.
P
Exercise 7.7. Prove that i
^
ei
ri vi
mi
ri
i
k
Figure 7.9 gives three different views of the precession of the angu-
lar momentum vector around the vertical. The leftmost figure shows
the angular momentum vector and the circle traced out by the tip of
this vector. Note that as time passes, the angular momentum goes
from L to L + ∆L. The magnitude of the angular momentum does not
change, only its direction, so ∆L lies in the plane of the circle (and
is perpendicular to both L and L + ∆L). The middle figure is a side
view, indicating that θ is the angle between L and the vertical, and
that the horizontal component of L is L sin θ. The rightmost figure is a
top view, looking down on the circular path, indicating that ∆L is the
chord of the arc subtended by the angle ∆φ. Setting the chord equal
to the arc we have
∆L = L sin θ∆φ.
Divide by ∆t and let ∆t → dt, to obtain
dL dφ
= L sin θ .
dt dt
dL dφ
Now recall that dt
= RM g sin θ, and dt
= Ωp so
RM g sin θ = L sin θΩp ,
and consequently
RM g RM g
Ωp = = .
L Iω
It is interesting to note that the precession rate increases as the top
slows down.
Since the angular momentum vector and the angular velocity vector
(ω) lie along the same line, they vary together. The angular velocity
vector is precessing at the same rate as the angular momentum vector.
A gyroscope also “nods” as it precesses. This is called nutation.
We shall consider it in Chapter 16.
z z
Lsinθ
ΔL ΔL
ΔΦ
L L+ΔL
L Lsinθ
θ
7.8. Summary
The angular momentum of a particle is defined as
l=r×p
and, consequently, it depends on the choice of origin. Since
dl d
= (r × p) = r × F = N
dt dt
256 7. CONSERVATION OF ANGULAR MOMENTUM
7.9. Problems
Problem 7.1. Two particles of equal charge and equal mass travel
at the same speed v but in opposite directions. Their initial trajecto-
ries are straight lines separated by a distance b. Due to their mutual
repulsion, they will reach a point of closest approach and then move
away from each other. The electrical force between the charges is given
by F = (q 2 /4π0 r2 )r̂. The vector rr̂ is drawn from one particle to the
other. Although it is unrealistic to do so, ignore the magnetic force
between the moving charges. (a) Draw the trajectories of the two
particles, indicating the electric forces acting on them. (b) Explain
whether or not the angular momentum of the system is constant. (c)
Do your answers depend on your choice of coordinate origin?
Problem 7.2. A uniform rod of length d and mass M is moving
at a constant velocity v in the direction of its length. (a) Obtain its
angular momentum relative to an arbitrary origin and show that this
is the same as the angular momentum of a particle of mass M moving
along the trajectory of the rod. (b) If the rod is inclined at an angle α
to its velocity vector, what is the position of the equivalent particle?
Problem 7.3. Prove that the total angular momentum of a system
of two particles is independent of the displacement of the origin only if
the total linear momentum is zero or if the displacement of the origin
is parallel to the total linear momentum.
Problem 7.4. In proving the law of conservation of angular mo-
mentum for an extended body (or system of particles) we assumed
that Newton’s third law is obeyed in the strong form. In this problem
you will investigate this assumption for the case of particles interacting
through magnetic forces. (The magnetic force between moving charged
particles is not directed along the line joining the particles.) The or-
bital motion of an electron around the nucleus is equivalent to a small
current loop, which we usually call a “magnetic moment.” A magnetic
moment generates a magnetic field. Another electron orbiting around
another nucleus will feel a force due to this magnetic field. Let us gen-
eralize the problem to two current loops of arbitrary shape that carry
currents I1 and I2 . Considering small elements of the current loops,
I1 dl1 and I2 dl2 , the force on I1 dl1 due to I2 dl2 is
dF12 = I1 dl1 × B
where
I2 dl2 × r12
I
µ0
B= .
4π |r12 |3
2
258 7. CONSERVATION OF ANGULAR MOMENTUM
Show that the force between the two current loops obeys Newton’s third
law in the strong form. (Hint: this involves showing that dl1 · r12 / |r12 |3
is an exact differential.)
Problem 7.5. An electron is fired horizontally with speed voy be-
tween parallel charged electrical plates. Assume the plates are oriented
horizontally and that the top plate is negatively charged and the bot-
tom plate is positively charged, so the electron feels a uniform force
downward. Denote this force by Fe . Using Cartesian coordinates, cal-
culate the angular momentum as a function of time and show that the
time rate of change of angular momentum is equal to the torque exerted
on the electron by the electric field. Evaluate the angular momentum
(and torque) relative to an origin located at the position of the elec-
tron when it first enters the electric field. See Figure 7.13. Answer:
l = −(1/2)v0y Fe t2 ı̂.
for the moment of inertia about a parallel axis that is tangent to the
edge of the disk.
Problem 7.13. Determine the moment of inertia of an annular
cylinder (or ring) about an axis perpendicular to the plane of the ring
and passing through its center. The ring has uniform density, mass M,
inner radius R1 and outer radius R2 .
Problem 7.14. Derive an expression for the moment of inertia of
a uniform, solid sphere about a diameter.
Problem 7.15. Determine the moment of inertia of a flat disk of
mass M and radius R about an axis tangent to the edge of the disk
and lying in the plane of the disk.
Problem 7.16. Assume the Earth is a sphere of mass M and radius
R. Remove the Southern Hemisphere. By direct integration determine
the moment of inertia of the remaining hemisphere about an axis per-
pendicular to the equatorial plane and passing through the North Pole.
Problem 7.17. A disk of mass M and moment of inertia I has a
hole in its center. An axle, which is slightly smaller in diameter than
the hole, passes through the hole. The axle is horizontal and the plane
of the disk is vertical. The disk spins smoothly on the axle. There is
friction between the disk and the axle, and the coefficient of friction
is µ. Assume the disk has an initial angular speed ω0 . Determine the
number of turns and the time for the disk to stop. (Hint: the point of
contact is not over the center of mass.)
Problem 7.18. A dumbbell is made up of a rod of length l and
mass m connected on either end to spheres of mass M and radius
R. (a) Determine the moment of inertia of the dumbbell about an axis
perpendicular to the rod and going through its center. (b) You suspend
the dumbbell from its center of mass and apply a force of 1.5 N at the
center of mass of one of the spheres. The force is in the horizontal plane
and perpendicular to the rod. Determine the angular acceleration of
the system. Assume M = 1 kg, R = 10 cm, l = 20 cm and m = 0.1
kg. Answer: (b) 0.1 rad/s2 .
Problem 7.19. The gyroscope of Figure 7.6 is made up of a disk
of mass M and radius a and an axis of negligible mass that is free to
swivel in any manner about the bottom end. It is inclined at an angle
θ = 90◦ so the axis is horizontal and the gyroscope is precessing. The
disk spins about the axis with angular speed ω. Write an expression
for the total angular velocity vector in Cartesian coordinates in terms
of M, a, R, ω, the time, and any other appropriate parameters.
7.9. PROBLEMS 261
8.1. Symmetry
Nature exhibits many symmetries, from the the structure of atoms
to the patterns of snowflakes. As physicists, we exploit the concept
of symmetry to understand the nature of physical reality. On a more
mundane level, we find symmetry concepts to be a valuable aid in
solving physics problems.
All of us have an intuitive concept of what is meant by symmetry.
We recognize spatial symmetry in flowers and snowflakes and in the
bodies of most living creatures. We say these things “look the same”
when viewed from different points of view. For example, a snowflake
looks the same when rotated through 60 degrees (known as hexagonal
symmetry). A human body looks the same when viewed in a mirror.
A cylindrical vase looks the same when rotated through an arbitrary
angle. Although its mirror image may be indistinguishable from the
vase itself, it will probably look different when it is turned upside down.
A physicist would say the vase has rotational symmetry and symmetry
under reflection, but it does not have “right side up - upside down”
(axial) symmetry.
From the examples above, you can appreciate that a symmetry is
related to an invariance under some particular operation. In simpler
words, a symmetry means that a system is not changed when something
is done to it. If there is no way we can tell whether or not a vase has
been rotated through some angle, then it has rotational symmetry. If a
pendulum oscillates in exactly the same way on one side of the room as
on the other, then it has translational symmetry. A mechanical system
that does not change in time has temporal symmetry.
265
266 8. CONSERVATION LAWS AND SYMMETRIES
1The
name comes from the Greek word for “hand.” The word “chiral” is
pronounced “kai-ral” and rhymes with “spiral.”
8.2. SYMMETRY AND THE LAWS OF PHYSICS 267
Therefore,
∂L
= constant.
∂ q̇
But recall that
∂L
= p = generalized momentum.
∂ q̇
Therefore, p = constant and the proposition is proved.
so
dL ∂L dq̇ ∂L dq̇ d ∂L
= − + q̇ .
dt ∂ q̇ dt ∂ q̇ dt dt ∂ q̇
Therefore,
dL d ∂L
= q̇ ,
dt dt ∂ q̇
or
d ∂L
L − q̇ = 0.
dt ∂ q̇
The definition of the Hamiltonian is
H = pq̇ − L,
∂L
where p = ∂ q̇
. Therefore the quantity in brackets is −H and
d
[−H] = 0.
dt
Consequently,
H = constant.
That is, for a system in which L is independent of the time, the Hamil-
tonian (and hence the energy) is constant.
8.5. Strangeness
Conservation laws are used in an unusual but interesting way by
high energy physicists. In studying the elementary particles, it is found
that some reactions never occur. There is no apparent reason why a
reaction such as
π− + p → π◦ + Λ
is never observed. It does not violate any of the everyday conserva-
tion laws. But it does not happen. Using the principle that “what is
not forbidden is required,” high energy physicists decided that there
must be a conservation principle at work. They called it “conserva-
tion of strangeness.” The way it works is this: Each particle is given
a “strangeness quantum number.” For example, the strangeness of a
π particle is 0, that of a proton is also 0, and the strangeness of a Λ
particle is -1. Note that the total strangeness on the right hand side
4However, relativistic quantum mechanics shows that CPT must be invariant
in any physical process. Here the letter T stands for the time reversal process. It is
interesting to note that CP violation was confirmed experimentally at the Stanford
Linear Accelerator Center (SLAC) in 2001 based on the observation of 32 million
decay events by a 1,200 ton detector. The violation of CP explains why the universe
is predominantly made of ordinary matter rather than being 50% anti-matter. (See
Colin Macilwain, “Physicists show what really matters,” Nature, 412, 105, 12 July
2001.)
8.5. STRANGENESS 273
of the reaction is not equal to the total strangeness on the left hand
side. Therefore, in this reaction, strangeness is not conserved and the
reaction is “forbidden.” This is not any weirder than requiring that
linear momentum be conserved during a collision. However, we feel at
home with momentum conservation because we have an analytical ex-
pression for momentum and we know that conservation of momentum
implies that no net external forces are acting on the system. We do
not have an analytical expression for strangeness nor do we know the
symmetry implied by strangeness conservation. We can be sure, how-
ever, that there is some symmetry in nature that requires strangeness
to be conserved.5
As an example of the importance of strangeness, Figure 8.1 is a
plot of strangeness vs charge for the eight spin 1/2 particles called
baryons. Note the neutron and proton on the top of the plot, both
having strangeness 0, but charge 0 for the neutron and +1 for the pro-
ton. (The lines of constant charge are skewed.) A plot just like this one
can be generated for the eight spin zero particles called mesons. These
plots (known as “The Eightfold Way”) are the basis of the standard
model of particle physics.6
n p
s=0
s=-1 Λ
Σ- Σ0 Σ+
s=-2
Ξ- Ξ0
5Ifyou wish to delve further into these questions, a good source is, Brehm, J.
and Mullin, W. J., Introduction to the Structure of Matter. John Wiley and Sons,
1989, New York, NY. Chapter 16.
6An interesting article on the history of the standard model by one of its
founders is Gerard ’t Hooft, “The making of the standard model”, Nature, 448,
271-273 (19 July 2007).
274 8. CONSERVATION LAWS AND SYMMETRIES
8.7. Summary
The point of this chapter was to introduce you to the idea that con-
servation laws are a consequence of symmetries in the physical universe.
We showed that momentum conservation is related to the homogeneity
of space (symmetry under translation) and energy conservation is a
consequence of temporal symmetry. The technique we used to demon-
strate this was to consider Lagrange’s equation when some coordinate
is ignorable. An ignorable coordinate in the Lagrangian implies a sym-
metry with respect to that variable.
Finally, we considered examples from elementary particle physics in
which the consequences of conservation laws are evident even if we do
not have a mathematical expression for the conservation law or even
know the form of the conserved quantity.
8.8. Problems
Problem 8.1. Consider a double pendulum (as in Chapter 4).
What physical quantities are conserved?
Problem 8.2. (a) Prove that if a certain coordinate qi does not
appear in the Lagrangian, it is also ignorable in the Hamiltonian. (b)
7For
a good discussion see Whitten, E., “When Symmetry Breaks Down,”
Nature, 429, 507-508, (3 June 2004).
8.8. PROBLEMS 275
Show that
∂H ∂L
=− .
∂qi ∂qi
Problem 8.3. Does the law
dL
N=
dt
exhibit symmetry under reflections? Explain.
Problem 8.4. Are your footprints chiral? Does the pattern formed
by your footprints when you are walking exhibit chirality? Would chi-
rality be exhibited if the pattern were of infinite length? (Hint: What
is a “glide reflection”?)
Problem 8.5. The “Baryon Decuplet” consists of four ∆ parti-
cles with charges from -1 to +2 and strangeness (s)=0, three Σ∗ with
charges from -1 to +1 and s=-1, two Ξ∗ with charges -1, 0 and s=-2,
and one Ω with charge -1 and s=-3. Plot the decuplet on a graph like
that of Figure 8.1.
Problem 8.6. The position vector r has components (1,2,3). Ex-
press the components of this vector after (a) an inversion and (b) a
reflection in the xy plane. (c) Let r0 be the vector obtained by re-
flecting r in a plane perpendicular to the unit vector n̂. Show that
r0 = r − 2(r · n̂)n̂.
Problem 8.7. An electrmagnet consists of a current carrying coil
wrapped around an iron core, as illustrated in Figure ??. (a) What
is the direction of B in the mirror image? (b) If a positive charge is
moving towards the mirror, it will feel a force F = qv × B directed out
of the page. What is the direction of the force in the reflected system?
276 8. CONSERVATION LAWS AND SYMMETRIES
B
q
v
Figure 8.2.
Part 2
m1 ^
r
r2- r1
r1
m2
r2
the negative sign in the equation indicates that the force is attractive.
By Newton’s third law, the force on particle m1 due to particle m2 is
r1 − r2
F12 = − Gm2 m1 .
|r2 − r1 |3
What is the force law for extended bodies? If the distance between
two bodies is much larger than the dimensions of the bodies (as for
astronomical objects), it is usually safe to assume that the bodies are
particles and the quantity |r2 − r1 | is the distance between the centers
of the two bodies. If an extended body cannot be treated as a particle,
you will need to determine the gravitational field of the extended body,
as discussed below in Section 9.3. (For example, if a satellite is in a
near-Earth orbit, the Earth cannot be considered a point mass.)
M = source point
r-r'
r'
P = field point
r
The location of body M is called the “source point,” and its position
is denoted by r0 . The point specified by r is called the “field point.” It
is the place where the field is evaluated. See Figure 9.2. Note that the
test mass does not enter into Equation (9.3), the expression for g(r).
284 9. THE GRAVITATIONAL FIELD
6Richard Feynman, QED: The Strange Theory of Light and Matter. Princeton
University Press, Princeton, NJ, 1985.
286 9. THE GRAVITATIONAL FIELD
e e
density may vary from one part of the body to another so it is written
as ρ = ρ(r0 ) to remind us that ρ is a function of position.
In Figure 9.4 the point P is the field point. Note that the source
point is at r0 and the field point is at r. The infinitesimal portion of
the gravitational field at r due to an infinitesimal mass element dm at
r0 is
r − r0
dg(r) = − Gdm .
|r − r0 |3
or
r − r0
dg(r) = − Gρ(r0 )dτ 0 ,
|r − r0 |3
where dτ 0 is so small that ρ(r0 )dτ 0 can be considered a particle, yet
large enough to treat the mass as continuous. To obtain the field at P
due to the whole body, we integrate over the body. That is,
0
0 r−r
Z
0
g(r) = −G ρ(r ) 0 3 dτ . (9.5)
body |r − r |
dm' = ρdτ'
r - r'
dg
P
r'
r
resolving the vector dg into its components before carrying out the
integration.
P
dg
θ d
dr
D
r
φr
force is conservative. (See Exercise 9.4 above.) Using the fact that the
curl of the gradient of any scalar function is zero, we conclude that
since ∇ × F = 0, then F can be expressed as the gradient of some
scalar function. That is, we can define the potential energy V = V (r)
such that
F = −∇V.
Then
Z r2 Z r2 Z 2
F · dr = − ∇V · dr = − dV = V1 − V2 .
r1 r1 1
where r is the field point and r0 is the location of the source point
dm = ρ(r0 )dτ 0 .
2π π
σa2 sin θdθdφ
Z Z Z
σdA
Φ = −G = −G ,
surface |r − r0 | φ=0 θ=0 |r − r0 |
Z 2
σa sin θdθ
Φ = −2πG ,
|r − r0 |
where the factor 2π came from integrating over φ. Using the law
of cosines, the figure shows that since |r0 | = a,
√
|r − r0 | = R = a2 + r2 − 2ar cos θ,
and R2 = a2 + r2 − 2ar cos θ. Therefore 2RdR = 2ar sin θdθ and
the integral takes the form
Z π 2 Z R=r+a
σa sin θdθ 2 (R/ar)dR
Φ = −2πG = −2πGσa
0 R R=r−a R
Z r+a
−2πGσa
= −Gσa2 (2π/ar) dR = [(r + a) − (r − a)]
r−a r
4πa2 σ GM
= −G =− ,
r r
where we used the fact that as θ goes from 0 to π, R ranges from
r − a to r + a.
Our result shows that for points outside the shell, the potential
of a shell of mass M is the same as the potential of a point mass
M at the origin. As a problem you can show that the field inside
the shell is zero.
of lines per unit area falls off as 1/r2 , just like the gravitational field.
The lines representing the field are traditionally called “lines of force”
although most modern books refer to them with the better terminology
of “field lines.”
Another graphical representation of the field consists in sketching
the surfaces on which the potential is constant. For a point mass,
the equipotential surfaces are concentric spheres centered on the mass.
As we will show in a moment, the field and the potential are related
by g = −∇Φ, and you know from Chapter 5 that the gradient of
Φ is perpendicular to the surfaces of constant Φ. Therefore, the two
representations (field lines and equipotential surfaces) are equivalent.
^
n dS
S ^
n
dS
θ
r r
M dS
7The solid angle is the three dimensional angle formed at the vertex of a cone.
Let point O be at the origin and a surface of area dA be a distance r from the
origin. Then, if dA is perpendicular to r, the solid angle subtended by dA at O is
dΩ = dA/r2 . The units of solid angle are steradians. The solid angle subtended by
a spherical shell at a point inside the shell is 4π steradians.
296 9. THE GRAVITATIONAL FIELD
of superposition,
I X X
g · n̂ dS = − 4πGmi = −4πG mi = −4πGMenc , (9.12)
S
where Menc is the total mass enclosed by S. The equation
I
g · n̂ dS = −4πGMenc
S
is called Gauss’s Law. It is particularly useful in the study of electro-
statics.
Assume mass is continuously distributed in a volume V bounded
by a closedR surface S. The element of mass is ρdτ , so the mass enclosed
is Menc = V ρdτ. Therefore,
I Z
g · n̂ dS = − 4πGρdτ. (9.13)
S V
You may recall from your vector analysis course that Gauss’s divergence
theorem states that for any vector A,
I Z
A · n̂ dS = ∇ · A dτ. (9.14)
S V
Therefore,
R the left hand side of equation (9.13) can be expressed as
V
∇ · gdτ, and consequently
Z Z
∇ · g dτ = − 4πGρdτ,
V V
or
∇ · g = − 4πGρ. (9.15)
Thus, the divergence of g is proportional to ρ, the mass density. An
important theorem called the Helmholtz theorem states that for any
vector field, if one knows the divergence and the curl of the field, then
one can determine the field itself. For this reason, ∇ · g and ∇ × g are
called the “sources” of g. Equations (9.10) and (9.15) indicate that the
source of the gravitational field is the mass density ρ. In other words,
masses generate gravitational fields.
Worked Example 9.4. Use Gauss’s law to determine the
gravitational field outside of an infinitely long cylinder of radius a
with constant linear mass density λ.
Solution: To solve we construct a Gaussian surface which is a
closed surface that has the same symmetry as the mass distribution.
We select this surface so that on the surface g · n̂ is either constant
or zero. The field point lies at an arbitrary point on this surface.
9.6. THE NEWTONIAN GRAVITATIONAL FIELD EQUATIONS 297
P = field point
r
a
There are two things to note about this expression. First of all it is
customary to use upper and lower indices which are called “contravari-
ant” and “covariant”. The difference need not concern us at present,
but it is important to appreciate that dxµ does NOT mean dx raised to
the power µ! It is just a way to denote the component9. Secondly, note
9Superscript
indices are denoted contravariant and subscript indices are de-
noted covariant. The difference between covariant and contravariant vectors is the
way they transform. The components of a covariant vector transform according
9.8. EINSTEIN’S THEORY OF GRAVITATION (OPTIONAL) 301
that the indices µ and ν range from 0 to 3. Here 1,2,3 represent the
spatial coordinates, dx, dy, dz and zero denotes the “time” coordinate
cdt. Thus, dx3 represents dz and dx0 represents cdt. The quantity ds is
the “line element” in 4-dimensional “spacetime.” In special relativity,
the metric ηµν is quite simple. It is called the “Minkowski metric10”
and in matrix form, it can be written as
1 0 0 0
0 −1 0 0
ηµν = 0 0 −1 0
0 0 0 −1
Now Minkowski spacetime is described as “flat” because Euclidean
geometry is valid in this space. This means, for example, that parallel
lines never meet and that the angles in a triangle add up to 180 degrees.
Note that the geometry of the surface of a sphere is non-Euclidean.
Parallel lines that start out perpendicular to the equator, meet at the
pole. The angles of a triangle may not sum to 180 degrees. In 3-
dimensional space, the line element ds on a flat surface is
ds2 = dx2 + dy 2 + dz 2 ,
but on the surface of a sphere of radius R,
ds2 = R2 dθ2 + R2 sin2 θdφ2 .
Not all surfaces that we might think of as curved are curved in the
mathematical sense. For example, the curved side of a cylinder is
mathematically flat because (as you can easily show) parallel lines never
meet, triangle have 180 degrees, etc.
Bernhard Riemann11 used the concept of line element to generalize
geometry. He expressed the generalized line element in a curved 3-
dimensional space as
3
X
2
ds = gij dxi dxj ,
i,j=1
∂xj
to A0i = ∂x0j Aj , whereas the components of a contravariant vector transform as
∂x0j
A0i = ∂xj Aj .
10This metric was developed by Hermann Minkowski who was one of Einstein’s
mathematics professors. Minkowski is credited for having developed the concept of
spacetime.
11Bernhard Riemann (1826-1866), who was a student of Carl Friedrich Gauss, is
considered one of the worlds greatest mathematicians. Among other achievements,
he developed the geometry of curved surfaces.
302 9. THE GRAVITATIONAL FIELD
where the gij are the elements of the metric tensor (or “metric coef-
ficients”). The metric relates the coordinate differentials (the dxi ) to
a length ds in the space under consideration. Therefore, the metric is
related to the geometry of the space. Once the metric is known, the ge-
ometry of the space is entirely determined. As we shall see, the metric
can not only tell us whether the space is flat or curved, it can actually
give us the curvature of the space.12 In fact the curvature is given by
a complicated quantity called the Riemann tensor that is defined as
l ∂Γlik ∂Γlij X m l X
Rijk ≡ − + Γik Γmj − Γm l
ij Γmk
∂xj ∂xk m m
where
1 X il ∂glk ∂gjl ∂gjk
Γljk = g + k + .
2 l ∂xj ∂x ∂xl
The Riemann tensor is rank 4 (as you can tell by noting that there
are four indices) and Γlmk is a rank 3 tensor. Obviously, this is a very
complicated quantity. You will probably never have to evaluate it un-
less you become a theoretical physicist specializing in general relativity!
l
Nevertheless, it is worth noting that for a flat space, Rijk = 0 every-
where. Now so far we have been considering the geometry of ordinary
3 dimensional space (as indicated by our use of latin letters as indices).
Einstein’s job was to generalize to 4 dimensional spacetime. Thus, in
a curved Riemannian 3-D space,
3
X
ds2 = gij dxi dxj ,
i,j=1
where ds2 is positive and the gij are functions of the coordinates,
whereas in a flat 4-D Minkowski spacetime
3
X
2
ds = ηµν dxµ dxν
µ,ν=0
where the ηµν are constants equal to (1, −1, −1, −1) and ds2 may be
positive, negative or zero. In order to treat curved spacetime, the
constant ηµν are replaced by the functional quantities gµν . That is, in
12It
might be mentioned that most of the metrics that are encountered in
physics are orthogonal, and as we have seen in Chapter 5, this means the metric
tensor can be expressed as a diagonal matrix.
9.8. EINSTEIN’S THEORY OF GRAVITATION (OPTIONAL) 303
“Newtonian limit” this tensor reduces to the left hand side of Poisson’s
equation. It is, of course, a sign of Einstein’s genius that he realized
this correspondence and expressed the Einstein tensor as proportional
to the energy-momentum tensor. It turns out that the proportionality
constant is K = 8πG/c2 where G is the universal gravitational constant
in Newton’s law of gravity. Consequently,
8πG
Gµν = − 2 T µν . (9.18)
c
This is, essentially, the Einstein expression for the gravitational field. It
is a set of 10 coupled second order differential equations for the metric
g µν and is referred to as “Einstein’s field equations.”13
We still have not answered the question about how a mass particle
will move in curved spacetime. Recall our discussion of the calculus of
variations in Chapter 4 in which we showed that the shortest distance
between two points in a plane is a straight line and the shortest distance
between two points on the surface of a sphere is a section of a great
circle (or geodesic). It is an interesting consequence of the general rela-
tivity that a particle in a gravitational field will move along a geodesic.
At first Einstein thought that this would have to be incorporated into
his theory as a postulate, but some years after he had published his
theory he realized that geodesic motion follows from the condition
X
∇µ T µν = 0.
µ
9.9. Summary
A field is defined as a physical quantity whose value can be deter-
mined at every point in some given region of space. The gravitational
field g(r) due to a particle of mass M located at point r0 can be defined
in terms of the force it exerts on an infinitesimal point mass m located
9.9. SUMMARY 307
at r,
r − r0
F
g(r)= lim = −GM .
m→0 m |r − r0 |3
The field produced by an extended body can be determined by evalu-
ating the expression
r − r0
Z
g(r) = −G ρ(r0 ) 0 |3
dτ 0 .
body |r − r
To evaluate the integral it is usually necessary to resolve the vectors
into components, using the symmetry of the problem as a guide.
The gravitational potential Φ(r) is given by
ρ(r0 )dτ 0
Z
Φ(r) = −G 0
.
body |r − r |
9.10. Problems
Problem 9.1. A neutron star has a mass of 1030 kg and a radius
of 5 kilometers. A body is dropped from a height of 20 cm above the
surface. Determine the speed of the body when it hits the surface.
Problem 9.2. The Earth is suddenly brought to a standstill. Eval-
uate the time required for it to collide with the Sun. You can assume
the Sun does not move and that the center of mass of the system is at
the center of the Sun. How good is this approximation?
Problem 9.3. Determine the period of a surface skimming satellite
in a circular orbit about a uniform, perfectly spherical planet of radius
R and density ρ. Determine the period of such a satellite about a
different planet which has radius 2R but the same density as the first.
Explain your result.
Problem 9.4. Consider an infinitely long, straight string whose
linear mass density is λ (mass per unit length). By direct integration,
determine the gravitational field a distance r from the string.
Problem 9.5. By direct integration, find the field at a point on
the axis of symmetry of a cylinder of length L, radius R and uniform
density ρ. (Hint: Let the mass element be an infinitesimally thin disk
and use the result of Worked Example 9.2.)
Problem 9.6. By direct integration determine the gravitational
field a distance z above an infinite flat surface whose mass per unit
area is σ. Check your result by using the result of Worked Example
9.2,
Problem 9.7. Consider an infinite string with linear mass density
λ. A particle of mass m is a distance d from the string. Determine the
force on the particle.
Problem 9.8. Using the technique of Worked Example 9.3, de-
termine the potential and the field of a spherical shell of mass M and
radius a, at an interior point. Answer: g = 0 and Φ = −GM/a =
constant.
Problem 9.9. Assuming that the mass cylinder of Worked Exam-
ple 9.4 has a constant mass density ρ, determine the field at a point
inside the cylinder.
Problem 9.10. A sphere of radius R and constant mass density
ρ has a spherical cavity of radius r where r = R/2. The center of the
cavity is a distance R/4 from the center of the sphere. A particle of
9.10. PROBLEMS 309
there is no net force acting on the particle at any interior point. (This
is the way Newton showed that there was no force on a particle inside
a shell.)
A2
P
r2
r1
A1
Problem 9.17. Use Gauss’s law to determine the field inside and
outside of: (a) A sphere of uniform mass density, (b) A homogeneous
hollow spherical shell. Let the mass and radius be M and R in both
cases.
Problem 9.18. Use Gauss’s law to determine the field inside and
outside an infinite cylinder of radius R and uniform mass density ρ.
Express your answer in terms of r, the distance from the axis.
Problem 9.19. Use Gauss’s law to determine the field above and
below an infinite plane of surface mass density σ.
Problem 9.20. An infinite mass plane gives rise to a constant
gravitational field, and a sphere of mass M gives rise to the gravita-
tional field given by Newton’s law of universal gravitation. Consider
an object that looks like Saturn, but instead of rings, the equatorial
plane of the sphere is a uniform mass plane of infinite extent and mass
density σ (per unit area). Let the radius of the sphere be R. Deter-
mine the work done as a particle of mass m is moved from the “north
pole” of the sphere to a distance h above it (h > R). (The principle
of superposition tells us the field of two mass distributions is just the
vector sum of the field due to each.)
Problem 9.21. The gravitational field in some region of space is
given by g = −kr3 r̂ where k is a constant. What is the mass density
ρ? What is ρ if the field is given by g = −(k/r2 )r̂?
9.10. PROBLEMS 311
COMPUTATIONAL PROJECTS
14E. T. Lu and S. G. Love, Gravitational Tractor for Towing Asteroids, Nature,
438, 177-178 (2005).
312 9. THE GRAVITATIONAL FIELD
The orbital motion of a planet around the Sun was one of the first
important problems to be analyzed in terms of Newton’s three laws.
The gravitational force attracting a planet to the Sun is a central force.
The motion of a planet is a prime example of the more general problem
of the behavior of a particle acted upon by a central force.
Although we shall be primarily concerned with the motion of plan-
ets and satellites, the techniques you will learn in this chapter are ap-
plicable to any kind of central force. In this chapter, as well as learning
the laws governing the motion of celestial bodies, you will be exposed
to the concept of effective potential energy, and you will appreciate
how constants of the motion are used in solving physics problems.
Historically, the quantitative analysis of orbital motion began with
Kepler’s realization that the motion of planets can be described by
three empirical laws. An important point made in this chapter is that
Kepler’s laws of planetary motion can be derived theoretically from
Newton’s laws of motion. Additionally, Newton’s laws gives us a much
deeper understanding of Kepler’s laws. This application of Newton’s
ideas amazed and fascinated the “natural philosophers” of his era and
was one of the most important events in the history of science.1
Kepler went to visit Brahe to get his data. He was certain that Brahe
would be overwhelmed by his wonderful new theory. Brahe was not
overwhelmed. In fact, at first Brahe would not even let Kepler see
the data! However, Brahe soon realized that Kepler was an excellent
mathematician and he offered to let him work on a small portion of
the data to calculate the orbit of Mars. This was certainly not what
Kepler had in mind, but he grudgingly agreed.
This was, perhaps, the original “graduate student - professor” re-
lationship in science. To this day the same pattern exists. A young,
aspiring scientist makes the pilgrimage to the laboratory of the estab-
lished professor with the hope of being allowed to share in the profes-
sor’s knowledge and data. You yourself may be doing this a few years
from now. I hope your relationship with your graduate advisor will less
tempestuous than that of Kepler and Tycho. They did not get along
at all. Tycho was a man who loved a party, who spent his evenings
eating, drinking, and carousing, whereas Kepler was somber and rather
puritanical and did not at all approve of Tycho’s lifestyle.
Some years after the collaboration began, Tycho died. (Rumor has
it that he died of a burst bladder while on a drinking spree.) Kepler
inherited all of Tycho’s data as well as Tycho’s position as court as-
tronomer. After a great deal of analysis, much to his dismay, Kepler
found that Tycho’s data did not support his grandiose theory. In fact,
the data showed that the orbits of planets were not even circles, but
rather ellipses! Kepler tried to modify his theory by slipping elliptical
orbits between the inscribed perfect solids, but it did not quite work.
Kepler never did know that there are more than five planets, as the
discoveries of the planets Uranus and Neptune came many years after
his death. By the time these were discovered, Kepler’s theory on the
inscribed orbits of the five planets was no more than a historical oddity.
Kepler’s life is full of instructive incidents for physicists. His most
valuable contribution to science was his analysis and synthesis of the
observations of Tycho Brahe. It is interesting to consider that he was
obsessed by a beautiful theory that did not agree with experiment.
No matter how beautiful a theory may be, if it does not agree with
experimental measurements, it must be discarded! As a physicist you
must never let your theories carry you away. Physics is the study of
the physical universe and it is Nature that determines the way things
behave. It was to Kepler’s great credit that he respected and believed
the data, even though the data did not agree with his theory.
10.3. CENTRAL FORCES 321
mass of this system lies at the center of the Earth and we are perfectly
justified in considering the Earth to be at rest.4
Placing the origin of the coordinate system at the center of mass
(essentially the center of the Sun) the force on a planet is an attrac-
tive force directed toward the origin. The magnitude of this force is
inversely proportional to the square of the distance from the planet to
the origin. Any force that is directed toward or away from a fixed point
(usually taken as the origin of coordinates) and whose magnitude is a
function only of the distance to the origin is called a central force. In
general, a central force will have the form
F = f (r)r̂,
where f (r) is the magnitude of the force (a function only of r) and r̂
is the radial unit vector. The direction of the force is along the line
joining the particle and the origin. Two important central forces are the
gravitational force between masses and the electrostatic force between
charges. For the gravitational force,
Gm1 m2
F=− r̂,
r2
and for the electrostatic force
Q1 Q2
F= r̂.
4πo r2
We can also imagine other central forces such as, for example,
k
F = 5 r̂.
r
Such forces may or may not exist in nature, but we can analyze them
mathematically anyway. Although this seems like a useless theoretical
exercise, such studies may have a practical outcome. For example, to
a first approximation, the force between molecules can be expressed
as the sum of two central forces. This “Lennard-Jones” force can be
expressed in terms of the potential energy as
a b
V = − 6 + 12 .
r r
Here a and b are constants that depend on the properties of the partic-
ular molecules involved in the interaction. The force itself is obtained
from F = −∇V.
4Recallthe discussion of reduced mass, defined as µ = m1 m2 /(m1 + m2 ). See
Equation (6.20). If the central force problem is treated in terms of the reduced
mass, then no approximation is made. The relations obtained are the same as
those derived here except that m is replaced by µ and r is the position of one body
relative to the other rather than the distance from the center of mass.
10.3. CENTRAL FORCES 323
v
r
Figure 10.2. The angular momentum vector is perpen-
dicular to the plane containing r and v. Since the angular
momentum is constant, the plane is invariant.
Since the motion of the particle lies in a plane, two coordinates are
sufficient to specify its position. These can be x and y or r and θ.
The origin of the coordinate system is usually placed at the primary
(assumed to be at rest).5 To orient the coordinates, it is necessary to
specify a fixed direction. Astronomers pick an imaginary line from the
center of the Earth towards a position called “The First Point in Aries”
which is the position of the Sun at the vernal equinox.
In Figure 10.3 the symbol Υ indicates the fixed line in space. In
polar coordinates the angle θ is measured from Υ. In Cartesian coor-
dinates one usually defines the x-axis along this fixed line. The y-axis
is selected in the plane of the orbit and perpendicular to the x-axis.
The z-axis is then perpendicular to the orbit and along the angular
momentum vector. For many problems it is quite safe to assume that
the coordinate system is an inertial system.
5In
the language of astronomy, if one body is much more massive than the
other, the massive body is called the “primary.”The primary is often assumed to
be at rest. Actually, of course, both bodies move about the center of mass.
10.3. CENTRAL FORCES 325
r
θ
Five thousand years ago the position of the Sun at noon on the
vernal equinox was in the constellation Aries (“the ram”). Since the
Earth’s axis of rotation precesses with a period of about 25,000 years,
the position of the Sun at noon on the vernal equinox has changed and
it is presently in the constellation Pisces; within a few hundred years
it will enter into the constellation Aquarius. However, the name “First
Point in Aries” and the symbol Υ, representing ram’s horns, are still
used to represent this arbitrary fixed line in space. It is amusing to
note that due to the precession of the Earth’s axis, the positions of
the zodiacal constellations have shifted, but astrologers still use the
values of 5000 years ago. Thus, people born when the Sun was in the
constellation Pisces think they are Aries, those who were born when
the Sun was in Aries think they are Taurus, and so on. If astrology
had any validity, this horrible mix-up in the zodiacal signs would be
serious indeed!
Therefore, for an inverse cube force law, mr̈ = −f (r)r̂ yields two
relations, namely
k
−mrθ̇2 = −f (r) = − 3 ,
r
and
mrθ̈ = 0.
The second equation boils down to θ̈ = 0, telling us that the angular
velocity, θ̇ is constant. The first equation yields
mr4 θ˙2 = k.
But since l = mr2 θ̇ we see that the left hand side is l2 /m and
consequently
l2 = mk,
and hence the angular momentum is constant. Actually, this result
also follows from r = constant and θ̇ = constant.
But
dG d
= (p · r) = p · ṙ + ṗ · r
dt dt
= mv · v + F · r =2T − ∇V ·r
dV
= 2T − r,
dr
because V = V (r). Therefore,
1 τ
Z
dG dV
= 2T − r dt = 0,
dt τ 0 dr
and
dV
2 hT i − r = 0,
dr
1 dV
hT i = r .
2 dr
Since V = krn+1 we have
dV
r = (n + 1)krn r = (n + 1)krn+1 = (n + 1)V,
dr
and
dV
r = (n + 1) hV i ,
dr
so
n+1
hT i = hV i .
2
The gravitational potential has the form V = − kr so n = −2
and
1
hT i = − hV i .
2
This relationship between kinetic and potential energy is useful
when solving orbital mechanics problems. It might be mentioned
that the virial theorem has important applications in thermody-
namics and statistical mechanics.
d 2
0 = mr θ̇ = m 2rṙθ̇ + r2 θ̈ ,
dt
= mr 2ṙθ̇ + rθ̈ .
Since neither m nor r is equal to zero, this implies that
rθ̈ + 2ṙθ̇ = 0.
But this is just the θ equation! Therefore, we see that the θ component
of the equation of motion (the second of equations 10.2) is a statement
that angular momentum is constant. Consequently, that equation can
be replaced by the equivalent equation
l = mr2 θ̇ = constant.
This equation, in turn, gives a nice expression for θ̇, namely,
l
θ̇ = . (10.4)
mr2
Replacing θ̇ in the r equation (the first of Equations 10.2) by ex-
pression (10.4) yields the following equation for the radial motion of
mass m:
l2 GM
r̈ − 2 3 = − 2 . (10.5)
mr r
This equation involves only r. Thus, conservation of angular momentum
de-couples the equations of motion. The equation has only one variable
(r), so it is often called a “one-dimensional equation.” But you should
always keep in mind that the motion takes place in two dimensions.
330 10. CENTRAL FORCE MOTION: THE KEPLER PROBLEM
Once Equation (10.5) has been solved for r = r(t), you can use it in
Equation (10.4) to determine θ = θ(t).
We shall now use the Lagrangian technique to determine the equa-
tions of motion. (We had better obtain the same result!)
As you know, the Lagrangian is L = T − V. In this problem V is
the potential energy for a particle of mass m attracted gravitationally
to a body of mass M . According to Equation (9.6) this is
GM m
V (r) = − .
r
The kinetic energy for this two-dimensional problem is T = 12 m(ẋ2 +
ẏ 2 ), or in polar coordinates,
1 1
T = mṙ2 + mr2 θ̇2 .
2 2
Therefore the Lagrangian is
1 1 GM m
L = mṙ2 + mr2 θ̇2 + .
2 2 r
Recall that the Lagrange equations of motion have the form
d ∂L ∂L
− = 0,
dt ∂ q̇i ∂qi
where the qi ’s are now r and θ. So we have two equations, namely,
d ∂L ∂L
− = 0,
dt ∂ ṙ ∂r
and
d ∂L ∂L
− = 0.
dt ∂ θ̇ ∂θ
The partial derivatives are easily evaluated. You should prove for your-
self that the two equations of motion are
d GM m
(mṙ) − mrθ̇2 + = 0, (10.6)
dt r2
and
d 2 dl
mr θ̇ = 0, or = 0. (10.7)
dt dt
The second of these equations gives
l
θ̇ = .
mr2
Using this expression, Equation (10.6) leads to
l2 GM
r̈ − 2 3
=− 2 . (10.8)
mr r
10.5. ENERGY AND THE EFFECTIVE POTENTIAL ENERGY 331
1
Exercise 10.4. Carry out the steps to show that T = 2
mṙ2 +
1
2
mr2 θ̇2 .
E = constant = T + V
1 Mm
= m(ṙ2 + r2 θ̇2 ) − G .
2 r
with the remaining two terms. The effective potential Vef f is given by6
l2 GM m
Vef f = 2
− . (10.10)
2mr r
Although Vef f looks and acts like a potential energy and is a function
only of position, it is definitely not a potential energy since it actually
contains a kinetic energy term, namely, l2 / (2mr2 ) = 21 mr2 θ̇2 .
It is instructive to draw an energy diagram in terms of the effective
potential energy.7 Note that Vef f is the sum of two terms, one positive
and the other negative. For r → ∞, the negative term in Vef f is the
dominant term because
1 1
> 2 .
r r→∞ r r→∞
As r → 0, the positive term dominates because
1 1
< 2 .
r r→0 r r→0
Therefore the plot of Vef f vs r must have the general shape shown
in Figure 10.4. Study this figure carefully and convince yourself it is
qualitatively correct, specifically that Vef f is positive as r → 0 and
negative as r → ∞. Notice particularly that the l2 /2mr2 term bends
much more sharply than −GM m/r. Also, keep in mind that r can
only take on positive values.
Figure 10.5 is also a plot of Vef f (r) vs r. In this plot the effective
potential energy does not have quite the right shape because I drew it
to make it easy for you to appreciate various aspects of the effective
potential that are hard to see on a more accurate plot, such as Figure
10.4. In Figure 10.5 you see four possible values for the total energy,
denoted E0, E1 , E2 , and E3 . Consider first a particle with energy E1 .
From Equations (10.9) and (10.10), we can write
1 2
mṙ = E − Vef f . (10.11)
2
6Theterms “effective potential energy” and “effective potential” are used inter-
changeably, even though a “potential” is actually potential energy per unit mass.
Some books use the term “fictitious potential.” An older term is “centrifugal po-
tential.” In general, the effective potential is defined as the sum of l2 /2mr2 and
the potential energy. Thus for an electron orbiting about a proton, the effective
potential would be
l2 e2
Vef f = − .
2mr2 4π0 r
7Atthis time you may wish to review the material on energy diagrams in
Section 5.6.
10.5. ENERGY AND THE EFFECTIVE POTENTIAL ENERGY 333
100
80
40
Veff l2 /2mr2
20
-20
-40
-60
-GMm/r
-80
-100
0 1 2 3 4 5 6 7 8 9 10
r (arbitrary units)
You should keep in mind that this not a complete description of the
motion; it is only a description of the radial motion. Meanwhile the
particle is also moving in θ with a velocity given by θ̇ = l/mr2 . The
334 10. CENTRAL FORCE MOTION: THE KEPLER PROBLEM
Going back to the energy diagram, Figure 10.5, note that according
to Equation 10.11 the radial component of the velocity ṙ is given by
r
2
ṙ = [E − Vef f (r)]. (10.12)
m
Thus, the square of the radial speed is proportional to E − Vef f . (Note
that E − Vef f is not the kinetic energy. You may think of it, if you
like, as the “radial kinetic energy.”) On the energy diagram (Figure
10.5) the distance between the horizontal line at E1 and the heavy line
representing Vef f is proportional to ṙ2 . At r1 the value of Vef f is E1 ,
so ṙ = 0. That is, at the turning point the particle has zero radial
velocity, which makes perfect sense. At the turning point, the angular
component of the velocity is a maximum because rθ̇ = (r)(l/mr2 ) =
l/mr is greatest when r is smallest.
Next consider a particle with zero total energy (E = E0 = 0; see
Figure 10.5). This means that the positive “radial kinetic energy” 21 mṙ2
is equal in magnitude to the negative effective potential energy Vef f .
The motion of the particle as it comes from r = ∞ to the turning point
at r0 and then goes back out to r = ∞ is similar to the motion of the
particle with energy E1 . As we shall see, the main difference is that
the trajectory for energy E1 > 0 is a hyperbola and the trajectory for
energy E0 = 0 is a parabola. For the parabolic orbit, the radial speed
of the particle ṙ is zero at infinity as well as at the turning point. As
the particle comes in from infinity, ṙ increases, reaching a maximum
at r4 where Vef f reaches is greatest negative value, then slows down to
10.5. ENERGY AND THE EFFECTIVE POTENTIAL ENERGY 335
Circle
Ellipse
Hyperbola Parabola
Figure 10.7. The conic sections. If the cone is cut as
shown, the cross sections of the cuts will be a circle, an
ellipse, a parabola, and one branch of a hyperbola.
Integrating gives
t = t(r, r0 , E, l),
which can, in principle, be inverted to yield
r = r(t, r0 , E, l).
In this solution, the total energy E and angular momentum l are ar-
bitrary constants. Note the use of the term arbitrary constant. Such
constants are arbitrary in the sense that the differential equation is
satisfied by the solution no matter what values the constants happen
to have. But for a particular problem, these constants are anything but
arbitrary! In simple kinematics problems the “arbitrary constants” are
usually the initial position and initial velocity. In more complex prob-
lems such as the one we just solved, they tend to be other constants of
the motion such as the energy and angular momentum.
where I used the Greek letters α, β, and γ to illustrate the form of this
equation. Rewriting and integrating the last equation gives
Z r Z θ
dr
2 1/2
= dθ. (10.15)
r0 r(α + βr + γr ) θ0
Sophisticated Technique
There is a different way to obtain the equation of the orbit that
cleverly avoids evaluating the complicated integral in (10.15). Bear
with me for a little while because the math is a bit involved (but not
difficult).
We will begin with Equation (10.2), the radial equation of motion:
GM
r̈ − rθ̇2 = − .
r2
10.7. THE EQUATION OF THE ORBIT 339
dr 1 du 1 du dθ
=− 2 =− 2 ,
dt u dt u dθ dt
where the last step uses the chain rule and the fact that u = u(θ).
So,
dr 1 du du
= − 2 θ̇ = −r2 θ̇ .
dt u dθ dθ
2
But r θ̇ = l/m. Replacing,
dr l du
=− .
dt m dθ
Taking the derivative with respect to time again,
d dr d l du
= − ,
dt dt dt m dθ
l d du l d du dθ
r̈ = − =− ,
m dt dθ m dθ dθ dt
l d2 u l 2 d2 u
l
= − = − .
m dθ2 mr2 m2 r2 dθ2
But 1/r2 = u2 so
l 2 u 2 d2 u
r̈ = − .
m2 dθ2
Substituting this into Equation (10.8) and using 1/r3 = u3 we have
l2 u2 d2 u l2 2
− − u = −GM u2 ,
m dθ2 m2
or
d2 u m2 l2 u3
2
=− 2 2 − GM u .
dθ2 l u m2
So,
d2 u GM m2
+ u = . (10.16)
dθ2 l2
340 10. CENTRAL FORCE MOTION: THE KEPLER PROBLEM
This is a very interesting equation. It looks like the equation for simple
harmonic motion except for the additional constant term on the right
hand side. It can be made to look exactly like the simple harmonic
motion equation by defining a new variable
GM m2
w =u− .
l2
Then, since the last term is a constant,
dw du
= ,
dθ dθ
and
d2 u d2 w
= .
d2 θ dθ2
Equation (10.16) can now be written in the form of the SHM equation:
d2 w
+ w = 0.
dθ2
As noted in Worked Example (3.5), the solution is sinusoidal and we
can write
w = A cos(θ − θ0 ).
Consequently,
GM m2
u− = A cos(θ − θ0 ).
l2
Now u is just the inverse of r, so the derivation finally gives an equation
for r in terms of θ :
1 1
r = = GM m2 .
u l 2 + A cos(θ − θ 0 )
Dividing top and bottom by GM m2 /l2 puts this in a nicer form:
l2 /GM m2
r= Al2
. (10.17)
1 + GM m2
cos(θ − θ0 )
From a mathematical point of view, the problem is now solved
because r has been expressed in terms of θ and other known quantities
(such as the angular momentum) and two constants of integration, A
and θ0 .
Equation (10.17) describes all the possible orbits for the two-body
problem, i.e., circles, ellipses, parabolas, and hyperbolas. Since ellipti-
cal motion is of particular interest, let us assume that the total energy
E is negative. Then the motion is bounded and the particle (planet)
oscillates radially between turning points r2 and r3 as illustrated in
Figure 10.5 for energy E2 .
10.7. THE EQUATION OF THE ORBIT 341
l2 GM m
VefT Pf = 2
− = E.
2mrtp rtp
Therefore, in terms of u ,
l2 2
u − GM mutp − E = 0.
2m tp
The two solutions of this quadratic equation for utp are:
r !
m 4l 2E
u± = 2 GM m ± (GM m)2 + . (10.18)
l 2m
(Note that 1/u± are the r2 and r3 of Figure 10.5.)
But we have seen that
GM m2
u = A cos(θ − θ0 ) + .
l2
The maximum and minimum values of this expression occur when
cos(θ − θ0 ) = ±1. That is,
GM m2
u+ = A + ,
l2
and
GM m2
u− = −A + .
l2
Using u+ and equating the expression above to the relation given by
Equation (10.18) leads to
r
GM m2 GM m2 m 2+
2l2 E
A+ = + (GM m) ,
l2 l2 l2 m
or
1
(GM m)2 m2 2Em 2
A= + 2 . (10.19)
l4 l
Thus A is related to the total energy and the angular momentum in a
rather complicated way.
Let us now turn our attention to θ0 , the other constant of in-
tegration. Since the maximum and minimum values of r occur at
cos(θ − θ0 ) = ±1, it is common to set θ0 = 0 and measure angles
from the x-axis. Then the initial time, t = 0, is the time the planet
passes through perihelion - the point of closest approach to the Sun.
342 10. CENTRAL FORCE MOTION: THE KEPLER PROBLEM
See Figure 10.8. More generally, the angle θ is measured from some
fixed line in space. This would be appropriate if the ellipse is precessing
and the x-axis is rotating relative to the inertial frame. In that case
the angle θ0 gives the angle between the fixed line and the major axis
of the elliptical orbit at the initial time.
x
θ θ0
θ x
θ0=0
r'
r
P' θ P
F' f f F
a
Figure 10.10. An ellipse. The total distance r + r0 +
F 0 F is a constant, so r + r0 must be constant. Points P 0
and P are called the apsides.
Consequently
GM m2 a(e2 − 1) 2 1
2 2 1
v = + = GM + .
m2 a(e2 − 1) r a r a
At r = ∞, the velocity is v0 and we see that
GM
v02 = ,
a
so
GM
a= 2 .
v0
Also
l2
a(e2 − 1) = .
GM m2
Solving for e we obtain
2 !1/2
v0 b
e= 1+ ,
(GM )
and finally
v02 b2 /GM
r= r 2 2 .
bv0
1+ 1 + GM cos θ
P
b
ea ea
F' 0 F
a
d
c
b
dr
r
r +d
r
Exercise 10.15. (a)Using the form τ 2 = (4π 2 /GM )a3 and assum-
ing you do not know the mass of the Sun, determine the semimajor
axis of Saturn in AU, given its period is 29.5 Earth years. (b)Using
the correction (M + m) and looking up the appropriate values, obtain
a corrected value. Answers: 9.55 AU, 9.52 AU
r = a = constant.
Consequently,
r̈ = 0.
The radial equation of motion (Equation 10.31) then reduces to
l2
f (a) = − . (10.32)
ma3
I will come back to this equation shortly.
Now consider the question of the stability of a perturbed circular
orbit. For example, we might be considering the motion of a planet in a
10.10. A PERTURBED CIRCULAR ORBIT 357
perfectly circular orbit which is hit by a comet (as happened some years
ago when comet Shoemaker-Levy collided with Jupiter). We want to
know if the planet will continue moving in a stable orbit after a slight
perturbation, or if it will behave in an erratic manner. In other words,
we would like to know if a collision with a relatively small body could
cause Jupiter to go flying out of the solar system.
After a collision or some other sort of perturbation, the radial po-
sition is no longer exactly equal to a, but it is still nearly equal to a,
so we can write
r = a + η, (10.33)
where η << a, i.e., η is a very small quantity.
Inserting expression (10.33) for r into the radial equation of motion
in the form (10.31) we obtain
d2 l2
m (a + η) = f (a + η) + ,
dt2 m(a + η)3
or
l2
mη̈ = f (a + η) + . (10.34)
ma3 (1 + ηa )3
The force at the position r = a + η is very nearly equal to the force
at r = a so we are justified in expanding f (a + η) in a Taylor’s series
expansion and keeping only the leading terms. Thus:
1 2 d2 f
df
f (a + η) = f (a) + η + η + · · ·.
dr r=a 2 dr2 r=a
The term (1 + η/a)3 in the denominator of the last term of Equation
(10.34) can also be expanded. Using the binomial expansion:9
1 η −3 η η 2
= 1 + = 1 − 3 + 6 + · · ·.
(1 + η/a)3 a a a
Consequently, Equation (10.34) can be written (to first order in η) as
l2
df η
mη̈ = f (a) + η + · · · + 3
(1 − 3 + · · ·),
dr a ma a
9If
you have not already done so, you should immediately memorize the follow-
ing extremely important series expansions:
Taylor’s Series:
1 2 d2 F 1 3 d3 F
dF
F (a + δx) = F (a) + δx + δx + δx + ··· .
dx
a 2! dx2
a 3! dx3
a
Binomial Expansion:
n(n − 1) 2 n(n − 1)(n − 2) 3
(1 + x)n = 1 + nx + x + x + ··· .
2! 3!
358 10. CENTRAL FORCE MOTION: THE KEPLER PROBLEM
or
l2 η l2
. df
mη̈ = f (a) + η + − 3 .
dr a ma3 a ma3
Now recall that Equation (10.32) states that f (a) = −l2 /ma3 , so the
first and third terms on the right hand side cancel. This leaves
η l2
. df
mη̈ = η − 3
dr a ma3
a
3l2
df
= η − ,
dr a ma4
or 2
3l 1 df
η̈ + η − = 0. (10.35)
m2 a4 m dr a
This equation has the form
η̈ + Kη = 0.
We will be considering equations of this form in detail later, but for
now it is sufficient to note that the general form of the solution depends
on whether K is positive or negative. For positive K the solution has
the form √ √
η = A sin Kt + B cos Kt, (10.36)
and for negative K the general form of the solution is
√ √
η = Ce+ |K|t + De− |K|t . (10.37)
Thus, if K > 0, the solution is simple harmonic motion (see Section
3.6). On the other hand, if K is negative, the solution is the sum of an
exponential decrease plus an exponential increase. The exponentially
decreasing term quickly dies out and η(t) increases with time exponen-
tially. We can interpret this result physically to mean that for K < 0
the orbit is unstable because the “distance” η from the radius a of a
circular orbit grows without bounds. On the other hand, if K > 0,
then η oscillates back and forth around zero, meaning that the parti-
cle oscillates about r = a in simple harmonic motion. Therefore, the
condition for a stable orbit is that K > 0. That is,
3l2
1 df
− > 0.
m2 a4 m dr r=a
Once again we can use the fact that l /ma3 = −f (a), and rewrite this
2
equation as
3 df
− f (a) − > 0. (10.38)
a dr r=a
10.10. A PERTURBED CIRCULAR ORBIT 359
df
The gravitational force is f = −GM m/r2 and dr = +2GM m/r3 . So,
for gravity, Equation (10.38) becomes
3 GM m 2GM m GM m
− 2
− 3
= + 3 > 0,
a a a a
and consequently, the gravitational force leads to stable orbits.
But what if the force is not inverse square? A more general central
force might have the form
c
f =− .
rn
Here n is an integer that can be greater or less than zero. (For the
gravitational force, n = 2; notice that I explicitly incorporated the
negative sign on the right hand side because only attractive force laws
lead to orbital motion.)
If f = − rcn , then
df cn
= n+1 .
dr a a
or
n < 3.
The inverse square force law (n = 2) leads to orbits that are stable
under small perturbations. If the gravitational force had the form
F ∝ 1/r4 , a planetary system would be impossible because any tiny
perturbation would cause the planets to spiral away from their orbits.
Notice that in determining the stability of circular orbits, we also
solved the problem of determining the frequency of small oscillations
in a perturbed orbit because Equation (10.35), the equation of motion
for η, has the form
..
η + Kη = 0,
360 10. CENTRAL FORCE MOTION: THE KEPLER PROBLEM
spheroid and bulges slightly at the equator. This bulge exerts a force
on an orbiting artificial satellite, and gives rise to a radial oscillation of
frequency ωr that is not equal to θ̇. The satellite moves in an elliptical
orbit, but since the radial motion is not exactly synchronized with
the angular motion, the ellipse slowly precesses. In other words, the
major axis of the ellipse slowly rotates at a rate which is the difference
between the two oscillations. The angular frequency of the precession
of the ellipse, ωp is given by
ωp = ωr − θ̇
where ωr is the frequency of the radial oscillations.
10.11. Resonances
An orbit is usually thought of as the path of a particle that retraces
its motion over and over again, as in the case of a single planet around
a star. If the particle always passes through the same points, the orbit
is said to be closed. In a simple closed orbit, the radial period and the
362 10. CENTRAL FORCE MOTION: THE KEPLER PROBLEM
angular period are equal. That is, the planet goes from one turning
point to the other and back again in the same time as the angular
displacement goes from 0 to 2π. If the radial period is slightly greater
than the angular period, the planet will reach perihelion a short time
after its angular position has advanced 2π. The orbit is not closed
and the perihelion precesses. The orbit can actually be closed even
if the two periods are not equal as long as they are in the ratio of
integers, as in the middle picture of Figure 10.14. When two periods
or (equivalently) two frequencies are in the ratio of small integers, they
are said to be commensurable or in resonance. If the ratio of the radial
frequency
√ to the angular frequency is an irrational number such as π
or 2, the orbit is not closed; it will never repeat itself. Figure 10.14
illustrates two commensurable orbits and one incommensurable orbit.
Resonances are very important in celestial mechanics. One often
hears of a system being “locked” into a particular resonance. For ex-
ample, the rotation of the Moon is locked into a 1:1 resonance with
its orbital motion. Therefore, the same side of the Moon always faces
Earth. Similarly, the orbit of a 24 hour satellite is in a 1:1 resonance
with the rotation of the Earth. An interesting case is the 2:5 resonance
between the periods of Jupiter and Saturn. In old fashioned terminol-
ogy used in celestial mechanics, this resonance is called “The Great
Inequality.”
10.12. Summary
A central force is directed towards or away from the origin and has a
magnitude that depends only on the distance to the origin, F = f (r)r̂.
An example is the gravitational force acting on a planet as it orbits the
Sun.
Kepler studied the motion of the planets and determined that this
motion obeys three relations which we call Kepler’s Laws.
A particle moving in a central force field has constant angular mo-
mentum. The equations of motion for a particle of mass m in the field
of a body of mass M are
l2 GM
r̈ − 2 3
= − 2
mr r
d dl
(mr2 θ̇) = = 0.
dt dt
The effective potential energy for a particle in a gravitational field
is
l2 GM m
Vef f = 2
− .
2mr r
10.13. PROBLEMS 363
10.13. Problems
Problem 10.1. (Bohr model of the atom.) Consider an electron in
a circular orbit around a proton. Assume the angular momentum can
only take on values equal to nh where n is an integer (n = 1, 2, 3, ...)
and h is a constant. Determine the possible values of the radius for the
orbit and the possible values for the total energy. Make up a table of
the energy for the first four energy states (that is, for the electron in
the four smallest orbits).
Problem 10.2. Integrate Equation (10.15) directly and show that
it leads to the equation of an ellipse. Note that α = −1, β is positive
and γ is negative.
364 10. CENTRAL FORCE MOTION: THE KEPLER PROBLEM
rest.) (d) Express the radial equation (for the magnitude of r) in terms
of the angular momentum and the reduced mass. (e) By comparing the
radial equation with Equation (10.8), obtain an expression for Kepler’s
third law for this situation.
Problem 10.8. Show that for a planet in an elliptical orbit about
the Sun, the radial velocity at any time is given by
2 2 GM
r ṙ = (a[1 + e] − r) (r − a[1 − e]) .
a
where M is the mass of the Sun. (Hint: Use Equations (10.27) and
(10.25).)
Problem 10.9. A particle of mass m is acted upon by an attractive
central force given by K/r4 . The particle ispplaced a distance a from
the force center and given an initial velocity 2K/3ma3 at right angles
to the radius vector. (a) Show that the particle spirals into the force
center by deriving the equation for the orbit. (b) Determine the time
for the particle to collide with the force center. (Hint: For this force,
3
p energy is V = −K/3r .) Answers: (a) r = (a/2)(1+cos θ),
the potential
(b) 3π/8 3ma5 /2K.
Problem 10.10. Consider a central force given by F (r) = −K/r3
with K > 0. Plot the effective potential and discuss possible types of
motion.
Problem 10.11. A particle is subjected to a central force
K K0
F (r) = − 2 + 3 .
r r
Assume K > 0 and consider both signs for K 0 . (a) Draw the effective
potential and discuss possible types of motion. (b) Solve the orbital
equation and show that the bounded orbits have the form
a(1 − e2 )
r=
1 + e cos αθ
2 0
as long as l > −mK .
Problem 10.12. The first artificial satellite of the Earth was the
Russian Sputnik I. Its perigee was 227 km. above the Earth’s surface.
At this point its speed was 28,710 km/hr. Determine its period of
revolution. What is the maximum distance from the satellite to the
surface of the Earth? Given: the Earth has a radius of 6.37X103 km.
Problem 10.13. A certain satellite has a perigee of 360 km and an
apogee of 2549 km above the Earth’s surface. Find its distance above
366 10. CENTRAL FORCE MOTION: THE KEPLER PROBLEM
the surface when it is 90◦ from perigee, as measured from the center of
Earth.
Problem 10.14. For the two body problem there is a conserved
quantity called the Laplace vector. (It is also called the Runge-Lenz
vector.) The Laplace vector is a vector pointing towards periapsis. Its
magnitude is proportional to the eccentricity. It can be expressed as:
A = p × l−GM m2 r̂
where M is the mass of the primary, p is the linear momentum and l
is the angular momentum.
a) Show that A lies in the orbit plane.
b) Show that A is a constant of the motion.
c) Show that the magnitude of A is GM m2 e. (Hint: Evaluate r · A.)
Problem 10.15. Imagine a circular orbit passing through the ori-
gin. Then r = A cos θ. Show that a central force that gives rise to such
an orbit has a 1/r5 dependence.
Problem 10.16. A particle moves under the action of a central
force in a spiral described by r = Aebθ . Show that the force acting on
it is inversely proportional to the cube of r.
Problem 10.17. A satellite is in a circular orbit, a distance r from
the center of Earth. It has a known velocity
√ v. Show that this satellite
has an escape velocity (from r to ∞) of 2v.
Problem 10.18. Two stars of equal mass orbit around their com-
mon center of mass with period τ. Show that if they were suddenly
stopped, dead still, √
and allowed to fall toward each other, they would
collide in a time τ / 32.
Problem 10.19. Consider a binary star system. (a) Show that
Kepler’s third law can be written as
a3 = τ 2 (M1 + M2 )
where a is in AU (astronomical units) τ is in Earth years and M1 + M2
is the sum of the masses of the stars in solar masses. (Note: All the
equations we developed in this chapter assumed M >> m. Hint: Red-
erive Equation (10.1) and show that if the two masses are comparable,
that GM should be replaced by G(M1 + M2 ) everywhere.) (b) Star
A and star B are members of a binary star system. An astronomer
determines that the period of this system is 32 years and that the stars
are separated by 16 AU. Furthermore, star A is found to be 12 AU
from the center of mass of the system. Determine the masses of these
10.13. PROBLEMS 367
where M is the mass of the Sun, and r0 is the average radius of the
orbit.
Problem 10.25. Consider a particle that is acted upon by the
central force:
A B
F = 2 + 4,
r r
where A and B are constants. The orbit is a circle of radius a. Deter-
mine the condition for the orbit to be stable. (Answer: a2 A > B.)
Problem 10.26. A particle is moving in a circular path under the
action of an attractive central force given by
1 −r/a
F = e .
r2
Show that if the radius of the circle is greater than a the motion is
stable, and if the radius of the circle is less than a, the motion is
unstable.
Problem 10.27. An ideal massless string passes through a small
hole in a perfectly smooth table. Two equal masses are attached to the
ends of the string so that one mass is on the table a distance a from
the hole, and the other mass is hanging freely. See the sketch (Figure
√
10.15). The mass on the table is set in motion with a velocity ga
perpendicular to the direction of the string. Show that the mass on
the table moves in a circular path. Now the hanging mass is perturbed
slightly so that it
poscillates up and down. Show that the period of this
oscillation is 2π 2a/3g.
where
F1 = f (x, t),
1 1
F2 = f (x + τ F1 , t + τ ),
2 2
1 1
F3 = f (x + τ F2 , t + τ ),
2 2
1
F4 = f (x + τ F3 , t + τ ).
2
Computational Project 10.1. A comet is jostled loose from the
Oort Cloud and heads toward the Sun. Suppose that when it passes
the orbit of Pluto it has velocity components vx = −0.01, vy = 0.05
AU/year. Plot the trajectory of the comet using three algorithms:
Euler-Cromer, Second Order Runge Kutta and Fourth Order Runge
Kutta. Show that the angular momentum is constant. (The only bodies
involved are the comet and the Sun. Hint: Use GM = 4π 2 .)
at a constant angular speed. However, you can get the position of the
planet at equal intervals of time by determining the eccentric anom-
aly E which is related to the average angular velocity n by Kepler’s
equation10
E − e sin E = nt.
This is a transcendental equation so you will have to solve for E nu-
merically. Write a program that obtains E iteratively, using the fact
that e is a small quantity. That is, write Kepler’s equation as
E = nt + e sin E.
Since e is small, the first approximation to E is E = nt. Use this for E
on the right hand side and get the second approximation. Repeat over
and over. Thus:
E0 = nt
E1 = nt + e sin E0
E2 = nt + e sin E1
etc.
To solve for the average potential energy, let the planet orbit the Sun
10 times, determine the time average potential energy and compare it
with −GM m/a.
x
k
m
l
θ
toward the equilibrium point. This form for the equation of motion is
characteristic of oscillatory systems. Whenever you see it, you can be
sure that you are faced with a case of simple harmonic motion. You can
easily show by substitution that possible solutions to the two simple
harmonic motion equations above are:
r !
k
x(t) = A cos t+β , (11.4)
m
and
r
g
θ(t) = A cos t+β . (11.5)
l
These solutions describe
p a sinusoidal oscillation of amplitude Apand
angular frequency k/m for the spring and angular frequency g/l
for the pendulum. The quantity β is related to the initial displacement.
Consider now the somewhat more complicated case of an oscillating
system subjected to a retarding force such as air resistance or some sort
of friction. For the sake of a specific example, let the system be a mass
on a spring acted upon by a resistive force that is proportional to the
velocity. Then the total force acting on the mass is not just −kx, but
rather,
F = −kx − bẋ,
where b is a constant of proportionality. The minus sign guarantees
that the retarding force opposes the motion. The −bẋ term is called a
damping force. The equation of motion will now be
mẍ + bẋ + kx = 0. (11.6)
This is the equation of motion for a damped harmonic oscillator (as-
suming the damping force is proportional to the first power of the
velocity).
376 11. HARMONIC MOTION
mean that the dependent variable (x) and its derivatives appear only
to the first power. However, the independent variable (t) can be raised
dx 2
to any power. Thus dx dx
2 2
dt
= 3t is linear, but dt
= 3x and dt
= 3t
are not linear.1
11.2.1. Homogeneous Linear Differential Equations. A ho-
mogeneous differential equation of the form given above has a solution
x = ept .
The technique for obtaining the general solution is to plug x = ept into
the equation and carry out the obvious mathematical operations. This
yields an equation for p in terms of the constants. I will illustrate this
procedure in a moment. But first I want to mention that a first-order
differential equation yields a single value for p, a second-order differen-
tial equation yields two values for p and so on. Thus, depending on the
order of the differential equation, you get several different solutions.
Which one of these various solutions should you use? The answer is:
all of them!
Rule 1: If x1 (t) and x2 (t) are both solutions of a linear
homogeneous differential equation then x1 (t) + x2 (t) is also a
solution.
For simple harmonic motion we have a second order differential
equation so we will obtain two values for p. Let us call these two values
p1 and p2 . The two solutions are:
x1 (t) = ep1 t and x2 (t) = ep2 t .
By Rule 1, the sum of these two solutions is also a solution. That is,
x(t) = ep1 t + ep2 t
is a solution.
We can generalize even further using another simple fact.
Rule 2: If x(t) is a solution of a linear homogeneous differ-
ential equation, then Cx(t), where C is an arbitrary constant,
is also a solution.
Consequently,
x = C1 ep1 t + C2 ep2 t
is a solution to a second-order differential equation. In fact, it is the
most general form for the solution to a homogeneous linear second-order
1The “degree” of a differential equation is the power of its highest order deriv-
ative. Thus a linear differential equation is of the first degree. The “order” of a
differential equation is the order of the highest derivative appearing in the equation.
378 11. HARMONIC MOTION
You might think that since x is a real quantity, you should discard the
last term because it is imaginary; however, this would be unjustified
because C1 and C2 are not yet determined and they could very well be
complex numbers.
You now know basically everything you need for solving simple
harmonic motion problems, except for one small detail: If the auxiliary
equation for a second order differential equation has two equal roots,
then p1 = p2 = p and the general solution is
x = C1 ept + C2 tept . (11.9)
This is easily verified by substitution. (Note the factor of t in the
second term.)
Worked Example 11.1. Solve
d2 x
− 4x = 0,
dt2
given the initial conditions x|t=0 = 0, dx = 3.
dt t=0
Solution: Note that this is not the SHM equation because
the coefficient of x is negative. To solve, substitute ept into the
equation and obtain
p2 − 4 = 0.
Therefore, p = ±2 and the solution is
x = C1 e2t + C2 e−2t .
The initial condition x(t = 0) = 0 leads immediately to C2 = −C1
so
x = C1 (e2t − e−2t ) = 2C1 sinh(2t).
The initial condition ẋ = 3 leads to
ẋ = 2C1 (2) cosh(2t) = 4C1 cosh(0) = 3.
∴ C1 = 3/4.
and
3
x= sinh(2t).
2
x(t = 0) = x0 = C1 + C2 ,
and
ẋ(t = 0) = 0 = iωC1 − iωC2 .
The second equation requires that C2 = C1 . Then the first equation
implies that C1 = C2 = 12 x0 . Therefore, the solution is
1 1
x = x0 (e+iωt + e−iωt ) = x0 (2 cos ωt) = x0 cos ωt.
2 2
The last equation tells us that the mass oscillates back and forth be-
tween the values +x0 and −x0 . The quantity x0 is called the amplitude.
The motion takes place with a frequency
r
ω 1 k
f= = .
2π 2π m
11.2. SOLVING THE DIFFERENTIAL EQUATION (OPTIONAL) 381
x
x0
P/2
t
P
-x0
P=2π
Figure 11.3. A plot of position as a function of time
for a mass undergoing simple harmonic motion.
energy is a maximum but the kinetic energy is zero. Since the total
energy is constant,2
1 2 1 2
mv = kx .
2 0 2 0
You can use this equation to relate x0 and v0 , thus:
r
m v0
x0 = ± v0 = ± .
k ω
The result (11.10) can be written in the more familiar form
x = x0 sin ωt.
The two examples you have just considered had two very simple
sets of initial conditions. In both cases C1 and C2 were replaced by
a single constant. In general, however, the solution will contain two
different constants. The general solution can also be expressed in the
form
x = x0 sin(ωt + β).
Worked Example 11.2. A mass of 0.25 kg is attached to a
spring of force constant 1.0 N/m. The mass is displaced 0.15 m
from its equilibrium point and released with zero initial velocity.
Evaluate the total energy of the oscillator (kinetic plus potential).
What is the maximum velocity? What is the period of the oscilla-
tion?
Solution: p p
ω = k/m = 1/.25 = 2,
2Note that this result is in agreement with results from the virial theorem
discussed in Worked Example 10.2. In the present problem, n = 1 so
< T >= n+1 2 < V > yields < T >=< V >
11.3. THE DAMPED HARMONIC OSCILLATOR 383
f.s = -kx
k x
m
.
ff = -bx
Let us call the two solutions p1 and p2 . Since ep1 t and ep2 t are both
solutions to the differential equation, the general solution is the sum of
these;
x = C1 ep1 t + C2 ep2 t , (11.12)
where C1 and C2 are constants to be determined from the initial condi-
tions.
If p1 and p2 are complex, having an imaginary component such as
iω, then the solution contains terms of the form
eiωt = cos ωt + i sin ωt.
From your knowledge of the behavior of sines and cosines, you know
that such a solution represents oscillatory motion. On the other hand,
if p1 and p2 are real, then the solution is made up of terms having the
form
e−at and e+at ,
where a is a real quantity. Such a solution will either represent an
exponential decrease or an exponential increase (depending on whether
the sign is negative or positive). From physical considerations we can
rule out exponentially increasing solutions for the problem of a mass
on a spring.
From the equation for p (Equation 11.11) you note that p is complex
if and only if the quantity under the square root is negative, that is, iff
2
b k
< .
2m m
In this case the system oscillates. On the other hand, if
2
b k
≥ ,
2m m
then p is real and the motion dies out exponentially.
The relation of k/m to (b/2m)2 characterizes the three types of
motion for the damped oscillator. These are called “underdamped,”
“overdamped,” and “critically damped.” They are described in the fol-
lowing table:
11.3. THE DAMPED HARMONIC OSCILLATOR 385
An underdamped oscillator has γ 2 < ω02 so the last term is the square
root of a negative number, and p is complex. In that case it is conve-
nient to write p in the form
p = −γ ± iω1
p
where ω1 = + ω02 − γ 2 . Note that ω12 = ω02 − γ 2 is positive, so ω1 is a
real number.
Inserting the two values of p into the solution (11.12) gives
x(t) = C1 e(−γ+iω1 )t + C2 e(−γ−iω1 )t
= e−γt (C1 eiω1 t + C2 e−iω1 t ).
C1 and C2 are complex constants related to the initial conditions. These
constants can be written in a variety of different ways. You obtain a
386 11. HARMONIC MOTION
-A
Figure 11.5. The motion of an underdamped oscillator
consists of rapid oscillations [described by cos(ω1 t + θ)]
modulated by an exponentially decreasing amplitude
(described by Ae−γt ).
x
C2e- 2t
x(t)
t
C1e- 1t
seat a shove each time it comes back to you at the top of its swing.
The periodicity of your shoves is the same as the natural period of
the swing. Similarly, the pendulum in a “grandfather” clock receives
a small push on each swing. An old-fashioned pocket watch has an
oscillating spring wheel that receives small periodic impulses from a
spring and ratchet system.
The oscillator will, in general, be damped, so the equation of motion
for such a system is
where F (t) is the driving force. The purpose of the driving force is
to overcome the damping. If F (t) is too small, the oscillations will
gradually die out. On the other hand, if F (t) is too large, the amplitude
may grow unacceptably large. (If you push a swing too hard it will “go
over the top” leading to unpleasant consequences.) As you shall see
shortly, the amplitude of the oscillations also depends on the frequency
of the driving force. This is related to the phenomenon of resonance.
The equation of motion of the driven harmonic oscillator (Equation
11.17) is an inhomogeneous second order linear differential equation. I
will now make a short aside to describe how to solve such an equation.
an Dn x + an−1 Dn−1 x + · · · + a1 Dx + ao x = 0.
The forcing term, F (t), is usually periodic because that will keep the
oscillator in motion, overcoming the effect of damping. Another com-
mon situation is an oscillator that has been subjected to an impulsive
force. In general, the forcing term can be any function of time. It
could even be a constant. In fact, for the sake of making the analysis
simple, I will begin with a constant force. For example, a mass hanging
from a spring in an external gravitational field (see Figure 11.7) has
the equation of motion
where the damping could be due to air resistance, for example. The
mass is only allowed to oscillate vertically; it does not simultaneously
swing like a pendulum.
mẍ + kx = F0 sin ωd t,
or
F0
ẍ + ω02 x = sin ωd t. (11.19)
m
11.4. THE FORCED HARMONIC OSCILLATOR 393
3See, for example, Mary L. Boas, Mathematical Methods in the Physical Sci-
ences, 3rd ed. John Wiley & Sons, Inc., Hoboken, NJ, 2006. pp 417 ff.
394 11. HARMONIC MOTION
and hence,
u = C1 e2t + 4t + 1.
But since u = (D + 2)x we have
dx
+ 2x = C1 e2t + 4t + 1.
dt
Now the integrating factor is e2t so we multiply through by this
factor to get
d
(xe2t ) = C1 e4t + 4te2t + e2t ,
dt
and consequently,
1 1
x(t) = C2 e−2t + C1 e2t + 2t − .
4 2
Inserting initial conditions leads to C1 = 4 and C2 = −1/2, and
finally
1 1
x(t) = e2t − e−2t + 2t − .
2 2
ξp (t) = Ceiωd t .
(I wrote ξp rather than xp because this is not yet the particular solu-
tion.) To obtain C, plug ξp back into the differential equation, thus:
2
d F0 iωd t
2
2
+ ω0 Ceiωd t = e .
dt m
Hence,
F0
−ωd2 C + ω02 C = ,
m
and
F0 /m
C= .
ω02 − ωd2
Finally, the particular solution is
F0 /m iωd t F0 /m
xp (t) = Im (ξp (t)) = Im 2 2
e = 2 sin ωd t,
ω0 − ωd ω0 − ωd2
and the general solution of Equation (11.19) is
F0 /m
x(t) = A cos(ω0 t + θ0 ) + sin ωd t.
ω02 − ωd2
The problem is now solved, except for determining θ0 from the
initial conditions. However, before leaving this problem, I want you
to notice that there is something quite interesting about the particular
solution,
F0 /m
xp = sin ωd t.
ω02 − ωd2
11.4. THE FORCED HARMONIC OSCILLATOR 397
You have already seen that the general solution of the homogeneous
equation for the underdamped harmonic oscillator is given in Equation
(11.15) as x = Ae−γt cos(ω1 t + θ). The general solution of the forced,
underdamped harmonic oscillator is, then,
2 2
F0 /m [(ω0 − ωd ) − iωd b/m] iωd t
x(t) = Ae−γt cos(ω1 t + θ) + Im h i e .
2
(ω02 − ωd2 ) + (ωd b/m)2
Eventually, the first term will die out. (This is, of course, also true
for overdamped and critically damped oscillators.) Dropping the first
term, you are left with
2 2
F0 /m [(ω0 − ωd ) − iωd b/m]
x(t) = Im h i (cos ωd t + i sin ωd t) .
2 2 2 2
(ω0 − ωd ) + (ωd b/m)
where
ω02 − ωd2
−1
φ = tan − .
ωd b/m
It is interesting to note that the amplitude of x(t) is not maximized
at ωd = ω0 but rather at a nearby frequency called the resonant fre-
quency which you can determine by evaluating the derivative of the
11.4. THE FORCED HARMONIC OSCILLATOR 399
R
C L
ω
ω
k3 k2
k1
m1 m2
X1 X2
mẍ1 + (k + k3 )x1 − k3 x2 = 0,
mẍ2 + (k + k3 )x2 − k3 x1 = 0.
Denote k + k3 by k 0 . Dividing by m yields:
ẍ1 + (k 0 /m)x1 − (k3 /m)x2 = 0,
ẍ2 + (k 0 /m)x2 − (k3 /m)x1 = 0.
Now replace k 0 /m by ω02 . The reason for this notation is that ω0 is the
frequency of oscillation of either one of the masses if the other mass is
held at rest. Furthermore, let us denote k3 /m by ∆ω 2 , for reasons that
will soon become clear.
Using this new notation the equations become
ẍ1 + ω02 x1 − ∆ω 2 x2 = 0, (11.30)
ẍ2 + ω02 x2 − ∆ω 2 x1 = 0.
This system of coupled linear differential equations can be solved by
assuming the solutions
x1 = Aeiωt ,
x2 = Beiωt .
We assume oscillatory solutions (that is why the exponent is imagi-
nary) with both of the variables x1 and x2 having the same angular
frequency ω. (Except for a slight change in the notation, this is the
same technique previously used in assuming solutions of the form ept .)
Plugging the expressions for x1 and x2 into the coupled equations of
motion yields
A(iω)2 eiωt + ω02 Aeiωt − ∆ω 2 Beiωt = 0,
B(iω)2 eiωt + ω02 Beiωt − ∆ω 2 Aeiωt = 0,
or
(−ω 2 + ω02 )A − ∆ω 2 B = 0, (11.31)
−∆ω 2 A + (−ω 2 − ω02 )B = 0.
These two equations are a system of homogeneous coupled linear alge-
braic equations in the unknowns A and B. The technique for solving
systems of coupled algebraic equations is to apply Cramer’s rule. I be-
lieve you are familiar with this rule, but I will review it briefly anyway.
404 11. HARMONIC MOTION
ω 2 = +ω02 ∓ ∆ω 2 .
That is,
1
ω = ±(ω02 ∓ ∆ω 2 ) 2 .
11.5. COUPLED OSCILLATORS 405
This yields four possible values for ω. Two of them can be written
1
ω1 = ±(ω02 − ∆ω 2 ) 2 ,
and the other two are
1
ω2 = ±(ω02 + ∆ω 2 ) 2 . (11.32)
The general solutions for x1 and x2 are the sum of all possible solutions:
x1 = A1 eiω1 t + A−1 e−iω1 t + A2 eiω2 t + A−2 e−iω2 t , (11.33)
x2 = B1 eiω1 t + B−1 e−iω1 t + B2 eiω2 t + B−2 e−iω2 t .
These two expressions involve eight constants. But solving two second
order differential equations should only lead to four constants. There-
fore, not all of the eight constants are independent. We now determine
the four independent constant amplitudes.
Note that each term in our expression for x1 satisfies the differential
equation. Thus, for example, a possible solution involves oscillations
of frequency ω1 . Then,
x1 = A1 eiω1 t
x2 = B1 eiω1 t
where ω12 = ω02 −∆ω 2 . Plugging into the equations of motion (Equation
11.30) yields
(ω02 − ω12 )A1 − ∆ω 2 B1 = 0
and
−∆ω 2 A1 + (ω02 − ω12 )B1 = 0
Either one of these yields the same result, namely,
−ω02 − (ω02 − ∆ω 2 ) A1 − ∆ω 2 B1 = 0,
or
+∆ω 2 A1 − ∆ω 2 B1 = 0.
Therefore,
B1 = A1 .
Similary by considering the solution in −ω1 we obtain B−1 = A−1 , and
±ω2 lead to B2 = −A2 and B−2 = −A−2 .
So our equations reduce to
x1 = A1 eiω1 t + A−1 e−iω1 t + A2 eiω2 t + A−2 e−iω2 t ,
x2 = A1 eiω1 t + A−1 e−iω1 t − A2 eiω2 t − A−2 e−iω2 t .
These solutions are not in a convenient format because x1 and x2 must
be real, but e±iω1 t and e±iω2 t are complex, so the coefficients Ai must
406 11. HARMONIC MOTION
x1 (t = 0) = b,
x2 (t = 0) = b.
But in general,
x1 (t) = C1 cos(ω1 t + θ1 ) + C2 cos(ω2 t + θ2 ),
x2 (t) = C1 cos(ω1 t + θ1 ) − C2 cos(ω2 t + θ2 ).
so,
b = C1 cos θ1 + C2 cos θ2 , (11.37)
b = C1 cos θ1 − C2 cos θ2 .
The velocities of the blocks are:
408 11. HARMONIC MOTION
C1 = b,
C2 = 0.
Therefore the motion is described by Equations (11.36) as desired. Fur-
thermore, C1 = b and θ1 = 0, so
x1 (t) = b cos ω1 t, (11.38)
x2 (t) = b cos ω1 t,
p
where ω1 = ω02 − ∆ω 2 . This is the low frequency normal mode.
Similarly, a different normal mode can be generated by either pulling
the two masses apart or pushing them together by equal amounts, and
releasing them from rest. Assuming the initial displacements are equal
to b0 (but in opposite directions), we find C1 = 0 and C2 = b0 .
The motion is then described by
x1 = b0 cos ω2 t, (11.39)
x2 = −b0 cos ω2 t,
where 1
ω2 = (ω02 + ∆ω 2 ) 2 .
This means the system oscillates at a higher frequency than the “nat-
ural” frequency ω0 .
Note that the general solutions (Equations 11.34 and 11.35) are
simply linear combinations of the normal mode solutions, Equations
(11.38) and (11.39). As mentioned before, the motion of any system
of coupled oscillators is always a linear combination of normal modes.
The problem gets more complicated when the number of oscillators
11.6. SUMMARY 409
11.6. Summary
This chapter was an in-depth study of harmonic motion using a
mass connected to a spring as an example. There were several digres-
sions to discuss the mathematical techniques involved in solving the
equations of motion. These techniques are used frequently in many
areas of physics, so you should make a special effort to understand and
remember them.
The analysis started with an undamped simple harmonic oscillator
with equation of motion
mẍ + kx = 0.
Next we generalized to a damped harmonic oscillator in which there
was a frictional force acting to slow the system down, with equation of
motion
mẍ + bẋ + kx = 0.
There were three cases: the underdamped oscillator with (b/2m)2 <
k/m, the critically damped oscillator with (b/2m)2 = k/m, and the
overdamped oscillator with (b/2m)2 > k/m. The underdamped oscil-
lator is the only one exhibiting truly oscillatory motion with the mass
moving back and forth in a regular, repetitive way. Note, however, that
in all three cases the amplitude of the motion decreases as time goes
on. Mathematically this is described by the amplitude being given by
Ae−γt so that the motion of an underdamped oscillator is
x(t) = Ae−γt cos(ω1 t + θ),
the motion of a critically damped oscillator is
x(t) = (C1 + C2 t) e−γc t ,
and the motion of an overdamped oscillator is:
x(t) = C1 e−γ1 t + C2 e−γ2 t .
Note also that the frequency of the motion is affected; an underdamped
oscillator has a frequency smaller than the frequency of the undamped
oscillator. The overdamped oscillator has a “frequency” greater than
410 11. HARMONIC MOTION
that of the undamped oscillator, but it does not really oscillate, the
motion simply dies out before a single full oscillation can occur.
Next you studied the forced harmonic oscillator. In this case the
motion is represented by the following inhomogeneous differential equa-
tion:
mẍ + bẋ + kx = F (t).
Obtaining the motion required summing the general solution of the ho-
mogeneous differential equation and a particular solution of the inho-
mogeneous equation. You learned a number of “tricks” for determining
the particular solutions of inhomogeneous equations.
Finally, you learned about coupled undamped harmonic oscillators
whose equations of motion are
m1 ẍ1 = −k1 x1 − k3 (x1 − x2 ),
m2 ẍ2 = −k2 x2 − k3 (x2 − x1 ).
This is an important problem in physics because there are many phys-
ical systems that can be represented as coupled harmonic oscillators.
Normal modes were defined and it was noted that the general motion of
coupled oscillators can be expressed as a linear combination of normal
mode solutions.
11.7. Problems
Problem 11.1. Determine the condition such that x = Ctept is a
.. .
solution of mx + bx + kx = 0.
Problem 11.2. Determine the time average of kinetic energy and
potential energy for a simple harmonic oscillator and show that they
are equal.
Problem 11.3. A bathroom scale should not oscillate. Ideally it
would be critically damped. Show that if a scale is critically damped for
a person of weight W it will be overdamped for a person whose weight
is less than W. If it is desired that for critical damping, the platform
deflect 2 cm for a 70 kg person, determine the spring constant k and
the damping constant b.
Problem 11.4. A 0.25 kg mass is attached to a spring of force
constant 0.02 N/m. It is placed in a resistive medium and released
from rest at a distance of 0.1 m from equilibrium. After 5 seconds, it
is observed that the amplitude of the oscillations is 0.05 m. What is
the damping constant b? What is the frequency of the oscillations?
11.7. PROBLEMS 411
x1
t
x2
t
x1 l1 l2 x2
m1 m3 m2
k1 k2
which work is done on the oscillator by the applied force, and eval-
uate the average power delivered by the force per cycle. Answer:
P = 12 F0 ẋmax cos β.
A few Differential Equation Problems
Problem 11.21. Obtain the general solution to (D2 +1)x = 12 cos2 t.
Answer: x = A cos t+B sin t+6−2 cos 2t. (Hint: use the method of un-
determined coefficients. You will have to look it up in your differential
equations text.)
Problem 11.22. Obtain the general solution to (D2 +4)x = 4 sin2 t.
(Hint: use the method of undetermined coefficients.)
COMPUTATIONAL PROJECT
Computational Project 11.1. A physical pendulum is illus-
trated in figure 11.13. Assuming that it is damped and driven by a
force of frequency f, its equation of motion can be written in the (sim-
plified) form
d2 θ dθ
2
+ b + sin θ = T sin(2πf t).
dt dt
It will be easiest to solve this equation numerically if you introduce
three new variables defined as follows:
dθ
x1 = ,
dt
x2 = θ,
x3 = 2πf t.
Then the equations of motion are
dx1
= T sin(x3 ) − sin(x2 ) − bx1 ,
dt
dx2
= x1 ,
dt
dx3
= 2πf.
dt
Write a computer program and plot θ (or x2 ) as a function of time for
t ≤ 250. First assume b = 0 and T = 0 to be sure that your solution
generates the oscillatory motion of a pendulum. Then set b = 0.1 and
obtain the motion of a damped pendulum. Finally, let T = 1, and
b = 0.1 and allow f to range from 0.001 to 0.1. Note that for f = 0.1
the motion is unpredictable. This is an example of chaos in a dynamical
system.
11.7. PROBLEMS 415
Torque
θ M
Mg
1Entire books have been written on pendulums. For example, Baken, G.L.,
and J. A. Blackburn, The Pendulum, A Case Study in Physics,Oxford University
Press, 2005, is a 300 page book on various aspects of pendulums.
417
418 12. THE PENDULUM
l
θ T
mg
Integrating,
Z θ(t)
dθ √ Z t
r = 2ω dt,
θ0 =0 E 0
mgl
+ cos θ
or Z θ
1 dθ
t= √ r , (12.3)
2ω 0 E
mgl
+ cos θ
where the lower limit, θ0 = 0, means the bob was at the bottom of the
swing when the timer was started.
√
r
p 1
cos θ − cos θm = k 2 1 − sin2 2θ .
k2
Therefore, Equation (12.5) is
Z θm
4 dθ
P =√ √ q . (12.6)
2ω 0 1 2 θ
2k 1 − k2 sin 2
Even this is not in the right form, but it is getting close. Define a new
angle φ by the relation
1 θ
sin φ = sin .
k 2
When θ = θm , the value of φ is π/2. Now,
1 θ 1 θ θ 1
cos 2θ dθ,
d(sin φ) = d( sin ) = cos d 2
=
k 2 k 2 2k
422 12. THE PENDULUM
so
2k cos φdφ 2k cos φdφ 2k cos φdφ
dθ = θ
=q =p .
cos 2 1 − sin2 2θ 1 − k 2 sin2 φ
Plugging into Equation (12.6) you obtain
Z φ=π/2
4 2k cos φdφ 1
P = √ p √ p ,
2 ω φ=0 1 − k 2 sin2 φ 2k 1 − sin2 φ
Z π/2
8 cos φdφ 1
= p .
2ω 0 1 − k 2 sin2 φ cos φ
Hence,
4 π/2
Z
dφ
P = p , (12.7)
ω 0 1 − k 2 sin2 φ
which has the desired form.
This somewhat tiresome derivation yielded the expression for the
period of a pendulum of arbitrary amplitude. The period is, however,
expressed in terms of an elliptic integral. Let me remind you that an
elliptic integral is a function like the sine or cosine in the sense that
its value can be looked up in any standard set of math tables. (I do
not know if any handheld calculators have elliptical integrals stored in
them.)
Elliptic integrals come in various different “flavors,” the primary
ones being:
(1) Elliptic Integrals of the First Kind:
Z θ
dφ
F (k, θ) = p . (12.8)
0 1 − k 2 sin2 φ
(2) Elliptic Integrals of the Second Kind:
Z θq
E(k, θ) = 1 − k 2 sin2 φ dφ. (12.9)
0
(3) Elliptic Integrals of the Third Kind:
Z θ
dφ
Π(n, k, θ) = 2
p . (12.10)
0 (1 − n sin φ) 1 − k 2 sin2 φ
(4) Complete Elliptic Integrals: These are the values of F and
E when θ = π/2. Thus, for example, the complete elliptical
integral of the first kind is:
Z π/2
π dφ
F (k, ) = F (k) = p .
2 0 1 − k 2 sin2 φ
12.1. A SIMPLE PENDULUM WITH ARBITRARY AMPLITUDE 423
You can expand the term in brackets using the binomial expansion
because x < 1. Keeping the first few terms,
4 π/2
Z
1 3
P = dφ[1 + x + x2 + · · · ] (12.11)
ω 0 2 8
or
4 π/2
Z
∼ 1 3
P = dφ[1 + k 2 sin2 φ + k 4 sin4 φ]
ω 0 2 8
(Z )
π/2
1 2 π/2 2 3 4 π/2 4
Z Z
4
= dφ + k sin φdφ + k sin φdφ .
ω 0 2 0 8 0
O
Rc
θ
G
Mg
|Rc |=h
where the distance from the axis of rotation to the center of mass is
denoted h.
When considering the motion of rigid bodies it is convenient to
define a quantity called the “radius of gyration,” denoted by k. The
radius of gyration is defined in terms of the moment of inertia. Recall
that the dimensions of the moment of inertia are mass multiplied by
the square of a distance. If I is the moment of inertia of a body relative
to some given axis, and M is the mass of the body, then the radius of
gyration is the distance k such that
I = M k2.
I0 = M k02 .
M k02 θ̈ = −M gh sin θ.
Consequently,
gh
θ̈ = − sin θ.
k02
Recall that the equation of motion of a simple pendulum is
g
θ̈ = − sin θ.
l
12.2. THE PHYSICAL PENDULUM 427
Comparing the two equations you see that the compound pendulum
oscillates with the same period as a simple pendulum of length
k02
l= . (12.12)
h
That is, a simple pendulum of length k02 /h will oscillate “in time” with
the compound pendulum. Figure 12.3 illustrates such a pendulum.
The point of suspension is O and the bob is at O0 (a distance l =
ko2 /h from O). The point O0 is called the “center of oscillation” of the
compound pendulum. (If all the mass of the compound pendulum
were somehow concentrated at O0 , it would be a simple pendulum that
oscillates with the same period as the real compound pendulum.) p For
small oscillations, the period of a simple pendulum is P = 2π l/g, so
the period of a compound pendulum is
p
P = 2π k 2 /hg.
O
l=k02/h
h l=k02/h
θ
G
h'
O'
Now according to the parallel axis theorem, the relation between the
moment of inertia about O (I0 ) and the moment of inertia about the
center of mass (IG ) is
I0 = IG + M h2 ,
so
M k02 = M kG
2
+ M h2 ,
where kG is the radius of gyration of a pendulum suspended at its
center of mass. Dividing by M,
k02 = kG
2
+ h2 . (12.14)
Comparing Equations (12.14) and (12.13) you can see that
h0 h = kG
2
. (12.15)
0 0
The “new” center of oscillation will be a distance l from O , where
l0 = k002 /h0 . Setting
k002 = h02 + hh0
leads to
1
l0 = 0 h02 + hh0 = h0 + h = l.
h
Therefore, a pendulum suspended at point O0 has its center of oscil-
lation at O. Note the symmetry between the location of the axis of
rotation and the location of the center of oscillation.
P
s
G
a
Force A
But F = dp
dt
so Z Z
dp
J= dt= dp = pf −pi .
dt
Assuming the initial momentum is zero and the final momentum is
M v, the velocity of the rod immediately after the blow is
J
v= .
M
To be a bit more explicit, this is the velocity of the center of mass of
the rod immediately after the blow.
However, it is clear from the physical situation that the rod will
also begin to rotate. Its angular momentum (taken with respect to the
center of mass) will change according to
dL
= N.
dt
2
The moment of inertia about the center of mass is I = M kG . The
magnitude of the torque on the rod (taken about the center of mass)
due to the impulsive force F is
N = |r × F| =aF,
where a is the distance from the center of mass to the line of action of
the force. Taking the derivative of L = Iω,
d
aF = (Iω).
dt
Therefore,
Z Z
aF dt = d(Iω),
aJ = Iω,
12.3. THE CENTER OF PERCUSSION 431
and the angular velocity about the center of mass, immediately after
the impulse, will be
aJ
ω= .
I
The motion of the rod is rather complicated. The velocity of a point
on the rod is the linear velocity of the center of mass, plus the rotation
about the center of mass (given by ω ×r where r is the vector from
the center of mass). These two velocities must be added vectorially,
of course. For example, consider the motion of point P, a distance s
from the center of mass, as shown in Figure 12.4. Immediately after
the impulse, the velocity of this point is
J
vP = + ω × s. (12.16)
M
But the right hand rule and Figure 12.4 tell us that J is directed toward
the right and ω × s is directed toward the left. The magnitude of the
velocity of point P is, consequently,
J aJ
vP = − 2
s
M M kG
J as
= 1− 2 .
M kG
This expression leads to the interesting conclusion that immediately
after the impulse there is a point on the rod having zero velocity. That
is, vP will be zero if
as
2
= 1.
kG
Consequently, immediately after the impulse, a point located at
2
kG
s= (12.17)
a
will have zero velocity. This point is called the center of percussion.
In the notation of the previous section, the distances a and s are
0
h and h, respectively. Inserting these quantities into Equation (12.17)
gives
hh0 = kG 2
,
which is just Equation (12.15). Thus the center of percussion is located
at the same point as the center of oscillation (assuming the axis is at
A, the point where the impulse is applied).
As time goes on, the impulsive force is no longer acting and the
linear and angular velocities will be constant. It is instructive to plot
the positions of a few points in the rod as a function of time to convince
yourself that the rod will rotate about its center of mass at a constant
432 12. THE PENDULUM
P s
F1
G
a
Force A
A
a
r d
question is: What is the correct height (d) for the “bumper” on
the edge of the table in Figure 12.6?
Solution: Since the ball does not slide, point C should be the
center of percussion for a blow at A. Therefore,
2
ar = kG .
For a sphere, I = (2/5)M r2 , so kG
2
= (2/5)r2 . Consequently,
ar = (2/5)r2 ,
and
d = a + r = (7/5)r.
θ
y
φ l
x
m
x = l sin θ cos φ,
y = l sin θ sin φ,
z = l cos θ.
The kinetic energy of the bob is
1
T = m(ẋ2 + ẏ 2 + ż 2 ),
2
1 2 2
= ml (θ̇ + sin2 θφ̇2 ), (12.18)
2
and the potential energy is
V = mgl cos θ.
Consequently, the Lagrangian is
1
L = T − V = ml2 (θ̇2 + sin2 θφ̇2 ) − mgl cos θ. (12.19)
2
The Lagrange equations of motion are
d 2
ml θ̇ − ml2 φ̇2 sin θ cos θ − mgl sin θ = 0, (12.20)
dt
d 2 2
ml sin θφ̇ = 0. (12.21)
dt
The second of these equations can be integrated immediately to give
ml2 sin2 θφ̇ = constant = pφ . (12.22)
The constant is named pφ because it is the generalized (angular) mo-
mentum associated with the (ignorable) coordinate φ.
It is convenient to introduce the energy of the pendulum. Adding
the kinetic and potential energies we obtain
1
E = ml2 (θ̇2 + sin2 θφ̇2 ) + mgl cos θ.
2
Using the expression for pφ to eliminate φ̇ leads to
1 2 2 p2φ
E = ml θ̇ + + mgl cos θ. (12.23)
2 2ml2 sin2 θ
The form of this expression suggests introducing an “effective poten-
tial” Vef f . (Recall the use of the effective potential for the central force
problem - see Section 10.5.) For the spherical pendulum, the term
(1/2)ml2 θ̇2 is a kinetic energy term depending on the velocity, but the
following two terms depend only on coordinates and not velocities so
436 12. THE PENDULUM
Veff = V
E3
mgl
E2
θ
π/2 π 3 /2 2
E1
-mgl E0
swing. Finally, for energies such as E3 > +mgl, the pendulum actually
“goes over the top” and the angle increases (or decreases) continuously.
A more interesting situation arises when pφ 6= 0. The pendulum now
moves azimuthally. It may or may not have motion in the θ direction.
If θ is constant, the pendulum bob traces out a horizontal circle of
constant radius. This special case is called the “conical pendulum.”
If pφ 6= 0, the effective potential is given by Equation (12.24). Vef f
is plotted in Figure 12.9 for various constant but nonzero values of
pφ . The effective potential rises to infinity at θ = 0 and at θ = π,
indicating that any motion of the pendulum involves angles between
these limiting values. The effective potential has a minimum at a point
whose location depends on pφ . This minimum gets closer and closer to
π/2 as pφ increases. For a given value of pφ the pendulum swings in
a circle about the vertical axis, while undergoing oscillations in θ. The
simplest motion of this type occurs when θ is constant and equal to its
value at the minimum in Vef f . This is, of course, a conical pendulum.
To treat the general case of the spherical pendulum, it is convenient to
start the analysis with the conical pendulum.
Effective Potential vs Theta
90
Effective Potential, V (arbitrary units)
80
70
10
60
50
40 8
30
6
20
10 4
0 2
-10
0 0.5 1 1.5 2 2.5 3
Theta (radians)
p2φ
dVef f d
= + mgl cos θ = 0, (12.26)
dθ θ0 2ml2 sin2 θ
dθ θ0
2
pφ d
−2
0 = sin θ − mgl sin θ ,
2ml2 dθ θ0
p2φ cos θ0
0 = − − mgl sin θ0 . (12.27)
ml2 sin3 θ0
Consequently,
Angle
π/2
θ1
θ2
θ = θ0 + θ1 . (12.31)
cos θ
θ̈ − β 2 − (g/l) sin θ = 0. (12.32)
sin3 θ
cos θ0 g
β2 3 + sin θ0 = 0, (12.33)
sin θ0 l
and
3 cos2 θ0
1 2g
θ̈1 + + β − cos θ0 θ1 = 0. (12.34)
sin2 θ0 sin4 θ0 l
Equation (12.33) gives
g sin4 θ0
β2 = − .
l cos θ0
Inserting this expression into equation (12.34) yields
1 + 3 cos2 θ0 g
θ̈1 − θ1 = 0. (12.35)
cos θ0 l
But this equation has the form θ̈1 +ω 2 θ1 = 0 (simple harmonic motion)
where
r
g β p
ω= − (1 + 3 cos2 θ0 ) = 1 + 3 cos2 θ0 (12.36)
l cos θ0 sin2 θ0
3Note that
cos(θ0 + θ1 ) = cos θ0 cos θ1 − sin θ0 sin θ1
=
˙ cos θ0 − θ1 sin θ0 ,
and
sin(θ0 + θ1 ) = sin θ0 cos θ1 + cos θ0 sin θ1
=
˙ sin θ0 + θ1 cos θ0 ,
where we used sin θ1 =θ
˙ 1 and cos θ1 =1.
˙
12.4. THE SPHERICAL PENDULUM 443
12.5. Summary
After studying this chapter you will probably agree that the “simple
pendulum” is not so simple after all! One of the benefits of studying
the motion of a pendulum is that it leads to a consideration of a variety
of mathematical techniques.
The period of a simple pendulum with arbitrary amplitude is
4 π/2
Z
dφ
P = p ,
ω 0 1 − k 2 sin2 φ
a complete elliptic integral of the first kind. These have been tabulated
but they can also be evaluated using a series expansion.
An analysis of the physical pendulum introduces a number of prop-
erties of extended bodies, including the radius of gyration k defined in
terms of the moment of inertia as
I = M k2.
A simple pendulum oscillating with the same frequency as a physical
pendulum has a length given by
l = k02 /h,
446 12. THE PENDULUM
where h is the distance from the axis of rotation to the center of mass,
and k0 is the radius of gyration relative to the axis. The distance l is
also the distance from the axis of rotation to the center of oscillation.
If a physical pendulum receives an impulse at the center of percussion,
there will be no reaction at the axis. The center of percussion lies at
the same point as the center of oscillation (but conceptually these two
points are quite different).
The motion of a spherical pendulum is most easily understood by
introducing an effective potential
p2φ
Vef f = + mgl cos θ.
2ml2 sin2 θ
For the special case of a conical pendulum the angle θ = θ0 is constant
and the value of pφ is also constant, leading to an azimuthal angular
velocity of p
φ̇ = −g/(l cos θ0 )
and a total energy of
1 mgl
Ecp = (2 − 3 sin2 θ0 ).
2 cos θ0
The spherical pendulum presents a number of difficulties, but it
warrants study because it can be treated as a conical pendulum sub-
jected to a small perturbation about a known solution. (This technique
was used in Chapter 10 to study the motion of a slightly perturbed
planet. It will be used again in the next chapter.) You will become
very familiar with perturbation techniques when you study quantum
mechanics. A conical pendulum which has been perturbed so the am-
plitude of oscillations in θ is , has an azimuthal angular position given
by
β 2 2π
φ= 2 t − sin ωt cot θ0 ' √ ,
sin θ0 ω 1 + 3 cos2 θ0
and the oscillations in θ have an angular frequency
p
ω = (β/ sin2 θ0 ) 1 + 3 cos2 θ0 ,
where the parameter β is defined by
g sin4 θ0
β2 = − .
l cos θ0
12.6. Problems
Problem 12.1. The rotational analogue to F = dp dt
is N = dL
dt
.
Starting with this equation, obtain a work-energy theorem for rota-
tional motion.
12.6. PROBLEMS 447
θ
R
2b
B
A
J
Figure 12.13. A rod of mass m and length 2b lies on a
smooth horizontal table.
and r
ml2
Z
dθ
t= i1/2 .
2 h p2φ
E − 2ml2 sin2 θ − mgl cos θ
(The first of these integrals can be put into the form of an elliptic
integral of the third kind and the the second can be expressed as an
elliptic integral of the first kind.)
COMPUTATIONAL PROJECTS
Computational Project 12.1. Write a program to evaluate el-
liptic integrals of the first kind.
Computational Project 12.2. Use the results of Computational
Project 12.1 to write a program that evaluates the period of a simple
pendulum with arbitrary amplitude. Vary the amplitude and obtain a
plot of period as a function of amplitude.
Computational Project 12.3. Plot position vs. time for several
points on a rod undergoing an impulse, as illustrated in Figure 12.4.
Assume the impulse is J = 10 Ns, the rod has mass M = 2 kg, and
length L = 0.5 m.
Chapter 13
Accelerated Reference Frames
By now you have heard me say many times that Newton’s laws of
motion are applicable only in inertial (non-accelerated) frames of refer-
ence. The question arises: How do we deal with motion in non-inertial
reference frames? After all, most real reference frames are accelerating,
and for many of them, the acceleration cannot be neglected. In fact,
since we live on the surface of large rotating sphere, it is important for
us to be able to solve physics problems in non-inertial systems.
In this chapter you will learn how to express Newton’s second law
in a non-inertial (accelerating) reference frame. We are, of course, par-
ticularly interested in rotating coordinate systems, which are a prime
example of accelerating reference frames. Our study of these systems
will introduce you to “fictitious” forces such as the centrifugal force and
the Coriolis force. The Coriolis force, as you may know, is responsi-
ble for a number of important geophysical processes such as hurricanes
and ocean currents. Finally, we will consider the motion of a very long
simple pendulum and show that the plane of this “Foucault” pendulum
precesses due to the rotation of the Earth.
ro ro'
O r O'
because they are really just mass times acceleration terms which have
been moved to the other side of the equation. Well known fictitious
forces include the centrifugal force and the Coriolis force.
z z'
Ω
y'
y
x x'
ai = ar + 2Ω × vr + Ω × (Ω × r).
Newton’s second law is valid in the inertial frame, so ai = F/m where F
is the net external (“physical”) force acting on the particle. Replacing
ai and rearranging leads to
Ωxr
Ω -Ω x Ω x r
r
= −ρ̂r(Ωb + Ω0 )2 ,
in agreement with the result of part (b).
However, that would only be true if the Earth were not rotating. On
a rotating Earth, the centrifugal force causes the plumb bob to be de-
flected slightly away from the line to the center. The proof is fairly
simple. You may want to try it on your own before reading my deriva-
tion.
As shown in Figure 13.4, the “real” forces acting on the bob are
the tension in the string and the gravitational attraction of the Earth.
That is,
ME m
F = T−G 2 r̂,
RE
where RE is the radius of Earth. Inserting this expression for the force
into the “F = ma” equation in form (13.6) yields
ME m
T−G 2
r̂−m [Ω × (Ω × r)] =mar = 0,
RE
Here I used the fact that the velocity and acceleration of the plumb
bob in the rotating frame are zero.
Ω T
ac
fc
Fg g
resultant = ge
ac
This relation indicates that the tension is not parallel to mg and thus
it is not directed along the line to the center of the Earth.
It is usually convenient to include the centrifugal acceleration into
the gravitational acceleration. The “effective” gravitational accelera-
tion is defined by
ge = g − Ω × (Ω × r),
and the equation reduces to
T+mge = 0.
Therefore the tension is directed opposite to ge rather than opposite to
g. See Figure 13.4. In the Northern hemisphere ge = g − Ω × (Ω × r)
points slightly below the center of the Earth while in the Southern
hemisphere it points somewhat above the center.
The centrifugal force is responsible for the fact that the Earth is not
a sphere; it has a “bulge” at the equator. To see how this comes about,
consider a particle on the surface of a perfectly spherical Earth. This
particle is subjected to the gravitational force, a normal force and the
centrifugal force. If you draw a force diagram for this situation, you
will find that the centrifugal force is an unbalanced force. Resolving
the centrifugal force into components parallel and perpendicular to the
Earth’s surface, you will find that the perpendicular component simply
adds to the normal force, but the unbalanced parallel component points
toward the equator. Thus, the centrifugal force is responsible for the
fact that the Earth is an oblate spheroid rather than a sphere. You
can easily convince yourself that if the Earth’s equatorial bulge is taken
into consideration there will be no unbalanced forces on the particle on
the surface. This argument also tells you that although a plumb line
does not point towards the center of the Earth, it is perpendicular to
the Earth’s surface, so it does define a local perpendicular.
Exercise 13.3. Draw the force diagram for a particle on the surface
of a perfectly smooth, spherical, rotating planet and show that the
particle accelerates towards the equator. Do this for both the Northern
and Southern hemispheres. Draw the same sort of sketch for a planet
with an equatorial bulge and demonstrate that in this case there is no
net force on the particle.
13.5. THE CORIOLIS FORCE 461
λ
λ
x
The time required for the particle to reach the ground, tf , is the value
of t when z = 0, that is,
p
tf = 2h/ge .
Note that none of the above relations include the Coriolis effect.
Now include the Coriolis acceleration, given by:
ac = −2Ω × vr .
You cannot evaluate ac exactly because you do not know the value of vr
which itself depends on the Coriolis acceleration. However, you know
the zeroth order approximation to the velocity. It is vr = − ge tk̂. The
first order correction is obtained by including the Coriolis term in the
acceleration but using the zeroth order velocity. That is, a =ge k̂ + ac
where
ac = −2Ω × vr
h i h i
= −2 Ω cos λk̂−Ω sin λı̂ × −ge tk̂
= 2ge tΩ sin λ̂.
Note that for the falling body, the Coriolis acceleration is in the positive
y direction, that is, towards the East. The eastward component of the
velocity (vy ) is initially zero and at any later time it is obtained by
integrating the expression above. Therefore,
vy = ge t2 Ω sin λ.
The eastward deflection (y) is obtained by integrating again.
1
y = ge t3 Ω sin λ.
3
To determine the total eastward deflection whenp the object is dropped
from height h, substitute the time of flight (t = 2h/ge ) into this last
expression.
The results obtained are the first order solution. As mentioned
above, these results can be plugged back into the equation of motion
to obtain the second order solution, and so on.
Students are often mystified by this result. The eastward deflection
is obviously not due to the Earth rotating under the falling body - any-
way, that would give a westward deflection! A physical explanation for
the horizontal (eastward) deflection of a dropped body comes from the
conservation of angular momentum. Let l be the angular momentum
of the particle in the inertial frame. There is no torque acting on it, so
l is constant. Initially the object is at position r (where r is a vector
464 13. ACCELERATED REFERENCE FRAMES
from the center of the Earth). In the inertial frame the body has an
eastward linear velocity v. (Since the body is at rest with respect to
.
the surface of the Earth, this initial velocity is v = Ω × r =ΩRe sin λ̂.)
The angular momentum of the particle is
l =mr × v.
For l to remain constant while r decreases (as the particle falls), the
velocity v must increase. Therefore, in the Earth based reference frame,
the object acquires a velocity towards the East.
Another strange and interesting effect of Coriolis forces is the fact
that if you throw an object straight up, it does not land at the point
from which it started. (You will prove this in problem 13.6.)
ay = 2Ω sin λge t
az = 2Ω2 sin2 λge t2 − ge
Consequently,
Z
2
vx = ax dt = Ω2 cos λ sin λge t3
3
and
1
x = Ω2 cos λ sin λge t4
6
Using the first order value for the time of flight we obtain
2 h2
x = Ω2 cos λ sin λ .
3 ge
13.5. THE CORIOLIS FORCE 465
Note that the time of flight will actually be slightly longer because
the Coriolis acceleration opposes the gravitational acceleration.
Comparing the southward and eastward deflections we have
x (2/3)Ω2 cos λ sin λ(h2 /ge )
= Ω cos λ
= p .
−y (1/3)Ω sin λ(2hge )1/2 2ge /h
Since Ω is about 10−5 and cos λ is near unity and the denominator
will be in the range of 1 to 10, we appreciate that the southward
deflection is significantly smaller than the eastward deflection.
Exercise 13.5. A stone is dropped from a height of 50 meters.
How far is it deflected towards the East? Assume the latitude is 60◦
N. Answer: 0.39 cm.
θ
T Path of pendulum bob
y
ψ
vh
x
mg
Projection of path onto xy plane
The pendulum bob is acted upon by two real forces: The tension in
the string (T) and gravity (mg). It is also acted upon by two fictitious
forces: the Coriolis force (−2mΩ × v) and the centrifugal force. For
convenience assume the centrifugal force is absorbed into mg.
Therefore,
ma = T+mg−2m(Ω × v).
In component form this relationship becomes
max ı̂+may ̂+maz k̂ = −T sin θ cos ψı̂−T sin θ sin ψ̂ + T cos θk̂−mg k̂
h i h i
−2m Ω cos λk̂ − Ω sin λı̂ × vx ı̂+vy ̂+vz k̂ .
That is,
The reason for doing this is because I want you to see several techniques
for solving coupled differential equations.
A Neat Trick
The set of coupled equations (13.11) can be solved very nicely by
introducing the complex quantity
ζ = x + iy.
√
Multiplying the second equation by i (= −1 ) and adding the two
equations gives
ζ̈ + 2iK ζ̇ + ω 2 ζ = 0,
which resembles the equation of motion for a damped harmonic oscilla-
tor. This linear differential equation is solved in the usual way, assum-
ing ζ = Aept and obtaining the “secular” or “characteristic” equation
p2 + 2iKp + ω 2 = 0,
with solution √
p = −iK ± i K 2 + ω 2 .
Now p K = Ω cos λ where Ω is the rotation rate of the Earth, whereas
ω = g/l is the angular frequency of the pendulum. So K << ω and
the K 2 under the root can be neglected. There are two possible values
for p, so the solution is
ζ = Ae−i(K+ω)t + Be−i(K−w)t (13.12)
where A and B are complex constants. Writing the complex constants
in the form (a + ib) and (c + id) the solution is
ζ(t) = (a + ib)e−i(K+ω)t + (c + id)e−i(K−w)t . (13.13)
This complex solution is expressed in terms of real quantities, but it
is still not obvious what the motion looks like. The following example
will help you to visualize the motion described by equation (13.13).
3You probably remember that this is the procedure we used to solve the coupled
oscillators problem. (See section 11.5.)
4It is interesting to note that the Foucault pendulum problem has many par-
allels to a seemingly completely unrelated problem, namely the violation of CP
symmetry in the decay of an elementary particle called the neutral kaon. This
illustrates the basic unity of physics. To explore further, see the paper “Classical
Illustration of CP Violation in Kaon Decays” by Jonathan L. Rosner and Scott A.
Slezak, Am. J. Phys., 69, 44-49, 2001.
13.6. THE FOUCAULT PENDULUM 473
y*
s cos(Kt)
y x
s sin(Kt)
Kt
s x*
13.7. Summary
A rotating reference frame is an accelerated reference frame. The
time derivative of a vector in a rotating frame is related to the time
derivative in an inertial frame by the operator equation
d d
= +Ω×.
dt inertial dt rot
The acceleration ar relative to a rotating frame is obtained by applying
the operator equation twice to the position vector, yielding:
ar = F/m − 2Ω × vr − Ω × (Ω × r),
that is often written as
13.8. Problems
Problem 13.1. A child runs on a merry-go-round with velocity V
with respect to the merry-go-round. Show that
dr
= V + (ωe +ω mgr ) × r
dt I
where ωe is the angular velocity of Earth and ωmgr is the angular ve-
locity of the merry-go-round.
Problem 13.2. A wedge of mass M and angle α is resting on a
frictionless horizontal plane. A rectangular box of mass m is on the
wedge. Since all surfaces are frictionless, the box will slide down the
wedge and the wedge will slide in the opposite direction on the plane.
So far, this is the same as Problem 4.7 which you may have solved.
There you used the Lagrangian technique to solve the problem and
found that the horizontal acceleration of the wedge is
!
m g sin α
Ẍ = − m cos α.
m + M 1 − m+M cos2 α
Here you are asked to find the horizontal component of the accelera-
tion of the box relative to the wedge. Solve the problem by applying
Newton’s second law directly, but using the fact that the wedge is ac-
celerating so the motion of the box relative to the wedge is motion in
an accelerated reference frame.
Problem 13.3. Obtain an equation analogous to equation (13.5)
but including the orbital motion of the Earth around the Sun. You may
assume the two rotation vectors are constant and parallel and that the
Earth’s orbit is circular. Compare the magnitude of any additional
term to the Coriolis term in equation (13.5).
Problem 13.4. Assume the Earth is an inertial coordinate system.
A child on a rotating platform runs (a) radially towards the edge. (b)
Runs in the azimuthal direction. Determine the fictitious forces acting
on the child in both cases.
Problem 13.5. A wheel of radius a rolls at constant speed V in
a circular path of radius R. Assume the wheel is perpendicular to the
plane of the path and that the surface of Earth is an inertial reference
frame. Consider a point at the top of the wheel. Determine the Coriolis
acceleration, and the centrifugal acceleration this point .
Problem 13.6. A particle is projected vertically upward with an
initial velocity v0 . Determine, to first order, the point where it lands
relative to the point where it was thrown.
476 13. ACCELERATED REFERENCE FRAMES
In this part of the book (Chapters 14 through 18) you will study
various aspects of extended material bodies. This chapter involves
a number of advanced concepts in statics. We will begin with a few
definitions and two simple theorems concerning systems of forces acting
on rigid bodies, then go on to analyze the statics of freely deformable
bodies such as a string or cable hanging from stationary supports.
This is followed by definitions of stress and strain and a generalization
of Hooke’s law. An important application is an investigation of the
properties of a fluid in equilibrium. The last topic is d’Alembert’s
Principle and the concept of virtual work. You will see how this is
related to the equilibrium of an extended body, and how the principle
can be used to derive Lagrange’s equations.
T1
T2
Mg
2M
2Mg
N
μN
m g cosθ
m g s in θ
θ mg
slips. Let the coefficient of static friction between spool and plane
be µ = 0.2. Determine the angle at which the spool begins to slip.
Solution: Forces up the plane equal forces down the plane so
if N is the normal force,
mg sin θ = T + µN.
The forces perpendicular to the plane also add to zero,
N = mg cos θ.
Clockwise torques equal counterclockwise torques. Taking torques
about the point of contact between spool and plane:
T (2R) = mgR sin θ,
1
T = mg sin θ,
2
where R is the radius of the spool. Since T = mg sin θ − µN =
mg sin θ − µmg cos θ we have
1
mg sin θ = mg sin θ − µmg cos θ,
2
tan θ = 2µ = 0.4,
θ = 21.8◦ .
Here are a few definitions and some concepts that are useful when
treating statics problems. Couple: A couple is a system of forces
P
whose sum is zero, i.e., Fi = 0. Clearly a couple cannot generate a
linear acceleration, but it can exert a torque on a body, as is obvious
from an inspection of Figure 14.4. Although a couple may be a system
of many forces, two forces are sufficient to describe it.
Exercise 14.6. A 1 meter rod lying along the x axis is acted upon
by the following three forces: A force F1 = 10̂ N at the left end, a
force F2 = −20̂ N at the midpoint, and a force F3 = 5ı̂ N at the right
end. Determine an equivalent system of two forces.
A B
Td
y D θ
Tc C
wx
x
O
The lines of action of the three forces (Tc , Td and wx) pass through
the same point (by Theorem 1 of section 14.1). By symmetry this point
is at x/2. The sum of the forces is zero so
Td sin θ = wx,
Td cos θ = Tc .
Dividing one equation by the other,
wx
tan θ = .
Tc
14.4. THE HANGING CABLE 489
Exercise p
14.7. Show that the tension at any point in the cable is
given by T = Tc2 + w2 x2 .
y
T
B
θ
A
s
T0 W=ws
C
x
force per unit length w thenRthe total “body force”1 acting on the piece
s
of rope of length s will be 0 wds. In equilibrium, the vector sum of
the three forces acting on the segment is zero. Equating the vertical
and the horizontal components of these forces leads to
T cos θ = T0 ,
T sin θ = ws.
Therefore
w
tan θ = s,
T0
or
T0
s= tan θ = c tan θ, (14.1)
w
where c = w/T0 and has the units of length. (In a little while I will
give you a geometric interpretation of c as the vertical distance from
the origin to the lowest point in the rope.)
If you are very good at analytical geometry, you might have recog-
nized equation (14.1) as the intrinsic equation of a catenary. In this
sense, the problem is solved, but most people would prefer to have a
formula for the curve in Cartesian coordinates x, y.
Note that the Cartesian coordinates shown in Figure 14.7 have the
y axis passing through the lowest point of the curve, but the origin is
displaced vertically from this lowest point.
Equation (14.1) can be written as
dy s
tan θ = = .
dx c
Consequently,
p
d2 y 1 ds 1 dx2 + dy 2
= = ,
dx2 csdx c dx
2
1 dy
= 1+ .
c dx
dy
Let p = dx
and write this relation as
dp 1p
= 1 + p2 ,
dx c
dp 1
p = dx.
1 + p2 c
1A
“body force” is a force that is proportional to the amount of matter being
considered. Thus, for example, the tension in the rope would not be a body force,
but the weight of the rope depends on the amount of rope and is a body force.
14.4. THE HANGING CABLE 491
Integrating
x
= sinh−1 p + constant.
c
Since the slope (p) is zero at x = 0, the constant of integration is zero.
Therefore, x
p = sinh .
c
But p = dy/dx, so
dy x
= sinh .
dx c
Integrating again leads to
x
y = c cosh + constant.
c
If the origin is placed a distance c below the lowest point, then y = c
at x = 0 and the constant of integration is zero. Thus the equation of
the curve formed by the hanging rope is
x
y = c cosh , (14.2)
c
which is the equation of a catenary in Cartesian coordinates.
The tension at any point in the rope is obtained from
T0 T0 ds
T = = dy = T0 .
cos θ ds
dy
But s = c sinh(x/c) so
x
T = T0 cosh .
c
Furthermore, y = c cosh xc so
T0
T = y,
c
and finally
T = wy.
That is, the tension at any point in the rope is directly proportional to
the height of that point. The maximum tension occurs at the highest
points, that is, at the end points. For a rope whose end points are at
the same height and are separated by a horizontal distance a (called
the “span”) the end points are at x = ±a/2, and y = h + c, where h
(the “sag”) is the vertical distance from the points of support to the
lowest point in the rope. The maximum tension is then
a
Tm = w(h + c) = wc cosh .
2c
492 14. STATICS (OPTIONAL)
Exercise 14.9. A rope of length 3 meters and mass per unit length
of 0.2 kg/m is hanging under its own weight. The endpoints are lo-
cated at (-1,1.6) and (1,1.6). Numerically determine the parameter c.
Determine the maximum value of the tension. Answer: c = 0.62.
Tension θ
Compression Shear
tabulated in various places, such as the CRC Tables.2 Hooke’s law for
a long thin rod of length l is
F ∆l
=Y , (14.3)
A l
where F is the tension, A is the cross-sectional area of the rod and Y is
a quantity characteristic of the composition of the rod called Young’s
modulus.
For a compression, Hooke’s law is
F ∆V
= −B , (14.4)
A V
where B is called the bulk modulus. In this case, the quantity F/A is
usually called the pressure, as it corresponds to the notion of pressure
as a force acting across an area that tends to cause a compression. The
negative sign in the law emphasizes that the stress is a compression.
Finally, for a shear, Hooke’s law is
F
= n tan θ, (14.5)
A
where n is called the shear modulus.
The Stress Tensor
You have certainly noticed that the left hand side of the Hooke’s
law equations all have the same form although they represent different
situations, specifically the direction of the force and the orientation of
the surface relative to the force. It turns out that we can put all of
these into a single expression by defining a quantity called the stress
tensor.
We are considering forces that act across surfaces. In general, the
material on one side of a real or imaginary surface exerts a force on the
material on the other side. For example, the molecules on one side of an
imaginary plane in a material body will exert forces on the molecules
on the other side. Of course, we are usually interested in real surfaces
with different materials on either side. The forces exerted across a
surface, are called stresses. They can be a tension, a compression or a
shear.
The force acting across a surface can have an arbitrary direction
and in general is given by
F =Fx ı̂+Fx ̂+Fz k̂.
2CRC Tables of Chemistry and Physics, 87th Edition, CRC Press, Boca Raton,
Fl. 2006.
14.6. THE CENTROID (OPTIONAL) 495
A s
disk be rotated about an axis lying on the straight side of the disk, as
shown in Figure 14.10. The volume of revolution generated is a sphere
of volume V = (4/3)πa3 . But according to theorem 6, this volume is
equal to the area of the disk times 2πY where Y is the distance from
the axis to the centroid of the shaded disk. That is
3 1 2
(4/3)πa = (Area of disk) 2πY = πa 2πY
2
4a
∴ Y = .
3π
all of the mass of the body can be concentrated to experience the same
gravitational force as the extended body.
As an example, think of a very long rod. The center of mass of the
rod does not depend on its orientation. But the center of gravity of
the rod will depend on whether the rod is held horizontally, parallel to
the surface of the Earth, or vertically, perpendicular to the surface of
the Earth. If the rod is horizontal, the center of mass and the center of
gravity will coincide; if the rod is vertical, the center of gravity will be
a little bit below the center of mass. This is because the gravitational
force decreases with distance from the center of the Earth and those
parts of the rod nearest the Earth are attracted a bit more strongly
than the points of the rod that are further away. Of course, the rod
would have to be extremely long for the distance between the two points
(CG and CM) to be at all noticeable.
The center of gravity of an extended body of mass M can be defined
in terms of its interaction with a particle (say m). The extended body
and the particle are illustrated in Figure 14.11. Suppose the extended
body shrinks to a point. Where should this point mass be located to
experience the same force as the extended body? The location of this
hypothetical point mass is called the center of gravity of M relative to
m. In Figure 14.11 it is indicated by the symbol CG.
m m
-F
CG M
M CG
force due to the pressure on the bottom surface, and (3) the downward
force of gravity on the fluid. Let the bottom of the cylinder be at z
and the top at z + dz. Then the equilibrium condition
force up = force down
is expressed by
p(z)A = p(z + dz)A + ρgdzdA.
Using the definition of derivative, this expression can be written
dp
= −ρg. (14.6)
dz
That is,
dp
k̂ = f .
dz
It is easy to generalize this expression to obtain an expression for the
change in pressure with respect to a translation in an arbitrary direc-
tion, leading to
∇p = f . (14.7)
That is, the gradient of the pressure is equal to the body force. This
implies that the surfaces of constant pressure are perpendicular to the
body force.
Using the vector definition of differential, we have
Z Z Z
∇p · dr = dp = f ·dr,
and so Z r
p(r) = p(r0 ) + f ·dr, (14.8)
r0
where the integral is along any path from r0 to r in the fluid. Note
that for a fluid in equilibrium, equation (14.8) defines a pressure field
(the pressure is defined at every point in the fluid). Furthermore, the
equation states that pressure changes are balanced by the body force,
f.
Equation (14.8) also tells us that if the body force is constant,
increasing the pressure at one point in a fluid increases the pressure by
an equal amount at all other points in the fluid. This is called Pascal’s
law. For example, if p(r0 ) is the pressure at the surface of the fluid,
then an increase in the surface pressure is communicated to all points
in the fluid.
To determine the pressure at any point in a fluid in equilibrium,
you can apply equation (14.8). For example, if a fluid is acted upon
502 14. STATICS (OPTIONAL)
An Ideal Gas
The volume of a gas depends on the pressure exerted on it. The
basic relation, based on Hooke’s law (stress ∝ strain) is given by equa-
tion(14.4). If the pressure is increased by an infinitesimal amount dp
the volume will decrease by dV and equation (14.4) can be expressed
as
dV dp
− = .
V B
504 14. STATICS (OPTIONAL)
m1 m2
θ1 θ2
Furthermore,
d ∂xi ∂xi d ∂xi
mi ẋi = mi ẍi + mi ẋi . (14.13)
dt ∂qj ∂qj dt ∂qj
The first term in this equation contains ∂xi /∂qj . A useful relationship
involving generalized coordinates is
∂xi ∂ ẋi
= . (14.14)
∂qj ∂ q̇j
The last term in equation (14.13) involves the expression
d ∂xi ∂ ẋi
= . (14.15)
dt ∂qj ∂qj
Using equations (14.13), (14.14) and (14.15), equation (14.12) becomes,
X X d ∂xi
∂ ẋi
ṗi δxi = mi ẋi − mi ẋi δqj ,
i i,j
dt ∂q j ∂q j
X d ∂ ẋi
∂ ẋi
= mi ẋi − mi ẋi δqj . (14.16)
i,j
dt ∂ q̇j ∂qj
and
X
∂T ∂ X 1 2 ∂ ẋi
= mi ẋi = mi ẋi .
∂ q̇j ∂ q̇j i 2 i
∂ q̇j
X X d ∂T ∂T
ṗi δxi = − δqj .
i j
dt ∂ q̇j ∂qj
14.10. SUMMARY 509
Using the fact that the δqj are independent it is clear that each term
in the summation must individually be equal to zero, so
d ∂T ∂T
− = Qj . (14.17)
dt ∂ q̇j ∂qj
This derivation gives Lagrange’s equations in a form involving the gen-
eralized forces and the kinetic energy. You have seen this alternate
form of Lagrange’s equations previously. (See equation 4.12 in Worked
Example 4.6.)
14.10. Summary
This chapter on the statics of extended bodies started with a simple
analysis of the conditions for static equilibrium, namely
X
Fi = 0,
X
Ni = 0.
That is, the sum of the forces is zero and the sum of the torques is
zero. You were then exposed to a number of theorems and definitions
to familiarize you with the concepts and terms used in the study of
statics. Thus, for example, you were given the definitions of a couple,
a resultant, and an equilibrant. You found out that any system of
forces can be reduced to a single force plus a couple.
The next topic was more advanced, being a study of a hanging
cable. There were two applications, namely, a suspension bridge and a
510 14. STATICS (OPTIONAL)
rope hanging under its own weight. The shape of the cable suspending
a bridge is a parabola,
w 2
y= x + c,
2Tc
and the shape of a rope hanging under its own weight is a catenary
x
y = c cosh .
c
Hooke’s law states that stress and strain are proportional. Defining
stress as force per unit area, leads to the three expressions
Law Stress Proportionality Constant
F ∆l
A
=Y l tension Y = Young’s modulus
F ∆V
A
= −B V pressure B = Bulk modulus
F
A
= n tan θ shear n = Shear modulus
A consideration of the orientation of the surface across which the
force acts led to the concept of the stress tensor which was described
but not utilized in the analysis.
The centroid is a geometrical concept similar to the physical concept
of center of mass. The centroids of a volume, a surface and a line are
defined as
Z
1
Rvol = rdτ,
V
Z
1
Rsurf = rdA,
A
Z
1
Rline = rds.
S
Some properties of centroids were summarized in the two theorems of
Pappus.
The center of gravity is the point in an extended body of mass M
at which a point mass M exerts the same gravitational force on a point
mass m as does the extended body.
The next topic was a brief study of the equilibrium of a fluid, in
which you found that the pressure field is related to the body force
f =ρg by
f = ∇p.
Two important applications are Archimedes’ principle and Pascal’s law
which were applied to fluids in equilibrium and to the atmosphere of
Earth.
14.11. PROBLEMS 511
T1 LL
L
T2
α
Mg Mg
Problem 14.3. A smooth rod that is nailed to the floor and to the
wall, makes an angle of 30o to the horizontal, as shown in Figure ??.
A ring of mass m can slide on the frictionless rod. A string is passed
through the ring; one end of the string is attached to a point on the
floor and the other end of the string supports a weight of mass 3m.
Determine the angle φ between the two segments of the string. Note,
the rod does not move, but the ring can slide along it.
Problem 14.4. A rigid flat plate 2 meters on a side is acted upon
at its corners by the following forces: F1 = 10̂ acting at (-1,1), F2 =
5ı̂−5̂ acting at (1,-1), F3 = 10̂ acting at (-1,-1), and F4 = 10ı̂ at
(1,1). The forces are in newtons. The center of the plate is at (0,0).
Determine the resultant and the equilibrant.
512 14. STATICS (OPTIONAL)
m 3mg
30o
ϕ 3m 4mg
S S'
Problem 14.7. Consider the rod of Figure 14.1. If the string with
tension T2 breaks, the rod will rotate and translate until it is hanging
straight down. Determine the equilibrant of the forces acting on the
rod at the instant the string breaks.
Problem 14.8. A ladder of weight W and length l stands on a
smooth floor and leans against a smooth wall. The angle between the
ladder and the floor is α. As you might expect, since there is no friction,
the ladder is slipping. What is the equilibrant? (The expression you
will obtain includes the normal forces)
14.11. PROBLEMS 513
Problem 14.10. A cube of side a has one corner at the origin and
the diagonally opposite corner at (a, a, a). The sides are oriented along
the axes. The forces acting on the cube are all equal in magnitude to
F. The force on the corner at (0, 0, 0) is Fı̂, the force on the corner at
(0, a, a) is F̂ and the force on the corner at (a, a, a) is F k̂. Find an
equivalent single force and couple. (See Figure 14.14.)
z
a
a
x
and w varies along the length of the cable. (a) Show that such a cable
forms a curve given by
x
y = k log sec + c
k
where c is a constant of integration. (b) Show the maximum possible
span for such a cable is πk.
Problem 14.13. Using the Taylor’s series expansion for the po-
tential, show that it leads to Hooke’s law for small displacements from
equilibrium. (For simplicity, you may assume a one dimensional defor-
mation.)
Problem 14.14. Obtain a formula for the volume of a torus. (Hint:
Use a theorem of Pappus.)
Problem 14.15. Two uniform spheres of mass M are located at
x = ±a. Determine the center of gravity relative to a point P on the
y-axis. Show that the center of gravity and the center of mass approach
one another as the distance to P increases.
Problem 14.16. A satellite consists of two spheres (which we ap-
proximate as particles) each having mass m. They are connected by a
massless rod of length l. The satellite is in a circular orbit of radius
r about the Earth with both masses lying along the same radial line
from the center of Earth. Determine the distance h between the center
of mass and the center of gravity of this satellite.
Problem 14.17. A cylindrical rod of length 20 km is held per-
pendicular to the surface of Earth. Determine the distance between
its center of mass and center of gravity. (Assume the Earth is a point
mass located at its center.) Hint: Use the binomial expansion.
Problem 14.18. The Earth is an oblate spheroid so its equato-
rial radius is somewhat greater than its polar radius. The following
“thought experiment” will allow you to estimate the difference between
these two radii. Imagine a tunnel is drilled from the North Pole to the
center of the Earth where it makes a ninety degree turn and emerges
at the equator. The tunnel is filled to the brim with water. How much
longer is the equatorial tunnel than the polar tunnel if the density of
the Earth is assumed constant? Next solve the problem if the density
of the Earth varies linearly with depth and that the density at the
center is ρ0 .
Problem 14.19. Under some circumstances the lower atmosphere
can be considered adiabatic. Assuming an adiabatic atmosphere, the
14.11. PROBLEMS 515
Problem 14.22. Assume that the two planes of Figure ?? are not
smooth, and the coefficient of friction is µ. The system is in equilib-
rium. Use the principle of virtual work to obtain an expression for the
coefficient of friction.
COMPUTATIONAL PROBLEMS
Computational Techniques: Solving a System of Linear Equations
Statics problems often involve solving a system of linear equations.
Computer Algebra Systems, such as Matlab, usually have built in func-
tions that allow one to solve such equations very simply. For example,
consider the system of three equations in three unknowns
a11 x1 + a12 x2 + a13 x3 = b1 ,
a21 x1 + a22 x2 + a23 x3 = b2 ,
a31 x1 + a32 x2 + a33 x3 = b3 .
Here the a0 s and b0 s are known quantities and one wishes to determine
the x0 s. Using matrix notation, these equations can be written
a11 a12 a13 x1 b1
a21 a22 a23 x2 = b2
a31 a32 a33 x b3
or
Ax = B.
Then, in the syntax of Matlab,
x =A\B.
516 14. STATICS (OPTIONAL)
(Note the backslash. Without going into the details, the process of
solving for the vector x uses a technique called Gaussian elimination
that basically solves the first equation for x1 and substitutes it into the
remaining equations. Then it solves the second equation for x2 and
substitutes it into the remaining equations, and so on, until the last
x is obtained in terms of the coefficients. Finally, “back substitution”
generates the values of all the x0 s.)
Computational Project 14.1. In Figure 14.14 assume the lengths
of the rod and strings are 80 cm. Let the angle α be 15◦ and the let
the hanging rod have mass 12 kg. Determine T1, T2 , and the mass of
the rod.
Computational Project 14.2. Meteorologists often plot the
variation of meteorological quantities as a function of the logarithm
of the pressure, decreasing with altitude. Generate such a plot for the
temperature as a function of pressure of the atmosphere assuming the
temperature decreases adiabatically, that is that the pressure and tem-
perature are related by P 1−γ T γ = constant. You can let γ = 1.4 and
assume the pressure and temperature at the surface are 1 atmosphere
and 298 kelvin.
Computational Project 14.3. We found the equation for a rope
x
hanging under its own weight is y = c cosh c . Recall that
1 u
e + e−u .
cosh u =
2
u −u 1 u −u
Plot e and e and 2 (e + e ) to generate a catenary. Show that
the lowest point in the catenary is a distance c above the origin.
Chapter 15
Rotational Kinematics
This chapter and the next are a study of the general motion of a
rigid body. This is a fairly complicated topic which involves mathe-
matical concepts that you may not have encountered before.
We have already considered the rotation of a symmetrical body
about a fixed axis. We now generalize to rotations about a fixed point.
For example, consider a spinning top. A top is a symmetrical body that
rotates about its axis of symmetry. When it is spinning very fast, the
top remains upright, and as long as the top is not slipping, its center of
mass is at rest. But as the top slows down, it leans over. The center of
mass traces a circle around the vertical. The axis of rotation precesses
about this vertical. Eventually, the top will begin to nutate or “nod”
up and down as it precesses. If the top is not sliding, it has one fixed
point, namely the point that is in contact with the ground.
Another example of rotation about a fixed point is illustrated in
Figure 15.1 which shows a disk with an axle passing through it. In this
case, the axle is not perpendicular to the disk. If the system is not
supported in any way (imagine it is a poorly designed satellite) this
nonsymmetrical body will wobble while it rotates. The axis of rotation
does not have a fixed direction. If the center of mass is at rest, this is
a rotation about a fixed point.
Analyzing the motion of a body rotating about a fixed point involves
developing a system of rotational kinematics. Recall that in Chapter 2
you studied the kinematics of a point particle. This entailed learning
how to describe the position (as well as the velocity and acceleration) of
a particle in terms of Cartesian, cylindrical and spherical coordinates.
In a similar manner, you will now learn how to describe the orientation
of a rigid body and how this changes as the body rotates.
517
518 15. ROTATIONAL KINEMATICS
z' z
y'
x'
y
x
Figure 15.2. An inertial reference frame (x, y, z) and a
reference frame (x0 , y 0 , z 0 ) fixed in the body and rotating
with it.
0
Denote the unit vectors for the two systems by ı̂,̂, k̂ and ı̂0 ,̂0 , k̂ .
The direction of the x0 -axis with respect to the unprimed axes is given
by its direction cosines α1 , α2 , α3 , defined by
α1 = cos(x0 , x) = ı̂0 ·ı̂, (15.1)
α2 = cos(x0 , y) = ı̂0 · ̂,
α3 = cos(x0 , z) = ı̂0 · k̂.
520 15. ROTATIONAL KINEMATICS
write a23 , and so on, so that the transformation equations (15.3) are
expressed as
x01 = a11 x1 + a12 x2 + a13 x3 , (15.4)
x02 = a21 x1 + a22 x2 + a23 x3 ,
x03 = a31 x1 + a32 x2 + a33 x3 .
It is convenient to use the Einstein summation convention to express
equations (15.4) in the very compact form
x0i = aij xj . (15.5)
The Einstein summation convention states that if a term contains the
same subscript
P twice, a summation over that subscript is implied. Thus
aij xj = j aij xj .
The orthogonality relationships (Equations 15.2) are then simply
aij aik = δjk . (15.6)
Note that the transformation from the unprimed to the primed coor-
dinates is carried out using the coefficients aij . The aij are related to
each other by the orthogonality conditions given by Equation (15.6).
Therefore, the transformation (15.4) is called an “orthogonal transfor-
mation.”
You will show in Problem 15.1 that the orthogonality relation (15.6)
leads to
x0i x0i = xi xi . (15.7)
But xi xi = x2 + y 2 + z 2 is the square of the length of the vector
xı̂+ŷ+z k̂. So Equation (15.7) tells us that an orthogonal transforma-
tion conserves the length of a vector.
The coefficients aij lend themselves quite naturally to being ex-
pressed as a matrix, thus
a11 a12 a13
A = a21 a22 a23 . (15.8)
a31 a32 a33
It is convenient at this time to express vectors as column matrices. For
example, the position vector in the inertial system is denoted x and is
expressed as a column matrix as follows:
x x1
x = y = x2 .
z x3
522 15. ROTATIONAL KINEMATICS
y
y' P
x'
r
θ
θ x
Figure 15.3. The coordinates of point P in the two
reference frames.
Interpretation
The matrix A describes the transformation from the unprimed sys-
tem to the primed system. Let us change the notation of the position
vector back to r. If (x, y) and (x0 , y 0 ) represent the components of the
vector r in the two coordinate systems, then A tells us how the two
sets of components are related. That is, A describes the result of a
counterclockwise rotation of the coordinate system through the angle
θ. This is called the “passive” interpretation of A, and is illustrated in
Figure 15.3.
There is, however, a different interpretation in which A is considered
to be an operator that rotates the vector r in a clockwise manner to
generate a new vector r0 . In this interpretation, the coordinate system
is fixed and (x0 , y 0 ) are the components of the new vector, as shown in
Figure 15.4. This is, of course, called the “active” interpretation.
r
y
y' r'
x x'
Figure 15.4. The “active” interpretation assumes the
vector rotates in a clockwise sense rather than the coor-
dinate system rotating in the counter-clockwise sense.
That is, the elements of the inverse matrix are obtained by transposing
the rows and columns of the original matrix. The transposed matrix is
denoted by AT . Thus, for an orthogonal matrix,
AT = A−1 , (15.18)
and hence,
AT A = 1. (15.19)
In physics we are often interested in the matrix obtained by taking the
transpose of the original matrix and then replacing each element by its
15.2. ORTHOGONAL TRANSFORMATIONS 525
Orthogonal Matrices
Name Symbol Element Property
Orthogonal A aij |A|2 = 1
Inverse A−1 a−1
ij = aji A−1 A =1
Transpose AT aTij = aji AT A =1
Adjoint A† aij = a∗ji
Unitary A† = A−1
Hermitian A† = A
Next, carry out a rotation through θ about the ξ-axis. This causes
the ζ-axis and the η-axis to change direction as indicated in Figure
15.6. The axes resulting from this rotation are denoted ξ 0 , η 0 , ζ 0 . The
old xy-plane is now the ξ 0 η 0 -plane and is inclined to the original plane
by an angle θ. (The line from the origin along the ξ- (or ξ 0 -) axis is
called the line of nodes.)
ϕ y
x
Finally carry out a rotation about the ζ 0 -axis through ψ. This oper-
ation leaves ζ 0 unchanged but carries ξ 0 to x0 and η 0 to y 0 , as indicated
in Figure 15.7. The new coordinate system is x0 , y 0 , z 0 .
It is easy to appreciate that the three Euler angles φ, θ, ψ can be
used to carry the original axes x, y, z to any desired final orientation
x0 , y 0 , z 0 .
y
x
direction of the axis of rotation and the third gives the angle through
which the body is rotated.)
A rotation has two important properties: (1) The length of a vec-
tor is unchanged by a rotation, and (2) A vector lying on the axis of
rotation is unchanged in direction and magnitude. If the matrix A
describes a rotation, then property (1) will be satisfied by requiring A
to be an orthogonal matrix (see Equation 15.7). To find the conditions
for property (2) to be satisfied, consider a vector R lying along the
axis of rotation. Then R0 = R, where R0 is the vector in the rotated
system. But R0 = AR. Consequently,
AR = R. (15.33)
The matrix A represents a rotation from an initial orientation to some
final orientation. A may have been generated by the three rotations
through Euler angles, for example. What Euler’s theorem states is
that the same final orientation can be achieved by a single rotation
about some axis. If this is so, then a vector that is unchanged by the
rotation must lie on the axis of rotation. That is, there exists a vector
R satisfying Equation (15.33).
But (15.33) is just a special case of the more general equation
AR =λR. (15.34)
where λ is a real or complex scalar constant. Equation (15.34) is called
an eigenvalue equation; λ is called the eigenvalue and R is called the
eigenvector. The eigenvalue problem consists in finding the eigenvalues
and eigenvectors for a given matrix (or operator) A.
Euler’s theorem requires that AR = R. Therefore, the theorem is
proved if we can show that any orthogonal matrix representing a rota-
tion will have at least one eigenvalue equal to +1.
Let us write Equation (15.34) as
(A − λ1) R =0, (15.35)
and express R as a column vector,
X
R = Y .
Z
Expanding Equation (15.35),
(a11 − λ)X + a12 Y + a13 Z = 0, (15.36)
a21 X + (a22 − λ)Y + a23 Z = 0,
a31 X + a32 Y + (a33 − λ)Z = 0,
534 15. ROTATIONAL KINEMATICS
That is,
1 1 0 c1 c1
−1 1 2
0√ c2 = √ λ c2 .
0 0 2/ 2 c3 2 c3
Carrying out the matrix multiplications we obtain the following three
equations:
√
(1 − 2λ)c1 + c2 = 0, (15.41)
√
−c1 + (1 − 2λ)c2 = 0,
(1 − λ)c3 = 0.
The third equation yields λ = 1. The first two are coupled homogeneous
algebraic equations and have a nontrivial solution if and only if the
determinant of their coefficients is zero:
√
(1 − 2λ) 1
√
= 0.
−1 (1 − 2λ)
This leads to the secular equation for λ, namely
√
2λ2 − 2 2λ + 2 = 0.
Solving for λ yields
1
λ = √ (1 ± i) .
2
The three eigenvalues are
λ(1) = 1,
1
λ(2) = √ (1 + i),
2
1
λ(3) = √ (1 − i).
2
There is one eigenvector associated with each eigenvalue. Inserting
eigenvalue λ(1) = 1 into the first of equations (15.41) gives (for the
components of C(1) )
√ (1) (1)
(1 − 2)c1 + c2 = 0,
from which
(1) (1)
c2 = 0.41c1 .
where the superscripts are reminders of which eigenvalue is being used.
The second equation of (15.41) gives
(1)
√ (1)
−c1 + (1 − 2)c2 = 0,
538 15. ROTATIONAL KINEMATICS
which yields
(1) 1 (1)
c2 = c .
0.41 1
(1) (1)
The two relations between c1 and c2 are contradictory. They can
(1) (1)
only be satisfied if c1 = 0 and c2 = 0. Now insert eigenvalue λ = 1
into the third equation of (15.41) to obtain
(1)
c3 (1 − λ) = 0,
(1) (1)
which is satisfied for any value of c3 . We usually set c3 = 1 for
convenience. This completes the determination of the first eigenvector,
giving
0
(1)
C = 0 .
1
Next use λ(2) = √12 (1 + i) in the three equations. The first equation
of (15.41) then reads
√
1
1 − 2 √ (1 + i) c1 + c2 = 0
2
from which
(2) (2)
c2 = ic1 .
Plugging λ(2) into the second equation of (15.41) gives the same rela-
tion, so the equations only give the relative values of the components.
(1)
At this point, one usually simply lets c1 = 1. Lastly, plugging λ(2) into
(2)
the third equation gives c3 = 0. So the second eigenvector is
1
(2)
C = i .
0
Similarly, using λ(3) leads to the following expression for the third eigen-
vector
1
C(3) = −i .
0
Once you have obtained the eigenvalues and the related eigenvec-
tors you can write down the modal matrix S by remembering the the
columns of S are the eigenvectors. That is, the columns of S are
15.4. EULER’S THEOREM 539
the angle of rotation φ. The transformed matrix will have the form
cos φ sin φ 0
A0 = − sin φ cos φ 0 .
0 0 1
The trace of A0 is
Tr A0 = 1 + 2 cos φ.
(Recall that the trace is the sum of the diagonal elements.) The trace
of an orthogonal matrix is invariant under a similarity transformation,
so the trace of the original matrix A is also 1 + 2 cos φ. Therefore, you
can determine the rotation angle by simply evaluating the trace of A.
15.6. Summary
The orientation of a rigid body is described in terms of the ori-
entation of a set of axes (x0 , y 0 , z 0 ) fixed in the body relative to a set
of inertial axes (x, y, z). Mathematically, the orientation is represented
by the set of nine direction cosines aij giving the angles between the
primed and the unprimed axes. Since three parameters are sufficient
to describe the orientation of a rigid body, the direction cosines are
not all independent; indeed they are related by the six orthogonality
conditions,
aij aik = δjk .
The rotation (or transformation) matrix A is a 3 × 3 matrix with ele-
ments aij . (Equation 15.8). The position vector in the rotated frame is
x0 and is related to the position vector in the inertial frame by equation
(15.9).
x0 = Ax.
546 15. ROTATIONAL KINEMATICS
15.7. Problems
Problem 15.1. Show that the length of a vector is unchanged by
an orthogonal transformation; that is, prove that
x0i x0i = xi xi .
Problem 15.2. Show that the determinant of the Euler transfor-
mation matrix, Equation (15.32) is +1.
Problem 15.3. Prove that the determinant and the trace of a
matrix are unchanged by a similarity transformation.
Problem 15.4. A “natural” coordinate system for the Earth has
the origin at the center of the planet, the z-axis passing through the
North Pole, and the x-axis in the equatorial plane and going through
the Greenwich meridian. (Where is the y-axis?) Suppose that you
wanted to change to a system with the z-axis going through London
(0◦ E, 51◦ 310 N ), and the x-axis also passing through the Greenwich
meridian. Determine the Euler angles to generate this transformation.
Problem 15.5. People in Fresno felt neglected and asked us to
determine an axis of rotation such that a single rotation would put
Fresno on the top of the world, that is, at the North Pole. Determine
the direction of the axis of rotation as well as the angle required to put
Fresno where it wants to be. At present, Fresno is at 119◦ 470 W and
36◦ 440 N.
Problem 15.6. The Earth is magically rotated through the fol-
lowing Euler Angles: φ = 30◦ , θ = 45◦ , and ψ = 30◦ . Determine
the latitude and longitude of the new directions of the x, y, z axes.
At present, the z-axis points through the North Pole, the x-axis goes
through (0◦ E, 0◦ N ) and the y-axis goes through (90◦ E, 0◦ N )
Problem 15.7. Determine the eigenvalues and eigenvectors for the
matrix
1 3 0
3 −2 −1 .
0 −1 1
Problem 15.8. Prove that the product of two orthogonal matrices
is an orthogonal matrix.
548 15. ROTATIONAL KINEMATICS
COMPUTATIONAL PROJECTS
Computational Project 15.1. Develop a program to determine
the eigenvalues of a 2 × 2 matrix.
Computational Project 15.2. Write a computer program to
find the inverse of a 2 × 2 matrix.
Chapter 16
Rotational Dynamics
from rotation about a fixed axis, to rotation about a fixed point leads
to a much more complicated formulation of the problem. (That is why
we will need to use tensors.) The study of rotational dynamics is of
particular value to physics students because many of the mathematical
techniques will carry over into the analysis of small oscillations about
stable equilibrium points and into quantum theory.
where the subscripts j, k take on the values 1, 2, 3, and δjk is the Kro-
necker delta. Note that Ijk = Ikj so there are only six independent
elements in our expressions for the moments and products of inertia.3
The nine quantities Ijk defined above can be represented as a 3 × 3
array. This array is called the inertia tensor. It is denoted I.
Finally, using matrix multiplication, the angular momentum can be
expressed as
L = Iω. (16.5)
Tensors
A scalar (mass, charge, etc.) is a physical quantity having one
component. If you rotate the coordinate system, the value of a scalar
does not change. A scalar is a tensor of rank zero.
A vector (position, velocity, etc.) is a physical quantity having three
components. If you rotate the coordinate system, the components of
a vector change according to the rule Vi0 = aij Vj where the a’s obey
condition aij aik = δjk . (See Equation 15.6.) A vector is a tensor of rank
one.
A tensor of rank two is a physical quantity having nine components4
that transform according to the rule,
Tij0 = aik ajl Tkl .
Note that the coefficients aik can be considered elements of a matrix A
that transforms T into T 0 by the similarity transformation
T 0 = AT AT .
A rank two tensor can always be represented as a square matrix. Note,
however, that a tensor and a matrix are not the same thing. A matrix
is simply an array of numbers. There are no a priori conditions on the
transformation of a matrix. A matrix usually has no physical signifi-
cance, whereas a tensor is a physical quantity that can be represented
as an orthogonal matrix. A second rank tensor is often represented by
a 3 × 3 matrix. For computational purposes there is no difference be-
tween a tensor and a matrix; if you represent a tensor by a matrix then
the tensor obeys all the rules of matrix algebra. (There are, however,
ways to represent tensors that do not involve matrices.)
A tensor of rank N is a quantity having 3N components which
transform in a particular way under orthogonal transformations. The
general transformation law for a tensor undergoing an orthogonal trans-
formation is:
0
Tijk··· (r0 ) = ail ajm akn · · · Tlmn··· (r),
where the summation convention is implied. Here T 0 (r0 ) is the ten-
sor represented in terms of the primed coordinate system and T (r) is
the tensor represented in terms of the unprimed coordinate system.
In mathematical terms, a tensor is often defined as a quantity that
transforms according to this relation.
You may be thinking that this is very complicated, but you will see
that in practice the only tensors we use are those of rank zero, rank
4I
am assuming three-dimensional space. In a four-dimensional space a rank
two tensor has sixteen elements.
16.1. ANGULAR MOMENTUM 553
one, and rank two (and you are very familiar with tensors of rank zero
and one).
In summary:
A tensor of rank zero (N = 0) has one component and it is invariant
under a coordinate transformation. We call such an object a scalar.
A tensor of rank one (N = 1) has three components and transforms
according to Ti0 = ail Tl . Such an object is called a vector.
A tensor of rank two (N = 2) has nine components and transforms
according to Tij0 = ail ajm Tlm .
We shall not be concerned with tensors of rank higher than two.
Dyads and Dyadics
Another useful quantity is a dyad, defined in terms of two vectors.
It is the object you would obtain if you multiplied two vectors together
in the most naive way possible. For example, the dyad formed from
the vectors A and B and denoted AB is defined as5
AB = Ax ı̂ + Ay ̂ + Az k̂ Bx ı̂ + By ̂ + Bz k̂
= Ax Bx ı̂ı̂ + Ax By ı̂̂ + Ax Bz ı̂k̂
+Ay Bx ̂ı̂ + Ay By ̂̂ + Ay Bz ̂k̂
+Az Bx k̂ı̂ + Az By k̂̂ + Az Bz k̂k̂.
A dyadic is a linear polynomial of dyads, as for example
AB + CD + · · · .
The unit dyadic is
1 =ı̂ı̂+̂̂+k̂k̂.
A dyad operates on a vector by the dot product, thus
AB · C = A(B · C),
and
C · AB = (C · A)B.
In dyadic notation the inertia tensor can be written as
I = mi (ri2 1 − ri ri ),
where the summation is implied, or as
Z
I = ρ(r) r2 1 − rr dτ.
5This kind of product of two vectors is also called the “outer product”, the
“direct product” and the “tensor product.”
554 16. ROTATIONAL DYNAMICS
Exercise 16.5. Show that Equations (16.9) and (16.10) are equiv-
alent.
6As
we shall see, there is a particular set of axes in which the inertia tensor has
only three non-zero components.
16.3. PROPERTIES OF THE INERTIA TENSOR 557
7See,
for example, Florian Scheck, Mechanics: From Newton’s Laws to Deter-
ministic Chaos, Springer Verlag, Berlin, 1990, pp 172-173.
8See, for example, David J. Griffiths, Introduction to Quantum Mechanics,
Prentice Hall, Englewood Cliffs, N.J.,1995, pp 79 - 80.
16.3. PROPERTIES OF THE INERTIA TENSOR 559
axes can be considered principal axes. In that case the inertia tensor
will be diagonal to begin with.
To determine the eigenvalues (or “principal moments of inertia”)
recall that the eigenvalue equation (16.11) can be written as
(I − I1) · R = 0.
A nontrivial solution can be found if and only if the determinant of the
coefficients vanishes, that is, iff
I11 − I I12 I13
I21
I22 − I I23 = 0.
I31 I32 I33 − I
This “secular equation” is a cubic in I and yields the three roots which
are the principle moments of inertia. Each one of these roots can then
be substituted into Equation (16.11) to obtain the directions of the
principal axes.
I 0 = AIAT . (16.17)
a
y
a
x
For a cube, these integrals are quite simple. I will evaluate Ixx just to
illustrate the process. Note that the density is ρ = M/a3 .
Z Z Z
Ixx = ρdτ (y 2 + z 2 )
Vol
Z a Z a Z a Z a Z a Z a
M 2 2
= y dxdydz + z dxdydz
a3 x=0 y=0 z=0 x=0 y=0 z=0
2
= M a2 .
3
It takes no imagination to appreciate that Iyy and Izz are also equal to
(2/3)M a2 . It is left as an exercise to show that the products of inertia
are equal to −(1/4)M a2 .
Therefore, 2
3
− 41 − 14
I = − 41 3
2
− 14 M a2
1 1 2
−4 −4 3
where it is understood that M a2 multiplies every term in the matrix.
Therefore
1
2
− 41 − 14
R2 1 − RR =a2 − 14 2
1
− 14 .
− 14 −4 1 1
2
Finally, the inertia tensor about parallel axes through the center of
mass is
2 1
3
− 14 − 14 2
− 14 − 14
ICM = M a2 − 14 2
3
− 14 − M a2 − 14 1
2
− 14
1 1 2 1 1 1
−4 −4 3
−4 −4 2
1
0 0 2 1 0 0
6 Ma
= M a2 0 16 0 = 0 1 0 .
0 0 1 6 0 0 1
6
You are probably not a bit surprised that the matrix is diagonal. From
the simplicity of the figure (and by LSD 4) you realize that axes parallel
to the cube sides and with origin at the CM will be the principal axes
for the inertia tensor of a cube. Furthermore, you note that the cube
has triply degenerate eigenvalues of M a2 /6.
The principal axes for the cube were found almost accidentally, but
it is not always that simple. To illustrate, let us determine the principal
axes for the object formed when you slice the cube along a diagonal.
As shown in Figure 16.2 the solid is a tetrahedron with three faces that
are right triangles and one face that is an equilateral triangle. We will
first determine the moment of inertia with respect to the (xyz) axes
indicated in the figure, and then determine the principal axes.
The inertia tensor is given by Equation (16.18). I will evaluate a
moment of inertia so you can see how it is done. If the mass of the
tetrahedron is M and its volume is (1/6)a3 , the density is ρ = 6M/a3
and hence,
16.3. PROPERTIES OF THE INERTIA TENSOR 563
z z
a z+x+y=a a
a a
y y
x'
a a
x
x
6M a
Z Z Z Z Z a−x Z a−x−y
2 2
(y 2 + z 2 )dz
Ixx = ρ y + z dτ = 3 dx dy
a x=0 y=0 z=0
Vol
6M a
Z Z a−x
2 1 3
= dx dy y (a − x − y) + (a − x − y)
a3 x=0 y=0 3
Z a 5
6M 1 4 Ma
= (a − x) dx = .
a3 x=0 6 a3 5
(The last two integrals are rather messy and it is easiest to evaluate
them with Maple or some other computer algebra system.)
By symmetry, Iyy = Izz = Ixx = 15 M a2 . It is left as a problem to
show that the products of inertia are all equal to −M a2 /20. Therefore,
the inertia tensor for the tetrahedron about the xyz axes is
1 1 1
− − 2 4 −1 −1
5 20 20 Ma
I = M a2 − 201 1
5
1
− 20 = −1 4 −1 .
− 1
− 1 1 20 −1 −1 4
20 20 5
Having obtained the inertia tensor about the xyz axes of Figure
16.2 let us now determine the principal axes. Using LSD 4, note that
the plane of symmetry indicated in the right-hand panel of 16.2 will
564 16. ROTATIONAL DYNAMICS
So,
5γ − λ = 0,
and
λ2 − 7γλ + 10γ 2 = 0.
Therefore, one eigenvalue is
M a2
(1)
λ = 5γ = 5 ,
20
and the other two are
M a2
λ(2) = 5γ = 5 ,
20
M a2
λ(3) = 2γ = 2 .
20
Note that the eigenvalue 5γ appears twice, indicating a degeneracy.
The first eigenvector is obtained by inserting λ(1) into Equations (16.19),
yielding
(1) (1) (1)
0R1 + 0R2 + 0R3 = 0,
(1) (1)
√ (1)
0R1 − 2γR2 − 2γR3 = 0,
(1)
√ (1) (1)
0R1 − 2γR2 + γR3 = 0.
(1) (1)
The first of these equations is satisfied by any values of R1 , R2 , and
(1)
R3 , the components of the eigenvector corresponding to λ(1) . How-
ever the second and third of these equations cannot be simultaneously
(1) (1)
satisfied by any nonzero values of R2 and R3 . Since it is convenient
for the eigenvectors to have unit magnitude we select the components
of R(1) to be (1, 0, 0). Note that these are the components of the first
eigenvector in the rotated coordinates, x0 , y 0 , z. Once again we see that
x0 is a principal axis.
566 16. ROTATIONAL DYNAMICS
Let us now consider eigenvalue λ(3) = 2γ. Inserting this value into
Equations (16.19) gives
(2)
3γR1 = 0,
(2)
√ (2)
γR2 − 2γR3 = 0,
√ (2) (2)
2γR2 + 2γR3 = 0.
(2)
The first equation requires R1 = 0, so the x0 component of this eigen-
(2) (2) √
vector is zero. The second equation states that R3 = R2 / 2 and
the third equation yields the same relation.√Thus, the second eigenvec-
tor could be written as R(2) = (0, −1, −1/ 2), but this does not have
magnitude unity, so we set it to R(2) = (0, −0.8165, −0.5774). √ [Note
that this gives a vector in the same direction as (0, −1, −1/ 2) but it
has magnitude unity.]
Finally, recall that for a degenerate eigenvalue you can find one
eigenvector by the usual technique, and the other eigenvector will be
orthogonal to the first one. We found the eigenvector R(1) to be (1, 0, 0)
and we know that there is another eigenvector perpendicular to this
one. It will therefore lie in the shaded plane of Figure 16.2. It will
also be perpendicular to R(2) . It is left as an exercise to verify that the
third eigenvector can be written as R(3) = (0, −0.5774, 0.8165).
And so, finally, the principal axes of the body sketched in Figure
16.2 have been determined. It required a significant amount of labor
but hopefully you appreciate the procedure.
(1) (1)
Exercise 16.9. Show that R2 and R3 must be equal to zero.
Exercise 16.10. Show that R(3) = (0, −0.5774, 0.8165) is a vector
of unit magnitude and is perpendicular to both R(1) and R(2) .
z P z' z'
x'
d y'
y y' x'
O
x
Here z 0 is the distance in the z 0 direction to the mass element, but since
we are assuming a zero thickness disk, it is reasonable to set it equal
to zero. Then
Z Z
0 M
Ix0 x0 = 2 y 02 dA.
πa
Changing to polar coordinates and noting that y 0 = r sin φ
Z a Z 2π
0 M M a2
Ix0 x0 = 2 (r sin φ)2 rdrdφ = .
πa r=0 0 4
2 2
Similarly, Iy0 0 y0 = Ix0 0 x0 = M4a , while Iz0 0 z0 = M2a . (You could have ob-
tained these results by elementary means using the perpendicular axis
theorem.) All of the products of inertia are zero. (You will show this
in an exercise.) The inertia tensor with respect to the body axes with
origin at the center of mass of the disk is
1
4
0 0
I 0 = M a2 0 14 0 .
0 0 12
Now translate the origin of the body axes along the z 0 -axis to the
origin of the space axes (x, y, z). The inertia tensor with respect to
568 16. ROTATIONAL DYNAMICS
Exercise 16.11. Show that in the body axes, the idealized gyro-
scope of our example has all the products of inertia equal to zero.
16.4. THE EULER EQUATIONS OF MOTION 569
Replacing dL
dt inertial
by N and dropping the subscripts, this equation
becomes
dL
+ ω × L = N.
dt
Keep in mind that this equation is valid in the body (rotating) reference
frame. Now L =I · ω and in the rotating frame the inertia tensor I is
constant, so
dω
I· + ω× (I · ω) = N. (16.20)
dt
Equation (16.20) is the equation of motion for a rotating body. An
interesting consequence of this equation is that if a body is to ro-
tate freely at constant angular velocity with no torques applied, then
ω× (I · ω) =0, which means that I · ω is parallel to ω. Therefore, ω
must lie along a principal axis. This may seem a bit far-fetched to you,
but just think about the last time you had your car wheels spin bal-
anced. The purpose of the dynamical balancing is to make sure that a
principal axis of the inertia tensor is aligned with the angular velocity
vector. This ensures that the wheel spins without any torques being
exerted on the axis. (Explain that to your mechanic!)
570 16. ROTATIONAL DYNAMICS
Assuming the body axes are principal axes, Equation (16.20) can
be expressed in component form as
I1 ω̇1 − ω2 ω3 (I2 − I3 ) = N1 ,
I2 ω̇2 − ω3 ω1 (I3 − I1 ) = N2 , (16.21)
I3 ω̇3 − ω1 ω2 (I1 − I2 ) = N3 .
These are called Euler’s equations of motion. They are rotational ana-
logues of Newton’s second law of motion. In the next two sections we
describe applications of Euler’s equations.
The third equation establishes that ω3 is constant. The first two equa-
tions can be written as
I1 − I3
ω̇1 = ω3 ω2 ,
I1
I1 − I3
ω̇2 = −ω3 ω1 .
I1
Denoting the constant ω3 (I1 − I3 ) /I1 by Ω, these become
ω̇1 = −Ωω2 , (16.22)
ω̇2 = +Ωω1 . (16.23)
Taking the derivative of the first of these equations and inserting it into
the second equation yields
ω̈1 = −Ω2 ω.
But this is the equation for simple harmonic motion! A solution is
ω1 = A cos Ωt.
Similarly
ω2 = A sin Ωt.
In conclusion, ω3 is constant but ω1 and ω2 vary sinusoidally out of
phase with each other. This means that the total angular velocity
vector, ω = ω1 ı̂+ω2 ̂+ω3 k̂, precesses about ω3 k̂ with a precession rate
Ω. (See Figure 16.4.)
ω3k
ω
y'
x'
y
ϕ
x
ξ
Figure 16.5 shows the orientation of the top in terms of the Euler
angles, θ, φ, ψ. The spinning motion of the top is given by the rotation
9Theasymmetrical top is a more complicated problem and is usually reserved
for graduate courses in mechanics. If you would like to look at the solution see L.D.
Landau and E. M. Lifshitz, Mechanics, 3rd ed, Pergammon Press, Oxford, 1976,
Section 37.
16.6. THE SPINNING TOP 573
about the z 0 -axis, and the spin angular velocity is ψ̇ =ψ̇ẑ0 If the top is
precessing, the rotation vector is tracing out a cone around the vertical
or z-axis. As the rotation axis traces out this trajectory, the angle φ is
changing, so the angular velocity associated with precession is φ̇ =φ̇ẑ.
Finally, the top may be nutating, that is, the axis may be nodding up
and down as it precesses. From the figure, it is clear that this type
of motion represents a change in the angle θ. The angular velocity
associated with nutation would be a vector along the line of nodes
(ξ) and can be expressed as θ̇ξ. ˆ Recall that the Euler angles form a
set of three independent coordinates so they are an appropriate set of
generalized coordinates for describing the motion of the system.
Using the Euler angles of Figure 16.5, the angular velocity ω is
written in terms of its components thus
where the unit vectors are directed along the lines ξ, z, z 0 . Let ê1 , ê2 , and
ê3 be unit vectors directed along the body axes, x0 , y 0 , z 0 . (By symmetry,
these are the principal axes.) In terms of ê1 , ê2 , and ê3 the unit vectors
along ξ, z, z 0 are given by
Consequently, the components of the velocity along the body axes are
Note that ω3 is not simply ψ̇; the precession of the axis of rotation also
has a component in the ê3 direction.
Let the top have mass m and assume that the center of mass lies a
distance d along the axis from the point of contact on the floor. Since
the top is symmetrical, the center of mass lies on the z 0 -axis. The
torque acting on the top is N = dẑ0 × mg, a vector pointing along the
line of nodes (ξ) and having magnitude mgd sin θ. This torque is the
source of the precession.
The top is spinning about its axis, but it is also precessing and
nutating. To develop a mathematical description of precession and
nutation, let us begin by writing the Lagrangian for the system.
574 16. ROTATIONAL DYNAMICS
with
0 (pφ − pψ cos θ)2 p2ψ
V (θ) = + mgd cos θ + . (16.32)
2I1 sin2 θ 2I3
Solving for θ̇ yields
r
dθ 2
θ̇ = = (E − V 0 ). (16.33)
dt I1
Given values for the constants E, pφ and pψ this equation can (in
principle) be integrated to obtain θ = θ(t). Inserting θ(t) in Equations
(16.28) and (16.29) will then yield expressions for φ = φ(t) and ψ =
ψ(t). The problem is thus solved. (It may not be easy to do this, but
conceptually it presents no difficulty.)
However, even without actually solving Equation (16.33) we have
obtained a great deal of information about the behavior of the system.
For example, note that the rotation of the top around its axis of rotation
is ω3 . The third of Equations (16.24) and the expression for pψ given
in Equation (16.29) lead to
pψ = I3 ω3 = constant.
Thus the rotational velocity of the top about its axis of symmetry is a
constant. (This assumes, of course, that the top does not slow down
due to friction at the point of contact with the ground or with the air.)
Figure 16.6 is a plot of the effective potential V 0 (θ) as a function
of θ. Note that the effective potential goes to infinity at 0 and π, as is
clear from Equation (16.32). The minimum in V 0 (θ) can be found by
setting the derivative of V 0 with respect to θ equal to zero. Denoting
the angle at which V 0 is minimized by θ0 , this yields
90
Effe ctive P ote ntia l (J )
80
70
60
50
40
30
20
10
0
0 0.5 1 1.5 2 2.5
The ta (ra d)
energy, the value of θ can vary between two limits (call them θ1 and
θ2 ) and the top nutates.
We shall first consider the case of the minimum energy top with
E = V 0 (θ0 ) and no nutation taking place. The precession rate of the
top is φ̇(θ0 ). To determine the value of this precession we insert θ0 into
Equations (16.28) and (16.29). Thus, from (16.29),
pψ
ψ̇ = − φ̇ cos θ0 ,
I3
so (16.28) reads
2 pψ
pφ = I1 φ̇ sin θ0 + I3 cos θ0 − φ̇ cos θ0 + φ̇ cos θ0 .
I3
Consequently,
pφ − pψ cos θ0 β
φ̇ = 2 = . (16.38)
I1 sin θ0 I1 sin2 θ0
Inserting the value of β from Equation 16.37 we obtain
" s #
pψ 4mgdI1 cos θ0
φ̇ = 1± 1− .
2I1 cos θ0 p2ψ
But pψ = I3 ω3 so
" s #
I3 ω3 4mgdI1 cos θ0
φ̇ = 1± 1− (16.39)
2I1 cos θ0 I32 ω32
Since φ̇ must be real, the quantity under the root must be positive,
that is
4mgdI1 cos θ0
1− ≥0
I32 ω32
from which
I1
ω32 ≥ 2 (4mgd cos θ0 ) .
I3
For precession to take place, ω3 must be at least as great as this mini-
mum value. If the minimum is exceeded, there are two possible values
for φ̇, as described in the following worked example.
Worked Example 16.3. :Show that the fast and slow preces-
sion rates are φ̇ ' (I3 ω3 /I1 cos θ0 ) and φ̇ ' mgd/I3 ω3 . (The slow
precession rate is the one most commonly observed.)
578 16. ROTATIONAL DYNAMICS
Exercise 16.18. Show that the slow precession rate agrees with
the result obtained in Exercise 7.15
Effective Potential V( )
E
0 Angle π
1 2
axis initially at some angle θ1 and released, its axis is observed to trace
out the path shown in 16.8c. The details are left to a problem.
θ1
a b c
Figure 16.8. The path traced out by the rotation axis
during nutation on the surface of a sphere centered on
the point of contact of the spinning top with the floor.
16.7. Summary
The angular momentum of a body of arbitrary shape that is rotating
about a fixed point is
L =I·ω,
10Aspinning coin will slow down and lean over, so that it will be spinning about
a point on its circumference. The axis of rotation does not pass through the coin.
The point of contact between the coin and the surface is not fixed (it moves around
the circumference) but at any instant it can be considered as the point about which
the coin is rotating. A highly polished disk on a smoothly machined flat plat is sold
as a toy called “Euler’s Disk.” A very interesting analysis of the motion is given in
the paper by H. K. Moffatt, “Euler’s Disk and its Finite-time Singularity” Nature,
404, 833, 2000.
16.7. SUMMARY 581
where I is the inertia tensor. The elements of the inertia tensor are
Z
Ijk = ρ(r)(r2 δjk − rj rk )dτ,
so
Z
I= ρ(r)(r2 1 − rr)dτ.
I 0 = AIAT ,
(I − λ1)R =0.
on the total energy. They can be evaluated as two of the roots of the
cubic equation
p2ψ
2I1 (E− )(1−cos2 θ) = p2φ −2pφ pψ cos θ+pψ cos2 θ+2I1 mgd cos θ(1−cos2 θ).
2I3
16.8. Problems
Problem 16.1. Prove that for a symmetrical body, the products
of inertia are zero.
Problem 16.2. For the tetrahedron of Figure 16.2 show that the
products of inertia are −M a2 /20.
Problem 16.3. A system is made up of three particles of mass M
located at (2a, 0, 0), (0, 2a, 0) and (0, 0, a). Find the principal moments
of inertia about the origin. Determine a set of principal axes.
Problem 16.4. Determine the inertia tensor for the ammonia mol-
ecule about its center of mass. Determine the principal axes. The am-
monia molecule (NH3 ) is made up of one nitrogen atom (MW 14) and
three hydrogen atoms (MW1) arranged in a pyramidical shape with the
nitrogen at the apex and the three hydrogens forming an equilateral
triangular base. The distance between the hydrogen atoms and the
nitrogen atom is 1.03 Å and the angles between the bonds are 36.4◦
and 107.2◦ .
Problem 16.5. Let Rn be an antisymmetric matrix whose ele-
ments are given by Rij = ijk xk where the xk are the coordinates of the
kth mass point of a body. Show that the matrix of the inertia tensor
of the body can be written as
I = −mn (Rn )2 .
By definition the Levi-Civita density tensor ijk is zero is any two of
the indices ijk are equal, +1 if ijk = 1, 2, 3 or any even permutation
of 1, 2, 3 and −1 for odd permutations of 1, 2, 3.
Problem 16.6. Prove Labor Saving Device 1, Equation (16.16).
Problem 16.7. Show that Θ = cos−1 (pφ /pψ ) is smaller than θ0 .
Problem 16.8. A uniform right circular cone of height h and base
R has a mass M. It is set on its side and it rolls without slipping in such
a way that the tip of the cone remains fixed. Note that instantaneously
the axis of rotation is along the line of contact between the cone and the
horizontal surface. The angle between this line and a fixed line on the
horizontal plane is θ so the angular velocity of the center of mass of the
584 16. ROTATIONAL DYNAMICS
cone is θ̇. (a) Obtain an expression for the kinetic energy. (b) Obtain
an expression for the angular momentum. (Hint: First determine the
inertia tensor relative to a set of principal axes.)
Problem 16.9. A gyroscope is both rotating about its axis and
precessing about the z-axis and has an angular velocity ω = αẑ + βẑ0 .
Determine the angular momentum about the body axes and the kinetic
energy. Assume I1 = I2 and I3 and θ are known.
Problem 16.10. A gyroscope is hanging from a fixed point (so
θ0 > π/2). Show that there is one positive and one negative value for
φ̇(θ0 ).
Problem 16.11. A spinning top is held with a fixed polar angle
θ = θ1 . Therefore, initially, θ̇ = 0, φ̇ = 0, and ψ̇ = ω3 . It is then
released and allowed to precess and nutate. (a) Write expressions for
pφ , pψ , V 0 , and E. (b) The turning points for the nutation are θ1 and
θ2 , (θ1 ≤ θ ≤ θ2 ). Obtain an expression for cos θ2 . (c) Assume the top
is spinning rapidly so that the quantity α = (2I1 mgd)/(I32 ω32 ) << 1.
Show that for this case, cos θ2 ' cos θ1 − α sin2 θ1 .
Problem 16.12. The spin axis of a rapidly spinning gyroscope
lies in the horizontal plane. That is, θ = π/2. It is precessing about
the vertical axis. (a) Show that elementary considerations lead to a
precession rate of φ̇ = mgd/I3 ω3 . (b) Show that a more sophisticated
analysis based on Equation
D E 16.33 yields the same expression for the
average precession rate φ̇ .
COMPUTATIONAL PROJECTS
Computational Project 16.1. Write a program to solve for θ0
using Equation (16.35). You may assume m = 1, d = 0.1, pφ = 3, pψ =
2, I1 = 3, and I2 = 6.
Chapter 17
Waves
1This is not quite true. A molecule of water also oscillates horizontally and
traces out an elliptical path - but you get the general idea.
2An ideal string is perfectly flexible and linearly elastic. This means the tension
is everywhere the same and always directed tangentially. Linear elasticity means
the tension depends linearly on the amount the string is stretched (Hooke’s law).
587
588 17. WAVES
equal to one-half the length of the string, L. The only waves that will
“fit” onto the string are those with wavelengths λ = (2/n)L where
n = 1, 2, 3, · · · .
λ
y
A
F(x+dx)
ds
y(x+dx,t)
y(x,t)
F(x)
dx x+dx
x
differential equation has then been separated into the following two
ordinary differential equations
c2 d 2 X
= −ω 2 ,
X dx2
1 d2 T
= −ω 2 ,
T dt2
or
d2 X ω 2
+ 2 X = 0,
dx2 c
2
dT
+ ω 2 T = 0.
dt2
Both of these equations have the form of simple harmonic oscillators.
You have seen the SHM equation many times, so I will simply write
down the general solutions
ω ω
X(x) = A cos x + B sin x = A cos kx + B sin kx,
c c
T (t) = C cos ωt + D sin ωt.
The solution for y = y(x, t) is the product of these two functions, thus:
y(x, t) = X(x)T (t) = (A cos kx + B sin kx) (C cos ωt + D sin ωt) .
(17.2)
Note that the partial differential equation (17.1) is satisfied by the four
expressions:
cos kx cos ωt
cos kx sin ωt
y(x, t) ∝ .
sin kx cos ωt
sin kx sin ωt
The general solution (Equation 17.2) is the sum of all of these. The
coefficients, A, B, C, D are determined from the boundary conditions,
as shown in the next section.
∞
X nπx
y(x, 0) = An sin .
n=1
L
But y(x, 0) is given by the function illustrated in Figure 17.3. As
you will show in Problem 17.1, this function can be expressed as
the following Fourier series:
8b πx 1 3πx 1 5πx
y(x, 0) = 2 sin − 2 sin + 2 sin · · · . (17.5)
π L 3 L 5 L
That is,
8b πx 1 3πx 1 5πx
y(x, 0) = sin − 2 sin + 2 sin ···
π2 L 3 L 5 L
∞
X nπx
= An sin ,
n=1
L
where
8b nπ
An = 2 2
sin n = 1, 3, 5, · · · .
nπ 2
Finally, the waveform of the plucked string is
∞
X 8b nπ nπx
y(x, t) = 2 2
sin cos ωn t sin .
n=odd
nπ 2 L
The factor sin(nπ/2) is included to give the sign of the term.
x=0 x=L
y
y(x,0) x
ct
y y(x,t)
x
The fact that the displacement at time t is the same as the dis-
placement at x − ct at an earlier time suggests that the waveform for
a wave traveling at speed c will have the functional form
y(x, t) = f (x − ct).
In the figure, the lower panel represents a later time, so the wave is
moving toward the right, that is, toward positive x. A wave traveling
in the opposite direction would have the functional form
For example, a sinusoidal traveling wave moving toward the right can
be represented as
y(x, t) = A sin k(x − ct).
Here I introduced k simply as a constant to make the argument of the
sine dimensionless; however, if you think about it, you will realize that
k is actually the wavenumber, so the expression becomes
y(x, t) = A sin(kx − ωt).
You might plot this expression for a few nearby values of time and
verify for yourself that this indeed is a representation of a traveling
sine wave moving in the positive x direction.
Although the expression
y = f (x − ct)
seems a reasonable way to describe a traveling wave, it remains to
be shown that it satisfies the wave equation for any function f that
depends on x and t in the combination x − ct. For example, a number
of different possible forms of f would be
y(x, t) = A cos k(x − ct)
y(x, t) = Aeik(x−ct)
y(x, t) = A sin k 2 (x − ct)2 .
The assumption that any function of (x − ct) satisfies the wave equa-
tion can be proved by plugging f (x − ct) into the equation and noting
that
∂ 2y
∂ ∂ ∂ ∂ ∂(x − ct)
= f (x − ct) = f (x − ct)
∂t2 ∂t ∂t ∂t ∂(x − ct) ∂t
∂ 2 f (x − ct)
∂ ∂f (x − ct)
= (−c) = +c2 ,
∂t ∂(x − ct) ∂(x − ct)2
and
2
2∂
y 2 ∂ ∂ 2 ∂ ∂f (x − ct) ∂(x − ct)
c = c f (x − ct) = c
∂x2 ∂x ∂x ∂x ∂(x − ct) ∂x
2
∂ ∂f (x − ct) ∂ f (x − ct)
= c2 = c2 .
∂x ∂(x − ct) ∂(x − ct)2
The right-hand sides of the two expressions are equal, so the wave
equation is satisfied for any function of x − ct.
A particularly useful function of x−ct that is often used to represent
a traveling wave is
y(x, t) = Aeik(x−ct) = Aei(kx−ωt) .
598 17. WAVES
Note that this can be expressed in terms of sines and cosines using the
Euler relation:
y(x, t) = Aei(kx−ωt) = A [cos(kx − ωt) + i sin(kx − ωt)] .
Although this expression satisfies the wave equation, it is not obvious
how to interpret the imaginary part. Therefore, we avoid the problem
by writing
y(x, t) = Re Aei(kx−ωt) ,
that is, only taking the real part of the expression. This formulation is
particularly useful when studying electromagnetic waves.
17.6. Energy
It is quite easy to determine the energy in a wave. As an example,
consider a standing wave in a string of length L.
The kinetic energy of the mass element ρdx as it oscillates up and
down is 2
1 dy
dT = ρdx .
2 dt
To determine the potential energy note that the element of unstretched
length dx has stretched length ds, so the displacement produces a net
stretch in the string of ds − dx. (See Figure 17.2.) The tension in the
string is F, so the work done to stretch the string (which is equal to
the increase in its potential energy) is
dV = F (ds − dx).
But s 2
p dy
ds = dx2 + dy 2 = 1+ dx,
dx
so s
2
dy
dV = F 1 + − 1 dx.
dx
Using the binomial expansion,
" 2 # 2
1 dy . 1 dy
dV = F dx 1 − + ··· − 1 = F dx.
2 dx 2 dx
Consequently, the total energy in a vibrating string of length L is
Z L " 2 2 #
1 ∂y 1 ∂y
E =T +V = ρ + F dx.
0 2 ∂t 2 ∂x
For example, if the wave is described by
πx
y(x, t) = A sin cos ωt,
L
602 17. WAVES
then,
∂y πx
= −ωA sin sin ωt,
∂t L
and
∂y π πx
= A cos cos ωt.
∂x L L
Therefore
Z L
1 2 2 2 πx 2 1 π 2 2 2 πx 2
E = ρω A sin sin ωt + F 2 A cos cos ωt dx,
0 2 L 2 L L
Z L 2 Z L
1 2 2 2 2 πx 1 2 π 2 2 πx
= A ρω sin ωt sin dx + A F 2 cos ωt cos dx ,
2 0 L 2 L 0 L
π2
1 2L 2 2 2
= A ρω sin ωt + F 2 cos ωt .
2 2 L
Now recall that F/ρ = c2 and ω = πc/L so F π 2 /L2 = ρc2 π 2 /(π 2 c2 /ω 2 ) =
ρω 2 and
1 2L 2
E = A ρω (sin2 ωt + cos2 ωt)
2 2
L
= A2 ρω 2 .
4
Note that the energy is proportional to the amplitude squared.
ρ ρ
1 2
x
x=0
yi = Ai sin(k1 x − ωt),
yr = Ar sin(k1 x + ωt), (17.8)
yt = Ai sin(k2 x − ωt).
Note the sign on the ωt term in the expression for yr and the subscript
on k for yt .
The speed of the wave is different in the two mediums. Since c =
ω/k, it is clear that if the speed changes there will be a change in either
ω or k or both. The three waves presented above all have the same
value of ω and different values of k. You may wonder why k changes
but ω stays the same. The reason is that the frequency of the wave
is determined by the physical mechanism that is generating the wave.
If you are standing at one end of the string and shaking it up and
down, you are controlling the frequency of the wave by how quickly (or
slowly) you move the end of the rope. This you can control. However,
you have no control over the wavelength. The wavelength depends on
the properties of the medium. When the properties of the medium
change the wavelength changes, and so the wavenumber k = 2π/λ also
changes.
The energy of a wave is proportional to the square of its amplitude.
An interesting and important question is how much energy is reflected
at an interface and how much is transmitted into the second medium.
The question can be answered by determining the amplitudes of the
reflected and transmitted waves relative to the incident amplitude. In
other words, it is desirable to determine the ratios Ar /Ai and At /Ai .
This can be done by realizing that both y and dy/dx must be continuous
at the junction between the two strings. The value of y is continuous
because the strings are attached to one another. The reason why dy/dx
is the same on both sides of the junction is that the tension is the same
604 17. WAVES
17.7. Momentum
The material particles in a transverse wave move perpendicular to
the motion of the wave. Thus, you might think that a transverse wave
could not transport momentum longitudinally. However, it is a known
fact that waves in strings do transport momentum. This means that
our assumption of purely transverse motion for a particle in a string
must be wrong. There must be some longitudinal motion as well as
transverse motion.3 This can be easily appreciated conceptually by
(mentally) replacing the string with a series of particles of mass m
connected by massless springs of constant k, as shown in Figure 17.6.
As indicated in the figure, if one particle is plucked so that it is displaced
3Electromagnetic waves in a vacuum are transverse. But the momentum density
associated with these waves is 0 E ×B which is perpendicular to the transverse
oscillations of E and B. For a string, however, perfectly transverse waves would not
have a longitudinal component of momentum.
606 17. WAVES
and
∂y ∂y ∂f ∂y
= = (−cT )
∂t ∂f ∂t ∂f
So
∂y 1 ∂y
=− .
∂x cT ∂t
Therefore,
Z L0 2 Z L0 Z L0
1 dy 1 ∂y 1 ∂y 1 ∂y ∂y
∆x = dx = − dx = − dt.
0 2 dx 0 2 ∂x cT ∂t 0 2 ∂x ∂t
But Z L0 Z L0
∆x = ẋdt = vx dt.
0 0
That is, the horizontal velocity of these waves is given by
1 ∂y ∂y
vx = − .
2 ∂x ∂t
Consider the longitudinal momentum of an element of the string of
length dx. The mass of this element is ρdx. The momentum in the x
direction of this element is
1 ∂y ∂y
(ρdx)(vx ) = ρdx − .
2 ∂t ∂x
Thus the momentum per unit length (or momentum density) is
1 ∂y ∂y
g = ρvx = ρ − .
2 ∂t ∂x
17.8. Summary
In this chapter you were exposed to an analysis of waves in strings.
The relations obtained, however, are applicable (with slight modifica-
tions) to other kinds of waves.
Considering the motion in an element of a string and applying New-
ton’s second law, it is easy to derive the relation
∂ 2y F ∂ 2y
= .
∂t2 ρ ∂x2
For a traveling wave, the speed of propagation (phase speed) is c =
p
F/ρ and the relation above becomes
∂ 2y 2
2∂ y
= c ,
∂t2 ∂x2
which is called the wave equation.
608 17. WAVES
Ar c1 − c2
= ,
Ai c1 + c2
At 2c2
= ,
Ai c1 + c2
where c1 and c2 are the speeds of the waves in medium 1 and medium
2. The average energy transmitted per unit time is:
F ω2 2
dW
= A.
dt t 2c2 t
17.9. Problems
Problem 17.1. Obtain expression 17.5 for the shape of the string
in Figure 17.3.
17.9. PROBLEMS 609
Problem 17.10. A stretched string has one fixed end, but the
other end (at x = L) is tied to a massless ring that slides on a fric-
tionless vertical rod. Determine the boundary conditions and obtain
an expression for a wave in this string.
Problem 17.11. One end of a string of length L is connected to
a mechanism that makes it oscillate according to y(0, t) = A sin ωt.
The other end (x = L) is fixed. Determine the motion. (Answer:
y = A sin ωt[cos kz − cot kL sin kz].)
Problem 17.12. In Worked Example 17.1 we solved the problem of
a plucked string and obtained a standing wave. Now solve the problem
by assuming the resultant standing wave is the sum of two traveling
waves f (x−ct) and g(x+ct). Make sure initial and boundary conditions
are satisfied.
Problem 17.13. A string of length L is fixed at both ends. At
time t = 0 the displacement at any point is given by u(x) and the
vertical speed is given by v(x). Recall that the general solution can be
written as
∞
X nπx
y(x, t) = (An sin ωn t + Bn cos ωn t) sin .
n=1
L
Show that Z L
2 nπx
An = v(x) sin dx,
ωn L 0 L
and Z L
2 nπx
Bn = u(x) sin dx.
L 0 L
Problem 17.14. A normal mode consists of an oscillation at a
single frequency. The general motion of a vibrating string is
∞
X nπx
y(x, t) = (An sin ωn t + Bn cos ωn t) sin .
n=1
L
This can be expressed as a sum of normal modes by defining the normal
coordinate φn as
φn = An sin ωn t + Bn cos ωn t,
leading to
∞
X nπx
y(x, t) = φn sin .
n=1
L
17.9. PROBLEMS 611
(a) Express the total energy in terms of the normal coordinates φn . (b)
Express the Lagrangian in terms of the normal coordinates and obtain
the equations of motion in terms of φn .
COMPUTATIONAL PROJECTS
Computational Project 17.1. Using the expression for y(x, 0)
obtained in Worked Example 17.1, sum the first five terms and plot.
Then sum the first ten terms and plot.
Computational Project 17.2. Consider the two traveling waves
y1 (x, t) = A sin(kx − ωt) and y2 (x, t) = A sin(kx + ωt). Show that the
sum of these waves is a standing wave. To do so, generate a three panel
(“movie”) plot showing the two traveling waves and the standing wave
as functions of time.
Computational Project 17.3. Using the model of Figure 17.6
as a basis and displacing the leftmost particle vertically, show the de-
velopment of a traveling wave. Compare your results with Figure 5 of
Rowland and Pask (Am. J. Phys.,67,1999, page 383).
Chapter 18
Small Oscillations (Optional)
18.1. Introduction
An important topic in mechanics is the study of small oscillations
of coupled systems of particles about their equilibrium positions. Ide-
alized examples of such systems are the particles connected by massless
springs in Figure 18.1 and the coupled pendulums shown in Figure 18.2.
These systems were treated in a more elementary manner in Chapter
11; you may wish to refer to that chapter as we go along.1
k k k k k
m m m m
xi = xi (q1 , · · · , qn ). (18.2)
Then the Lagrangian for the system is a function only of the q’s and
the q̇’s, and does not depend on time:
d ∂T ∂T ∂V
− =− . (18.4)
dt ∂ q̇i ∂qi ∂qi
1 2
k
m m
n n
d X ∂xi dqj X ∂xi
xi = ẋi = = q̇j .
dt j=1
∂qj dt j=1
∂qj
18.2. STATEMENT OF THE PROBLEM 615
have been a bit abstract, but the result itself merely states an obvious
fact.
From the figure it is obvious that only for case (a) in which the
potential is a minimum at the equilibrium point is the system in stable
equilibrium. Stable equilibrium implies that if the system is displaced
from the equilibrium point by a small amount it will tend to return
to the equilibrium point. For cases (b) and (c) a similar displacement
(even by an infinitesimal amount) does not result in the system return-
ing to equilibrium.
A small displacement from equilibrium can be represented by
qi = qi0 + ηi , i = 1, · · · , n, (18.7)
(18.9)
∂V
But ∂qj
= 0 and V (q10 , · · · , qn0 ) is simply a constant that can be set
q0
equal to zero by redefining the zero point of potential energy. Then,
18.2. STATEMENT OF THE PROBLEM 617
Note that mkj and vkj are symmetric so the order of the subscripts
makes no difference. Equations (18.13) are a set of n linear, homoge-
neous, coupled, second order, differential equations with real, constant
coefficients. Each equation has 2n terms involving the elements of the
n × n matrices we have named the mass matrix and the potential ma-
trix. If you can find the solutions to Equations (18.13), that is, if
you can find functions ηi (t) that satisfy (18.13), you have solved the
dynamical problem.
Note that Equation (18.13) can be written in matrix form as
Mη̈ + Vη = 0,
where M and V are the n × n mass and potential matrices and η is an
n-component vector giving the displacements of all the particles from
their equilibrium positions.
θ l
y m
= ml2 .
Similarly, since V = −mgy = −mgl cos θ, Equation (18.11) leads to
2
∂ V ∂ ∂
v11 = =− (mgl cos θ) ,
∂qj ∂qk q0 ∂θ ∂θ θ=0
v11 = mgl cos θ |θ=0 = mgl.
For this problem, the mass matrix and the potential matrix are 1 × 1
matrices having a single element. In general, if the system is charac-
terized by a single coordinate, you will find that mjk = m and vjk = κ
where m and κ are real numbers (but not necessarily equal to the mass
and the force constant) and the equation of motion (18.13) reduces to
mη̈ + κη = 0, (18.14)
or
κ
η̈ = − η.
m
18.3. NORMAL MODES 621
or
n
X n
X
(s)∗ (s) (s)∗ (s)
zk vkj zj = ωs2 zk mkj zj .
k,j k,j
Consequently
Pn (s)∗ (s)
k,j zk vkj zj
ωs2 = Pn (s)∗ (s)
. (18.25)
z
k,j k mkj z j
If you take the complex conjugate of this expression and use the fact
that the vkj and mkj are real, you obtain
Pn (s) (s)∗
2 ∗
k,j zk vkj zj
ωs = Pn (s) (s)∗
.
z
k,j k mkj z j
The elements of the mass matrix (mkj ) are positive (as is evident from
the definition, Equation 18.6) so the denominator in (18.27) is positive.
The numerator will be positive or negative depending on the value of
vkj (which is defined in Equation 18.11):
2
∂ V
vkj = vjk = .
∂qj ∂qk q0
If the potential energy V is a minimum at equilibrium point q 0 then its
second derivative is positive, so vkj > 0 and condition (18.27) is met.
Therefore, stable oscillations can only occur around points where the
potential energy is a minimum - a result that is intuitively obvious.
Returning to the problem of determining the motion of the system,
let us pick a particular eigenvalue (ωs2 ) and consider the form of the
solutions of Equations (18.23) which can be written in a more general
way as
Xn
(s)
(vkj − ωs2 mkj )zj = 0. (18.28)
j=1
18.3. NORMAL MODES 625
The value of φs is the same for all the zk ’s. But if eiφs is a common
factor, you can divide all of the equations by it and write Equations
(18.28) in the form
n
X n
X
(s) (s)
vkj ρj = ωs2 mkj ρj , k = 1, 2, · · · , n. (18.32)
j=1 j=1
(s)
Equation (18.32) is the eigenvector equation. Note that the ρj are a
set of n real quantities. They can be ordered into a column matrix,
626 18. SMALL OSCILLATIONS (OPTIONAL)
thus,
(s)
ρ1
(s)
(s)
ρ2
ρ = . (18.33)
..
.
(s)
ρn
Here ρ(s) is an eigenvector.
In like manner, the eigenvector equation corresponding to a different
frequency ωt2 is
Xn n
X
(t) 2 (t)
vkj ρk = ωt mkj ρk . (18.34)
k=1 k=1
P (t)
Now operate on Equation (18.32) with k ρk and on (18.34) with
P (s)
j ρj to obtain
X (t) (s)
X (t) (s)
ρk vkj ρj = ωs2 ρk mkj ρj , (18.35)
k,j k,j
and X X
(s) (t) (s) (t)
ρj vkj ρk = ωt2 ρj mkj ρk . (18.36)
k,j k,j
There are n solutions of this form for each of the eigenvalues ωs . Each
(s)
such solution has (in general) a different value of ρj but the factors
C (s) and φs are common to all the solutions corresponding to the same
eigenvalue. For example, for the system illustrated in Figure 18.1, the
displacements of particles 1, 2, 3, · · · , j, · · · , n correspond to the real
(s) (s) (s) (s)
parts of z1 , z2 , · · · , zj , · · · zn where all the particles are assumed to
oscillate at the same (normal) frequency ωs .
The solution of the coupled problem is obtained as a linear super-
position of all the particular solutions, thus,
Xn h i
(s) (s)
zj (t) = z+j eiωs t + z−j e−iωs t , j = 1, · · · , n, (18.41)
s=1
where + −
(s) (s) (s) (s) (s) (s)
z+j = C+ ρj eiφs , z−j = C− ρj eiφs . (18.42)
The general solution is the sum over all the normal modes. It is
clear that if the normal mode solutions satisfy the differential equation,
then a sum of normal mode solutions will also satisfy the differential
equation. This is just the principle of superposition.3
The actual physical motion, ηj , is obtained by taking the real part
of equation (18.41), yielding
n
X (s)
ηj (t) = C (s) ρj cos(ωs t + φs ). (18.43)
s=1
V − ωs2 M ρ(s) = 0.
(18.44)
The orthonormality condition (18.39) is
T
ρ(s) Mρ(t) = δst ,
and the general solution (18.43) can be expressed as a linear combina-
tion of the normal mode solutions,
X
η(t) = ρ(s) C (s) cos(ωs t + φs ). (18.45)
s
The constants C (s) and φs can be associated with the initial conditions
η(0) and η̇(0) by (18.45) and its derivative:
X
η(0) = ρ(s) C (s) cos φs ,
s
X
η̇(0) = − ρ(s) C (s) ωs sin φs .
s
Usually, you are given η(0) and η̇(0) and asked to find C (s) and φs .
T
These quantities are easily determined by multiplying by ρ(t) M
from the left and using the orthogonality conditions.
T X T
ρ(t) Mη(0) = ρ(t) Mρ(s) C (s) cos φs = C (t) cos φt ,
s
and
T X T
ρ(t) Mη̇(0) = − ρ(t) Mρ(s) ωs C (s) sin φs = −ωt C (t) sin φt .
s
or
2L = η̇ T Mη̇ − η T Vη. (18.52)
But η(t) = Aζ(t) and η̇(t) = Aζ̇(t), so
T
2L = Aζ̇ M Aζ̇ − (Aζ)T V (Aζ) ,
= ζ̇ T AT MA ζ̇ − ζ T AT VA ζ,
= ζ̇ T ζ̇ − ζ T ωD
2
ζ. (18.53)
Or, in component form,
n
1 X 2
L= ζ̇i − ωi ζi2 . (18.54)
2 i=1
Note that using the modal matrix to define the new generalized coor-
dinates has, in a sense, diagonalized the Lagrangian. Note also that
the ωi ’s in Equation (18.54) are the set of normal frequencies for the
system.
In terms of the normal coordinates, ζi , the Lagrange equations of
motion are
n
! n
!
d ∂ 1 X 2 ∂ 1 X
ζ̇ − ωi ζi2 − ζ̇ 2 − ωi ζi2 = 0, (18.55)
dt ∂ ζ̇i 2 i=1 i ∂ζi 2 i=1 i
or
ζ̈i = −ωi2 ζi , (18.56)
which is the equation for simple harmonic motion. Therefore diagonal-
izing the Lagrangian decoupled the equations of motion and reduced
the problem to a set of n independent, decoupled, simple harmonic os-
cillator equations. Each normal coordinate ζi oscillates independently
with angular frequency ωi .
The derivation has been perfectly general and implies that any me-
chanical system undergoing small amplitude oscillations around points
of stable equilibrium can be described in terms of normal coordinates.
The solution to Equation (18.56) can be expressed in component
form as
ζi = C (i) cos(ωi t + φi ), i = 1, · · · , n (18.57)
632 18. SMALL OSCILLATIONS (OPTIONAL)
Comparing Equations (18.59) and (18.57) you can see that the normal
coordinates ζi are the coefficients of the eigenvectors ρi in the expansion
of ηi .
. 1
= mgl sin2 θ1 + sin2 θ2 .
2
Therefore, to second order in small quantities,
mg 2
VG = (η + η22 ). (18.61)
2l 1
The potential energy of the spring is 21 k(d−d0 )2 where d is the distance
between the masses and d0 is the unstretched length of the spring.
18.6. COUPLED PENDULUMS: AN EXAMPLE 633
These quantities are illustrated in Figure 18.5, from which you can see
that
1
d = y 2 + (d0 − η1 + η2 )2 2 ,
! 21
2(η1 + η2 ) (η2 + η1 )2 y 2
= d0 1 + + + 2 + ··· ,
d0 d20 d0
.
= d0 + η2 − η1 .
To second order in small quantities,
1 1
VS = k(d − d0 )2 = k(η2 − η1 )2 . (18.62)
2 2
Having obtained the kinetic and potential energies, the Lagrangian
is
1 mg 2 1
L = T − V = m η̇12 + η̇22 − (η1 + η22 ) − k(η2 − η1 )2 . (18.63)
2 2l 2
The equations of motion are
mg
mη̈1 + k + η1 − kη2 = 0, (18.64)
l
mg
mη̈2 + k + η2 − kη1 = 0.
l
k m
m
η1
time
η2
time
τ
τ
y
φ
θ
a
So,
sin (k(N + 1)a) = 0,
and
k(N + 1)a = nπ, n = 1, 2, · · · , N. (18.87)
18.7. MANY DEGREES OF FREEDOM 639
= An e−iωn t 2i sin(kxj ),
nπxj
= 2iAn sin [cos ωn t − i sin ωn t] .
a(N + 1)
Since y(xj , t) is real, we only keep the real part of this solution. As-
suming An is real,
nπxj
y(xj , t) = 2An sin sin ωn t, (18.89)
a(N + 1)
where ωn2 is given by Equation (18.88).
Exercise 18.7. Show that ω 2 = (4τ /ma) sin2 (ka/2) follows from
Equations (18.82) and (18.84).
where c2 = τ /σ. Note that I am using σ for the mass density of the
string because in this chapter ρ is being used to represent the eigen-
vectors.
Worked Example 18.3. Solve the wave equation (18.95) us-
ing the techniques developed in this chapter. (Recall that in Chap-
ter 17 the wave equation was solved by separation of variables.)
Solution: Assume normal mode solutions of the form
y(x, t) = Cρ(x) cos(ωt + φ). (18.96)
Plugging this expression into the differential equation (18.95) yields
d2 ρ(x)
+ k 2 ρ(x) = 0, (18.97)
dx2
where k = ω/c. Equation (18.97) is an ordinary differential equa-
tion that is also an eigenvalue equation. Solutions exist only for
specified allowed values of k 2 . (The solutions are the eigenfunc-
tions.) Since Equation (18.97) has the form of the SHM equation,
we can write a solution immediately:
r
2
ρ(x) = sin kx, (18.98)
lσ
p
where the coefficient 2/lσ was selected for normalization.
For a string tied at both ends, the boundary conditions are
ρ(0) = 0 and ρ(l) = 0. The solution (18.98) meets the first bound-
ary condition, but the second condition is met iff sin kl = 0, i.e., iff
k is one of the set
nπ
kn = , n = 1, 2, · · · , ∞. (18.99)
l
Since k = ω/c, the normal mode frequencies are
nπc
ωn = kn c = .
l
The allowed wavelengths are
2π 2l
λ= = , n = 1, 2, · · · , ∞. (18.100)
k n
Thus, the string must contain an integer number of half wave-
lengths. The general solution is obtained as a superposition of
normal modes and is given by
X∞
y(x, t) = Cn ρ(n) (x) cos(ωn t + φ), (18.101)
n=1
18.8. TRANSITION TO CONTINUOUS SYSTEMS 643
or, equivalently,
∞ r
X 2
y(x, t) = sin kn x (an cos ωn t + bn sin ωn t) . (18.102)
n=1
lσ
Here an = Cn cos φn and bn = −Cn sin φn . Each term in this sum
satisfies the wave equation and the boundary conditions. The val-
ues of the constants can be determined from the initial conditions:
∞ r
X 2
y(x, 0) ≡ an sin kn x,
n=1
lσ
and ∞ r
X 2
ẏ(x, 0) ≡ ω n bn sin kn x.
n=1
lσ
The expressions in parentheses are Fourier series so you can use
the orthonormality property of these series to write
r Z l
2σ
an = y(x, 0) sin (kn x) dx, (18.103)
l 0
and r Z l
2σ
ωn bn = ẏ(x, 0) sin (kn x) dx. (18.104)
l 0
Exercise 18.8. Using condition 18.100, plot the first four waves in
a string with fixed endpoints.
We can write
∂y ∂
δ = δy,
∂x ∂x
and
∂u ∂
δ = δu.
∂t ∂t
Carrying out integrations by parts as usual, leads to
Z t2 (Z l " # )
∂L ∂ ∂L ∂ ∂L
0= dt dx − ∂y
− δy .
t1 0 ∂y ∂x ∂ ∂x ∂t ∂ ∂y∂t
18.9. Summary
Interacting particles often exhibit small oscillations about equilib-
rium. Two classical examples of such systems are illustrated in Figures
18.1 and 18.2. The problem could be set up in terms of Cartesian
coordinates (xi ), but it is much more convenient to introduce a set
646 18. SMALL OSCILLATIONS (OPTIONAL)
The solution is obtained by making the usual assumption that all the
particles oscillate at the same frequency ω (normal mode assumption)
so that
zj = zk0 eiωt .
Plugging this into the equations of motion leads to a set of n coupled
algebraic equations that have a nontrivial solution iff
det vkj − ω 2 mkj = 0.
18.9. SUMMARY 647
The normal mode frequencies ωs can then be determined. Once the fre-
quencies are determined, the complex coordinates zk can be evaluated
using Cramer’s rule. The coefficients in the Cramer’s rule procedure
are
akj = vkj − ωs2 mkj ,
(s)
and the solution gives n values of zk for each frequency ωs . Here the
superscript (s) indicates the particular normal mode. (Actually, we
(s) (s)
obtain n − 1 ratios of the form zk /zn .)
(s)
The complex coordinates zk can be expressed in polar form, thus,
(s) (s)
zk = ρk eiφs .
(s)
There is a different ρk for each value of k, but the φs are all the same
(for a given s). The set of ρ values for a given normal mode (s) is
conveniently expressed as a column vector, thus,
(s)
ρ1
ρ(s)
ρ(s) = . 2 .
..
(s)
ρn
This is called the eigenvector for the normal mode. The various eigen-
vectors are orthogonal. Once the eigenvectors have been determined,
the physical coordinates can be obtained from
n
X (s)
ηj (t) = C (s) ρj cos(ωs t + φs ),
s=1
18.10. Problems
Problem 18.1. Determine the mass matrix and the potential ma-
trix for a system of two equal masses and three equal springs arranged
as in Figure 18.1.
Problem 18.2. A particle of mass m finds itself in a region of space
where the potential is given by
V0
V (x, y, z) = 2 4x2 + 5y 2 + 6z 2 − 2ax − 5ay .
a
V0 and a are positive constants. (a) Determine the location and value
of V at its minimum. (b) Evaluate the mass matrix and the potential
matrix. (c) Determine the normal mode oscillation frequencies about
the equilibrium point.
Problem 18.3. A massless spring of constant k is fixed to the
ceiling and is hanging vertically. A particle of mass m is attached to
the bottom of the spring. Then another identical spring is attached
to the particle and finally a second particle (also having mass m) is
attached to the second spring. The system only moves in the vertical
18.10. PROBLEMS 649
of Equation (18.12).
Problem 18.5. Three equal masses m are connected by four equal
springs of force constant k and length a, as in the system of Figure
18.1. Somehow the masses are constrained to oscillate transversely,
that is, perpendicular to the line of the springs. (Perhaps the masses
are mounted on thin rods.) The oscillations all take place in a single
plane. Find the normal mode frequencies.
Problem 18.6. Three equal springs are placed in a frictionless
circular trough of radius a. At the junctions of the springs are masses
m, m, and 2m. Determine the normal coordinates and frequencies of
oscillation.
Problem 18.7. Assume the system of Figure 18.1 consists of two
identical masses and three identical springs. (a) Find the Lagrangian
and Lagrange’s equations. (b) Determine the normal mode frequencies
and the eigenvectors. (c) Construct the modalmatrix. Answers: (b)
1 1 1
ω12 = k/m, ω32 = 3k/m. (c) A = √2m ..
1 −1
Problem 18.8. Consider the system of Figure 18.1 as described in
the previous problem Let one mass be displaced from equilibrium by a
small distance a in the direction of the other mass. Obtain the motion.
Problem 18.9. A system like the one in Figure 18.1 consists of
two equal masses and three equal springs. The system is initially at
rest. Then the mass on the right is subjected to an additional force
F = A sin ωt. Determine the motion.
Problem 18.10. Consider a linearly symmetrical triatomic mole-
cule. In equilibrium two atoms of mass m are located on either side
of an atom of mass M. All three atoms lie along a straight line and
in equilibrium the distance between them is b. The interatomic forces
can be assumed to obey Hooke’s law. Assume vibrations take place
along the line of the molecule. Determine the normal mode frequen-
cies. Describe the normal modes. (Note, the zero frequency normal
mode corresponds to a translation of the entire molecule.)
650 18. SMALL OSCILLATIONS (OPTIONAL)
Special Topics
Chapter 19
The Special Theory of Relativity
This chapter presents some of the basic ideas of the special theory of
relativity. Relativity theory involves serious modifications to classical
mechanics as well as to many of our basic notions about time, space
and causality. For this reason, relativity theory is of as much interest
to philosophers as it is to physicists.
large
ClassicalMechanics RelativityTheory
Size
Relativistic
Quantum
Quantum Mechanics
small Mechanics
slow fast
Speed
Figure 19.1. Different realms of physics. Classical me-
chanics applies to objects that are “large” and “slow.”
When the Nazi party came into power in Germany and began to
persecute Jews, Einstein left Berlin and emigrated to the United States
where he took a position at the Princeton Institute for Advanced Stud-
ies. He was convinced that German scientists would develop nuclear
weapons and he signed a letter to President Roosevelt urging him to
begin a project to develop the atomic bomb even though this went
against his pacifist philosophy. Eventually he took out American citi-
zenship. His last years were a model of a tranquil academic life. After
he died a medical examiner removed his brain. The story of Einstein’s
brain and its subsequent travels is a fascinating but gruesome tale.2
y y*
v
R R*
O x O* x*
returns to the first mirror, and the clock ticks. (Nobody said this was
a practical clock, but it certainly is a possible one.) The left panel of
Figure 19.3 illustrates the clock. Assume Mary has the clock in her
rocket ship (reference frame R∗ ). In this reference frame the clock is at
rest. The mirrors are separated by a distance d/2, so the time required
for the light pulse to go from A to B and back to A is d/c where c
is the speed of light. The time between ticks (according to Mary) is
∆tM = d/c.
B vt
B
d/2
A A A
l
rest in his reference frame. He says her clock is running slow. This
effect is known as “time dilation.”
Remember that ∆tM is the time between ticks on a clock at rest
with respect to Mary and ∆tJ is the time between ticks on that same
clock as determined by John. Mary’s heart is a kind of clock. She says
her heart is beating at a normal 60 beats per minute. But John says
that Mary’s heart is beating more slowly, say, at 40 beats per minute.
He decides she is aging more slowly than he is.
The time measured on the clock at rest in R∗ is usually denoted t∗ .
Time intervals measured by the clocks in the two reference frames are
related by
p
t∗2 − t∗1 = (t2 − t1 ) 1 − v 2 /c2 . (19.3)
Mary is at rest with respect to the clock in the rocket ship. If she
observes the clock in John’s reference frame, she will say it is his clock
that is running slow. (After all, this is the theory of relativity!)
Having determined that moving clocks run slow (as observed from
a stationary frame), let us consider the length of a moving object as
observed from a stationary frame. Imagine Mary has a table in her
rocket ship and the table is aligned along the common x-axis of the
two systems. To determine the length of the table, she sets up a mirror
at one end and a lamp at the other. She (somehow) measures the time
for a light signal to go from the lamp to the mirror and back to the
lamp. This time is measured on her clock and is denoted ∆tM (or ∆t∗ ).
She concludes that the length of the table is l∗ = (1/2)c∆t∗ . Note that
the time of flight is ∆t∗ = ∆tM = 2l∗ /c.
John observes this experiment. He notes that the light ray that
left the lamp had to travel a distance l + v∆t1 to get to the mirror,
because the rocket moved the mirror a distance v∆t1 while the light
was traveling. The time for the light signal to get to the mirror is
1
∆t1 = (l + v∆t1 ),
c
and solving for ∆t1 we find
l
∆t1 = .
c−v
The light signal now must return to the position of the lamp, but this
time (according to John) it only has to travel a distance l − v∆t2 where
∆t2 is the time of flight. Once again
1
∆t2 = (l − v∆t2 ),
c
19.4. THE LORENTZ TRANSFORMATIONS 661
y y*
vt (in R)
-vt* (in R*)
v
t
t*
E
x x*
O*
O
Our job is half done, as we have determined the relation between x and
x∗ . Now we need to obtain the relation between t and t∗ . This is quite
easily done by noting that Mary will obtain the same sort of transfor-
mation equation between the starred and unstarred x coordinates. She
finds that
x∗ + vt∗
x= p . (19.6)
1 − v 2 /c2
Plugging Equation (19.5) into (19.6) yields
" #
1 x − vt
x= p p + vt∗ .
2
1 − v /c2 2
1 − v /c 2
two flashes (or events) and calls it τ0 (= t∗2 − t∗1 ). John also sees the
two events. According to his clock the time between the events is
τ = t2 − t1 . What is the relation between the time measured by
Mary and the time measured by John?
Solution: Note that relative to Mary the two events occur at
the same place, so x∗2 = x∗1 . The time interval is calculated using
the inverse relation, the last equation of (19.10). The result is
t∗ − t∗1
t2 − t1 = p 2 .
1 − v 2 /c2
Consequently,
τ0
τ=p .
1 − v 2 /c2
Thus τ > τ0 . The time interval in the rest frame (τ0 ) is shorter
than the time interval in the reference frame that is moving with
respect to the events. Although John is at rest on the asteroid, in
this example, his is the moving clock and moving clocks run slow.
It is interesting to consider the time read on a distant clock by
a moving observer. Suppose that John has set out three synchro-
nized clocks at positions x = −l, 0, +l. Mary is passing by in her
rocket ship and at the instant she passes the origin of John’s refer-
ence frame, she checks the times on the clocks. (Actually, it might
take her a long time to make the determination, but she can al-
ways figure out later what the clocks read at the instant she passed
John’s origin.) In Problem 19.3 you are asked to show that accord-
ing to Mary, the three clocks will read times of +lv/c2 , 0, −lv/c2 .
Thus, the clock she is approaching reads an earlier time than the
clock at rest by −lv/c2 .
C
B
A
and
q
u∗y 0.70c sin 60◦ 1 − (0.75c)2 /c2
uy = =
γ(1 + u∗x v/c2 ) (1 + (0.70c cos 60· ) (0.75c) /c2 )
0.10
= = 0.08c
1.263
L L
v
Mary
John W1 W2
Event = E1 Event = E2
x1 = γ(x∗1 + vt∗1 ) = γ(−L + vt∗1 ) x2 = γ(x∗2 + vt∗2 ) = γ(L + vt∗2 )
t1 = γ(t∗1 + v ∗ L
+ cv2 (−L) t2 = γ(t∗2 + cv2 x∗2 ) = γ Lc + cv2 L
x)
c2 1
=γ c
events was
x2 − x1 = γ(L + vt∗2 ) − γ(−L + vt∗1 )
= γ (2L + v(t∗2 − t∗1 ))
= 2γL,
where the last step uses the fact that t∗2 = t∗1 . Therefore, if a signal were
to propagate from E1 to E2 in John’s reference frame it would have a
speed given by
distance 2γL c
speed = = = c > c.
time 2γLv/c2 v
Consequently, E1 cannot be the cause of E2 in any coordinate system.
A nice geometric way to understand the causality problem is based
on the concept that an event is a point in a four-dimensional space-time
coordinate system. An event is specified by three spatial coordinates
(x, y, z) and one temporal coordinate (t). I cannot draw (or even vi-
sualize) a four-dimensional space, but if, for example, z = 0 at all
times, then an event can be represented graphically as a point in the
three-dimensional projection of the four-dimensional space-time dia-
gram. See Figure 19.7.
t E(x,y,z,t)
y
x
previous one and slightly higher along the time axis. These circles gen-
erate a cone in the three-dimensional slice of space time, as illustrated
in Figure 19.8. Events that are causally related must lie within the
same light cone.
Also
p
distance l l0 1 − v 2 /c2
t = 30 = = =
velocity v v
p
50v 1 − v 2 /c2
30 = .
v
Consequently,
3 p
= 1 − v 2 /c2
5
and
4
v = c.
5
ct
3
2
1
x
-4 -3 -2 -1 1 2
world line of
origin of B ct**
ct* ct ct*
ct ct
3 θ ray of light x*
2
1 x**
x x
1 2 3 (a) x (c)
(b)
Figure 19.10. Sketch (a) shows the world line for the
origin of observer B’s coordinate system. Sketch (b)
shows the oblique coordinate system for observer B and
sketch (c) shows the coordinate system of observer C who
is moving to the left.
ct*
ct
x*
B
A
x
l0
Now if you were very attentive, you would have noticed something
strange. The ruler, as shown in the figure, is longer in B than in A,
as you can appreciate from the fact that in B it is the hypotenuse
of a triangle and in A it is along the adjacent side. You know that
676 19. THE SPECIAL THEORY OF RELATIVITY
from the moving system, the ruler must appear shorter. The problem
disappears, however, when you realize that we have not decided on the
scale of the various axes. (The ruler must be shorter according to B.
This means that the “tics” along the x∗ -axis must be further apart
than the tics on the x-axis. I’ll get to this point in a moment.) Just
to make things easier to analyze, in Figure 19.12 I moved the origin
of coordinates so that one end of the ruler coincides with the origin in
both systems.
ct*
ct
x*
b
tic*
y
Ob
O x x
f2l0 tic
l0
Figure 19.12 indicates three lengths. The first one is l0 , the length
of the ruler in the system A, which you can take to be one meter. The
next one is Ob, the length of the ruler as measured by B. It is shortened
by a factor f that has not yet been determined. Also we do not (yet)
know the scale for the starred axes. With all these limitations, you
might well ask how you could possibly determine the length Ob. Well,
there is a clever way around the problem, and that involves going back
to the A system that does have a scale. Imagine that B happens to have
a rod that is exactly the length Ob. If A observes this rod, it will appear
to have shrunk by another factor of f. So in the A system, this rod has
a length f 2 l0 . All of this fiddling around with rulers and going from
one inertial system to another, will allow you to determine the amount
f by which moving objects shrink, because now it is a straightforward
geometry problem. In the figure the angle θ = tan−1 (v/c) is indicated
in several places. You can see that
f 2 l0 = l0 − δx
19.9. 4-VECTORS 677
f 2 = 1 − tan2 θ = 1 − (v 2 /c2 ).
p
Exercise 19.13. Show that cos θ = 1/ 1 + (v 2 /c2 ).
19.9. 4-Vectors
An event is specified by its position and time in a particular co-
ordinate system. It is convenient to introduce a four-dimensional co-
ordinate system in which an event is specified by the four coordinates
678 19. THE SPECIAL THEORY OF RELATIVITY
(x0 , x1 , x2 , x3 ) where
x0 = ct,
x1 = x,
x2 = y,
x3 = z.
(By multiplying the time coordinate by c, all coordinates are distances,
although x0 is obviously the coordinate that gives the time when the
event occurred.)
Just as three spatial coordinates describe an ordinary three-dimensional
vector, so too, the four space-time coordinates (x0 , x1 , x2 , x3 ) describe
a “4-vector” that starts at the origin and ends at some event E. A
4-vector from event E 1 to event E 2 has components (x20 − x10 ), (x21 −
x11 ), (x22 − x12 ), (x23 − x13 ). Note that the superscripts indicate the event.
Unfortunately, this makes for a somewhat confusing notation because
powers are also superscripts. For example, the 4-distance (or “space-
time interval”) between two events is given by
S = (x21 − x11 )2 + (x22 − x12 )2 + (x23 − x13 )2 − (x20 − x10 )2 .
Also note the change in sign on the temporal term. This 4-distance is,
of course, just the space-time distance defined previously by Equation
(19.12).
An important property of S is the fact that its value is unchanged
under a Lorentz transformation that expresses S in some other inertial
coordinate system. A quantity that is invariant under the transforma-
tion of a 3-D coordinate system is called a scalar, as you recall from
the discussion in Section 16.1. Analogously, S is called a four-scalar
(or sometimes, a “world scalar.”)
A more elegant way to write S uses the following “metric” (see
Section 5.3.4):
g0 = −1,
g1 = g2 = g3 = 1,
so that X
S= gµ (x2µ − x1µ )2 .
µ
positive, the interval between the two events is called space-like. (In
this case each event is within the light cone of the other one.) For such
an interval, it is possible to define the real quantity “proper distance”
between the two events by
σ = S 1/2 .
This quantity is the distance between the two events in a coordinate
system in which the events are simultaneous.
If S is negative, a real “proper time” can be defined by
(−S)1/2
τ= .
c
Physically, this is the time interval between two events in a coordinate
system in which the two events occur at the same place. A simple
example is when a light in my rocket ship flashes twice. In my reference
frame the two events occur at the same location.
5A
purist would point out that the scalar product is the product of a contravari-
ant vector with a covariant vector. See Section 9.8.
19.9. 4-VECTORS 681
break apart nor do particles fuse together). The second condition will
be removed shortly.
Next, assume that the 4-momentum is conserved. (I will not prove
that the 4-momentum is conserved, but I assure you that this conser-
vation is confirmed by a vast body of experimental evidence and it can
also be derived theoretically.) Then the total 4-momentum is the same
at all times, i.e.,
(Pµ )t1 = (Pµ )t2 . (19.17)
What are the consequences of this statement? Equation (19.17) states
that the 3-momentum is conserved, and this is as expected, but it also
means that P0 is constant. That is,
X
γj mj c = constant,
j
where
−1/2
u2j
γj = 1 − 2 .
c
Expanding this expression, using the binomial expansion, yields
−1/2
X X u2j
P0 = γj mj c = c 1− 2 mj ,
j j
c
X 1 u2j 3 u4j
= c 1+ + + · · · mj ,
j
2 c2 8 c4
X 1X1
= mj c + mj u2j + · · ·
j
c 2
or !
X
cP0 = mj c2 + T + O(1/c2 ). (19.18)
j
P1
The last term is negligible. I wrote T for m u2 because that is
2 j j
the kinetic energy of the system of particles. Note that in relativity,
energy and momentum have become enmeshed in the single quantity 4-
momentum. Einstein identified the quantity cP0 with the total energy
E.
E = cP0 .
We can also express the total energy as E = γmc2 .
Writing Equation (19.18) for a single particle,
.
E = mc2 + T (19.19)
19.10. RELATIVISTIC DYNAMICS 685
which states that the total energy of particle is given by its kinetic
energy plus a quantity mc2 . You might suspect that mc2 is a constant
because it is nearly always acceptable to add a constant to the energy
and only consider energy differences. Thus, you might feel that rel-
ativity theory has only resulted in adding a constant to the energy.
But if the total energy E is constant, this interpretation would mean
that T, the kinetic energy, is also constant, and obviously that is not
necessarily true. In the interaction between two particles (a collision),
the kinetic energy is conserved only if the interactions are perfectly
elastic. Otherwise, kinetic energy is lost, and according to the usual
description, energy has been converted into heat. According to rela-
tivity theory, the term mc2 represents all other forms of energy (heat,
potential energy, chemical energy, etc.). Equation (19.19) states that
the total energy E is equal to the kinetic energy T plus all other forms
of energy. Thus, if you heat an object (at rest), the value of mc2 must
increase. Recall that m is the rest mass. This means that the heated
object is more massive than the same object when cool.
The total energy of an object at rest (so that T = 0) is given by
the well known equation
E = mc2 .
For example, if you heat a gram of water by one degree centigrade,
you supply it with one calorie (4.186 joules) of heat. You will have
increased the mass of the water by 4.18/c2 = 4.64 × 10−17 kg. This is
unmeasurable. In like manner, if a nucleus splits apart (fission) and
the fragments are less massive than the original nucleus, then energy
is emitted. This energy is, as we all know, quite significant.
It is sometimes convenient to write the rest mass of a particle as
m0 . Then the momentum of a particle is
Pi = γm0 ui , (i = 1, 2, 3)
and the “relativistic mass” mrel is defined as
m0
mrel = γm0 = p ,
1 − v 2 /c2
where v is the velocity of the particle. This notation suggests that the
mass of a moving particle increases with its velocity and tends toward
infinity as the speed approaches c. You might consider the concept of
relativistic mass to be a needless variant, and, in fact, modern physi-
cists tend to dismiss it. Nevertheless, it is helpful in solving collision
problems involving relativistic particles, particularly since relativistic
mass is conserved in collisions.
686 19. THE SPECIAL THEORY OF RELATIVITY
19.11. Summary
In this brief review of the theory of relativity, I only considered
aspects of the theory that relate to mechanics and did not delve at all
into electromagnetic phenomena. Nevertheless, you should be aware
that it was electromagnetic considerations that led Einstein to develop
his theory.
The most important concepts of the theory are embodied in the
Lorentz transformations, from which time dilation and length contrac-
tion follow. These transformations lead to a number of unusual effects,
including paradoxes such as the relative aging of moving twins. The
Lorentz transformations are
x − vt
x∗ = p ,
1 − v 2 /c2
y ∗ = y,
z ∗ = z,
t − (xv/c2 )
t∗ = p .
1 − v 2 /c2
Time dilation is described by
τ0
τ=p ,
1 − v 2 /c2
and length contraction is given by
p
l = l0 1 − v 2 /c2 .
The relativistic addition of velocities is given by
u∗ + v
u= ∗ .
1 + uc2v
19.12. PROBLEMS 687
19.12. Problems
Problem 19.1. An observer in R measures an event at x = 3 m
and t = 7 ns. The R∗ frame is moving along the common x-axis at
v = 0.6c. (a) What are the relativistic space time coordinates measured
by the observer in R∗ ? (b) What space time coordinates would be
measured in R∗ if the Galilean transformation law held?
Problem 19.2. Use the Lorentz transformations to show that an
observer moving at speed v toward a clock at rest in some reference
frame, will find the clock to read a time −lv/c2 compared to the reading
by an observer in the rest frame of the clock. At the instant of the
measurements the two observers are at the same location and at a
(rest frame) distance l from the clock.
Problem 19.3. John, the observer in R, sets out three synchro-
nized clocks at locations x = −l, 0, +l. Mary (in frame R∗ traveling at
speed v along x) observes the times on the clocks at the instant she
688 19. THE SPECIAL THEORY OF RELATIVITY
the Earth were at rest and a particular star were directly overhead, you
would observe it by pointing your telescope straight up. But since the
Earth is moving, you must compensate by pointing your telescope at
a small angle away from straight up.) The aberration of light means
that a ray of light that makes an angle θ∗ with the x axis in the rest
frame will make an angle θ in the moving frame where
p
sin θ∗ 1 − v 2 /c2
tan θ = .
cos θ∗ + v/c
Consider a light source at rest in R∗ . The fraction of light emitted into
a cone of half angle θ∗ is
1
f = (1 − cos θ∗ ) .
2
In this problem, you will show that in a moving reference frame, the
same amount of light is emitted into a smaller cone of half angle θ. (a)
Derive the relationship f = 0.5 (1 − cos θ∗ ) . (b) Assume that θ∗ = 40◦
and determine f in a coordinate system in which the light source is
at rest. (c) Calculate the half angle of a cone in a frame moving at
v = 0.9c into which the same amount of light will be emitted. (d) Why
do you think this is called the “headlight” effect?
Problem 19.18. Evaluate the work done on an electron when it
is accelerated from rest to 0.98c. Answer: 3.3 × 1013 J.
Problem 19.19. Determine the speed and kinetic energy of a par-
ticle whose momentum is equal to m0 c.
Problem 19.20. Show that a particle that travels at the speed of
light must have zero rest mass.
Problem 19.21. A proton is traveling at 0.995c in the laboratory
frame. What is its kinetic energy? What is its momentum?
Problem 19.22. Two identical particles of rest mass m0 approach
one another at speeds ±0.5c, and undergo a completely inelastic col-
lision. What is the momentum and energy of the final (composite)
particle? Note that the momentum of a particle is given by p = γm0 v,
where γ = γ(v) depends on the speed of that particular particle.)
Problem 19.23. In some particular reference frame (perhaps the
laboratory), a particle traveling at 0.8c collides with and sticks to an
identical particle at rest. (a) Determine the momentum of the resulting
combined particle. (b) What is its speed? (c) What does classical
physics yield for the speed? (You may assume that when the particles
are at rest they have mass m0 .)
19.12. PROBLEMS 691
693
694 20. CLASSICAL CHAOS (OPTIONAL)
Again, at any instant of time, the state of a system is given by the values
of p and x. For a one-dimensional system, these can be represented by
a point on a p vs x plot. As time goes on the values of p and x will
change and the point will trace out a path in the px plane. This plane
is called “phase space” and the path is called a phase space trajectory.
For most simple systems, the momentum and the velocity are related
by p = mẋ, so plots of ẋ vs x are also called phase space plots (or phase
space “portraits”).
Although chaotic motion is not periodic, it is convenient to begin
our study by considering some aspects of periodic motion. You are
familiar with two periodic systems: simple harmonic oscillators and
the Kepler (two-body) problem. The Kepler problem is interesting
because it involves two different periodicities, namely, the periodicity
in the radial motion and the periodicity in the angular motion.
Recall that the radial position of a particle in an elliptical orbit is
a (1 − e2 )
r= .
1 + e cos θ
The angular velocity is given by
l
θ̇ = .
mr2
Now the unperturbed two body system, as considered in Chapter 10
is degenerate, that is, the period of the radial motion is equal to the
period of the angular motion, and the orbit is repeated over and over
again. A configuration space plot of r vs θ is simply the elliptical orbit
of the particle.
The Hamiltonian for a Keplerian particle is a function of pr , pθ , r, θ.
It is interesting to generate phase space plots for this motion. Note,
again, that one normally plots velocity rather than momentum vs po-
sition, that is, one plots ṙ vs r and θ̇ vs θ. Figure 20.1 gives these plots
for various values of e, using arbitrary units.
H = H(p1 , p2 , · · · , pn ; q1 , q2 , · · · qn ; t)
. ∂H . ∂H
pi = − and qi = + .
∂qi ∂pi
20.2. PERIODIC MOTION 697
(See Equations 4.20.) Recall, further, that under most usual conditions,
the Hamiltonian reduces to the total energy and one can write
H = T + V.
Let us consider, as an example of periodic motion, a double os-
cillator in which the motions in mutually perpendicular directions are
independent. The Hamiltonian for such a system can be written as
(p01 )2 1 2 (p0 )2 1 2
H= + m1 ω12 (q10 ) + 2 + m2 ω22 (q20 ) (20.1)
2m1 2 2m2 2
which is just E1 + E2 = total energy. It is convenient to transform to
“normalized” coordinates pi and qi defined by
r
p0i 0 1
pi = √ and qi = qi mi ωi2 .
2mi 2
In terms of these p’s and q’s the Hamiltonian is
H = p21 + q12 + p22 + q22 = E1 + E2 .
√
It is clear that the phase space plots are simply circles of radius E,
as shown in Figure 20.2
p1 p2
q1 q2
p2
q1
q2 p1
20.3. Attractors
If a bounded, periodic, dynamical system is subjected to a small
perturbation, its behavior will change somewhat, but you would expect
it to still be bounded. The Russian mathematician Kolmogorov and,
later, Arnold and Moser developed and proved a theorem to that effect.
It has become known as the KAM theorem and it states that a bounded
system subjected to a small perturbation will remain bounded. How-
ever, the theorem does allow for chaotic motion to occur in some sys-
tems if the initial conditions are just right.
Consider a periodic system acted upon by a small perturbing force.
The orbit may decay and the phase space trajectory might end up at
some point in phase space and remain at that point from then on. Or
the system might evolve to a stable orbit (called a “limit cycle”) and
then go around in that orbit from then on. In these cases, the point or
the stable orbit are referred to as “attractors.”
In other words, an attractor is a set of points in phase space to
which the solution evolves, usually after a long interval of time. As an
example, consider a simple pendulum. The motion of an unperturbed
pendulum can be described by a circle in phase space. But if there is
a small perturbation, say a weak drag force, then the amplitude of the
pendulum swings will grow smaller and smaller and the momentum
will gradually decrease, until the pendulum comes to rest. If you were
to plot this on a phase space diagram you would see the trajectory is
20.3. ATTRACTORS 699
Exercise 20.3. Plot the phase space trajectory for a damped os-
cillator.
A’ A
x
px
A’ A
x
Note that when y = 0 the planet is not always at the perihelion and
the point going through the x, px plane will gradually move along the
trajectory shown in Figure 20.8. (The point representing the system
configuration gradually moves from A to B to C and so on, generating
the x, px curve shown.)
Thus, a Poincaré section is a 2-D slice through a 3-D energy hy-
persurface in a 4-D phase space. For higher dimension phase spaces
it is possible to draw Poincaré surfaces, but their usefulness is mainly
limited to 4-D phase space.
20.6. THE HENON HEILES HAMILTONIAN 703
px
C B
A x
V = 1/6
V = 1/8
V = 1/12
V = 1/24
20.7. Summary
In this very brief introduction to classical chaos, I have attempted
to give you some of the vocabulary used in the theory and to give you a
general idea of the basic ideas of chaos. Note that nonlinear equations
are sometimes very sensitive to initial conditions, in which case they
will lead to chaotic motion. Chaos is characterized by the generation
of random points on a phase portrait. The effect of initial conditions
706 20. CLASSICAL CHAOS (OPTIONAL)
[6] A. L. Fetter and J. D. Walecka Theoretical Mechanics for Particles and Con-
tinua McGraw Hill, 1980.
[7] H. Goldstein, C. P. Poole and J. Safko Classical Mechanics, 3d Ed. Pearson,
2011.
[8] P. Hamill A Student’s Guide to Lagrangians and Hamiltonians Cambridge
University Press, 2013.
[9] L. D. Landau and E. M. Lifshitz Mechanics, 3d Ed. Butterworth-Heinemann,
1976.
709
Appendix A
Formulas and Constants
Universal Constants
Gravitational constant = G = 6.67 × 10−11 Nm2 /kg2
Speed of light = c = 3 × 108 m/s
Astronomical Constants
Mass of Earth 5.97 × 1024 kg
Mass of Sun 1.99 × 1030 kg
Radius of Earth 6.37 × 106 m
Radius of the orbit of Earth around Sun 1.5 × 1011 m
ez − e−z
sinh z =
2
ez + e−z
cosh z =
2
sinh z
tanh z =
cosh z
711
712 A. FORMULAS AND CONSTANTS
Integrals
Expressions involving terms like a2 ± x2
Z
dx 1 x
= arctan
a2+ x2 a a
Z
dx 1 x
= tanh−1
a − x2
2 a a
Z
dx x
√ = arcsin
a2 − x 2 a
Z
dx √
√ = ln(x + x2 − a2 )
x 2 − a2
Z
dx 2 2cx + b
= √ tan−1 √
a + bx + cx2 4ac − b2 4ac − b2
Z
dx √1 sinh−1 √2cx+b 2 if c > 0
c
√ = 4ac−b
a + bx + cx2 √1 sin−1 √−2cx−b if c < 0
−c b2 −4ac
Error Function Z x
2 2
erf(x) = √ e−t dt
π 0
Appendix B
Answers to Selected Problems
1.1 35 mph
1.5 (a) v1 +v
2
2
(b) v2v1 +v
1 v2
2
1.7 0.69 mi
1.9 (c) v = v0 [(sin θ − µ cos θ)/(sin θ + µ cos θ)]1/2
1.11 29 cm
1.13 2.58 × 1016 N
1.15 176 watts
1.17 (b) 2857 rad
1.19 Stable
1.21 (b) 189 sec
2.1 6m/s
2.5 48.5m/s
2.7 23.65◦ and 66.35◦
2.11 5609m h i1/2
2 x2
2.12(b) v = v02 − 2gx tan θo + v2gcos 2θ
gx
, tan θ = tan θ0 − v2 cos2θ .
0
0 0
2.13 15.7m2 2
v cos θ sin θ
2.15 s = 0 2g [ cos 2 θ + log(sec θ + tan θ)]
2.18 v =30m/s
2.19 R=455km
2.20 ac =75m/sp
2.21(a) τ = 2π r/a.
2.22 ω = 10.35π cos πtî + 10.35π sin πtĵ + 39.64π k̂
2.26 -24
2.30 (a) This is the spiral of Archimedes. (b) v = (4 + 100t2 )1/2
m/s.
2.31 v = (cos θ)r̂ + (1 + sin t)e−t θ̂
715
716 B. ANSWERS TO SELECTED PROBLEMS