Part IA Dynamics and Relativity
Based on lectures by G. I. Ogilvie
Notes taken by Dexter Chua
Lent 2015
These notes are not endorsed by the lecturers, and I have modified them (often
significantly) after lectures. They are nowhere near accurate representations of what
was actually lectured, and in particular, all errors are almost surely mine.
Familiarity with the topics covered in the non-examinable Mechanics course is assumed.
Basic concepts
Space and time, frames of reference, Galilean transformations. Newton’s laws. Dimen-
sional analysis. Examples of forces, including gravity, friction and Lorentz. [4]
Newtonian dynamics of a single particle
Equation of motion in Cartesian and plane polar coordinates. Work, conservative forces
and potential energy, motion and the shape of the potential energy function; stable
equilibria and small oscillations; effect of damping.
Angular velocity, angular momentum, torque.
Orbits: the
u
(
θ
) equation; escape velocity; Kepler’s laws; stability of orbits; motion in
a repulsive potential (Rutherford scattering). Rotating frames: centrifugal and coriolis
forces. *Brief discussion of Foucault pendulum.* [8]
Newtonian dynamics of systems of particles
Momentum, angular momentum, energy. Motion relative to the centre of mass; the
two body problem. Variable mass problems; the rocket equation. [2]
Rigid bodies
Moments of inertia, angular momentum and energy of a rigid body. Parallel axes
theorem. Simple examples of motion involving both rotation and translation (e.g.
rolling). [3]
Special relativity
The principle of relativity. Relativity and simultaneity. The invariant interval. Lorentz
transformations in (1 + 1)-dimensional spacetime. Time dilation and length contraction.
The Minkowski metric for (1 + 1)-dimensional spacetime. Lorentz transformations
in (3 + 1) dimensions. 4-vectors and Lorentz invariants. Prop er time. 4-velocity
and 4-momentum. Conservation of 4-momentum in particle decay. Collisions. The
Newtonian limit. [7]
Contents
0 Introduction
1 Newtonian dynamics of particles
1.1 Newton’s laws of motion
1.2 Galilean transformations
1.3 Newton’s Second Law
2 Dimensional analysis
2.1 Units
2.2 Scaling
3 Forces
3.1 Force and potential energy in one dimension
3.2 Motion in a potential
3.3 Central forces
3.4 Gravity
3.5 Electromagnetism
3.6 Friction
4 Orbits
4.1 Polar coordinates in the plane
4.2 Motion in a central force field
4.3 Equation of the shape of the orbit
4.4 The Kepler problem
4.5 Rutherford scattering
5 Rotating frames
5.1 Motion in rotating frames
5.2 The centrifugal force
5.3 The Coriolis force
6 Systems of particles
6.1 Motion of the center of mass
6.2 Motion relative to the center of mass
6.3 The two-body problem
6.4 Variable-mass problem
7 Rigid bodies
7.1 Angular velocity
7.2 Moment of inertia
7.3 Calculating the moment of inertia
7.4 Motion of a rigid body
8 Special relativity
8.1 The Lorentz transformation
8.2 Spacetime diagrams
8.3 Relativistic physics
8.4 Geometry of spacetime
8.5 Relativistic kinematics
8.6 Particle physics
0 Introduction
You’ve been lied to. You thought you applied for mathematics. And here you
have a course on physics. No, this course is not created for students taking the
“Maths with Physics” option. They don’t have to take this course (don’t ask
why).
Ever since Newton invented calculus, mathematics is becoming more and more
important in physics. Physicists seek to describe the universe in a few equations,
and derive everyday (physical) phenomena as mathematical consequences of
these equations.
In this course, we will start with Newton’s laws of motion and use it to derive
a lot of physical phenomena, including planetary orbits, centrifugal forces
a
and
the motion of rotating bodies.
The important thing to note is that we can “prove” all these phenomena just
under the assumption that Newton’s laws are correct (plus the formulas for, say,
the strength of the gravitational force). We are just doing mathematics here.
We don’t need to do any experiments to obtain the results (of course, we need
experiments to verify that Newton’s laws are indeed the equations that describe
this universe).
However, it turns out that Newton was wrong. While his theories were
accurate for most everyday phenomena, they weren’t able to adequately describe
electromagnetism. This lead to Einstein discovering special relativity. Special
relativity is also required to describe motion that is very fast. We will have a
brief introduction to special relativity at the end of the course.
a
Yes, they exist.
1 Newtonian dynamics of particles
Newton’s equations describe the motion of a (point) particle.
Definition
(Particle)
.
A particle is an object of insignificant size, hence it can
be regarded as a point. It has a mass m > 0, and an electric charge q.
Its position at time
t
is described by its position vector,
r
(
t
) or
x
(
t
) with
respect to an origin O.
Depending on context, different things can be considered as particles. We
could consider an electron to be a point particle, even though it is more accurately
described by the laws of quantum mechanics than those of Newtonian mechanics.
If we are studying the orbit of planets, we can consider the Sun and the Earth
to be particles.
An important property of a particle is that it has no internal structure. It can
be completely described by its position, momentum, mass and electric charge.
For example, if we model the Earth as a particle, we will have to ignore its own
rotation, temperature etc.
If we want to actually describe a rotating object, we usually consider it as a
collection of point particles connected together, and apply Newton’s law to the
individual particles.
As mentioned above, the position of a particle is described by a position
vector. This requires us to pick an origin of the coordinate system, as well as an
orientation of the axes. Each choice is known as a frame of reference.
Definition
(Frame of reference)
.
A frame of reference is a choice of coordinate
axes for r.
We don’t impose many restrictions on the choice of coordinate axes. They
can be fixed, moving, rotating, or even accelerating.
Using the position vector
r
, we can define various interesting quantities which
describe the particle.
Definition (Velocity). The velocity of the particle is
v =
˙
r =
dr
dt
.
Definition (Acceleration). The acceleration of the particle is
a =
˙
v =
¨
r =
d
2
r
dt
2
.
Definition (Momentum). The momentum of a particle is
p = mv = m
˙
r.
m
is the inertial mass of the particle, and measures its reluctance to accelerate,
as described by Newton’s second law.
1.1 Newton’s laws of motion
We will first state Newton’s three laws of motion, and then discuss them individ-
ually.
Law
(Newton’s First Law of Motion)
.
A body remains at rest, or moves uniformly
in a straight line, unless acted on by a force. (This is in fact Galileo’s Law of
Inertia)
Law
(Newton’s Second Law of Motion)
.
The rate of change of momentum of a
body is equal to the force acting on it (in both magnitude and direction).
Law
(Newton’s Third Law of Motion)
.
To every action there is an equal and
opposite reaction: the forces of two bodies on each other are equal and in opposite
directions.
The first law might seem redundant given the second if interpreted literally.
According to the second law, if there is no force, then the momentum doesn’t
change. Hence the body remains at rest or moves uniformly in a straight line.
So why do we have the first law? Historically, it might be there to explicitly
counter Aristotle’s idea that objects naturally slow down to rest. However, some
(modern) physicists give it an alternative interpretation:
Note that the first law isn’t always true. Take yourself as a frame of reference.
When you move around your room, things will seem like they are moving around
(relative to you). When you sit down, they stop moving. However, in reality,
they’ve always been sitting there still. On second thought, this is because you,
the frame of reference, is accelerating, not the objects. The first law only holds
in frames that are themselves not accelerating. We call these inertial frames.
Definition
(Inertial frames)
.
Inertial frames are frames of references in which
the frames themselves are not accelerating. Newton’s Laws only hold in inertial
frames.
Then we can take the first law to assert that inertial frames exists. Even
though the Earth itself is rotating and orbiting the sun, for most purposes, any
fixed place on the Earth counts as an inertial frame.
1.2 Galilean transformations
The goal of this section is to investigate inertial frames. We know that inertial
frames are not unique. Given an inertial frame, what other inertial frames can
we obtain?
First of all, we can rotate our axes or move our origin. In particular, we can
perform the following operations:
Translations of space:
r
0
= r r
0
Translations of time:
t
0
= t t
0
Rotation (and reflection):
r
0
= Rr
with R O(3).
These are not very interesting. They are simply symmetries of space itself.
The last possible transformation is uniform motion. Suppose that
S
is an
inertial frame. Then any other frame
S
0
in uniform motion relative to
S
is also
inertial:
S
S
0
y
x
y
0
x
0
v
Assuming the frames coincide at t = 0, then
x
0
= x vt
y
0
= y
z
0
= z
t
0
= t
Generally, the position vector transforms as
r
0
= r vt,
where
v
is the (constant) velocity of
S
0
relative to
S
. The velocity and acceleration
transform as follows:
˙
r
0
=
˙
r v
¨
r
0
=
¨
r
Definition
(Galilean boost)
.
A Galilean boost is a change in frame of reference
by
r
0
= r vt
t
0
= t
for a fixed, constant v.
All these transformations together generate the Galilean group, which de-
scribes the symmetry of Newtonian equations of motion.
Law
(Galilean relativity)
.
The principle of relativity asserts that the laws of
physics are the same in inertial frames.
This implies that physical processes work the same
at every point of space
at all times
in whichever direction we face
at whatever constant velocity we travel.
In other words, the equations of Newtonian physics must have Galilean
invariance.
Since the laws of physics are the same regardless of your velocity, velocity
must be a relative concept, and there is no such thing as an “absolute velocity”
that all inertial frames agree on.
However, all inertial frames must agree on whether you are accelerating or
not (even though they need not agree on the direction of acceleration since you
can rotate your frame). So acceleration is an absolute concept.
1.3 Newton’s Second Law
Newton’s second law is often written in the form of an equation.
Law. The equation of motion for a particle subject to a force F is
dp
dt
= F,
where
p
=
mv
=
m
˙
r
is the (linear) momentum of the particle. We say
m
is the
(inertial) mass of the particle, which is a measure of its reluctance to accelerate.
Usually, m is constant. Then
F = ma = m
¨
r.
Usually,
F
is specified as a function of
r,
˙
r
and
t
. Then we have a second-order
ordinary differential equation for r.
To determine the solution, we need to specify the initial values of
r
and
˙
r
, i.e.
the initial position and velocity. The trajectory of the particle is then uniquely
determined for all future (and past) times.
2 Dimensional analysis
When considering physical theories, it is important to be aware that physical
quantities are not pure numbers. Each physical quantity has a dimension.
Roughly speaking, dimensions are what units represent, such as length, mass
and time. In any equation relating physical quantities, the dimensions must be
consistent, i.e. the dimensions on both sides of the equation must be equal.
For many problems in dynamics, the three basic dimensions are sufficient:
length, L
mass, M
time, T
The dimensions of a physical quantity
X
, denoted by [
X
] are expressible
uniquely in terms of L, M and T . For example,
[area] = L
2
[density] = L
3
M
[velocity] = LT
1
[acceleration] = LT
2
[
force
] =
LMT
2
since the dimensions must satisfy the equation
F
=
ma
.
[energy] = L
2
MT
2
, e.g. consider E = mv
2
/2.
Physical constants also have dimensions, e.g. [
G
] =
L
3
M
1
T
2
by
F
=
GMm/r
2
.
The only allowed operations on quantities with dimensions are sums and
products (and subtraction and division), and if we sum two terms, they must
have the same dimension. For example, it does not make sense to add a length
with an area. More complicated functions of dimensional quantities are not
allowed, e.g. e
x
again makes no sense if x has a dimension, since
e
x
= 1 + x +
1
2
x
2
+ ···
and if
x
had a dimension, we would be summing up terms of different dimensions.
2.1 Units
People use units to represent dimensional quantities. A unit is a convenient
standard physical quantity, e.g. a fixed amount of mass. In the SI system, there
are base units corresponding to the basics dimensions. The three we need are
meter (m) for length
kilogram (kg) for mass
second (s) for time
A physical quantity can be expressed as a pure number times a unit with the
correct dimensions, e.g.
G = 6.67 × 10
11
m
3
kg
1
s
2
.
It is important to realize that SI units are chosen arbitrarily for historical reasons
only. The equation of physics must work in any consistent system of units. This
is captured by the fact that physical equations must be dimensionally consistent.
2.2 Scaling
We’ve had so many restrictions on dimensional quantities equations must
be dimensionally consistent, and we cannot sum terms of different dimensions.
However, this is not a hindrance when developing new theories. In fact, it is a
very useful tool. First of all, it allows us to immediately rule out equations that
do not make sense dimensionally. More importantly, it allows us to guess the
form of the equation we want.
Suppose we believe that a physical quantity
Y
depends on 3 other physical
quantities
X
1
, X
2
, X
3
, i.e.
Y
=
Y
(
X
1
, X
2
, X
3
). Let their dimensions be as
follows:
[Y ] = L
a
M
b
T
c
[X
i
] = L
a
i
M
b
i
T
c
i
Suppose further that we know that the relationship is a power law, i.e.
Y = CX
p
1
1
X
p
2
2
X
p
3
3
,
where
C
is a dimensionless constant (i.e. a pure number). Since the dimensions
must work out, we know that
a = p
1
a
1
+ p
2
a
2
+ p
3
a
3
b = p
1
b
1
+ p
2
b
2
+ p
3
b
3
c = p
1
c
1
+ p
2
c
2
+ p
3
c
3
for which there is a unique solution provided that the dimensions of
X
1
, X
2
and
X
3
are independent. So just by using dimensional analysis, we can figure
out the relation between the quantities up to a constant. The constant can
then be found experimentally, which is much easier than finding the form of the
expression experimentally.
However, in reality, the dimensions are not always independent. For example,
we might have two length quantities. More importantly, the situation might
involve more than 3 variables, and we do not have a unique solution.
First consider a simple case if
X
2
1
X
2
is dimensionless, then the relation
between
Y
and
X
i
can involve more complicated terms such as
exp
(
X
2
1
X
2
), since
the argument of exp is now dimensionless.
In general, suppose we have many terms, and the dimensions of
X
i
are not in-
dependent. We order the quantities so that the independent terms [
X
1
]
,
[
X
2
]
,
[
X
3
]
are at the front. For each of the remaining variables, form a dimensionless quan-
tity λ
i
= X
i
X
q
1
1
X
q
2
2
X
q
3
3
. Then the relationship must be of the form
Y = f(λ
4
, λ
5
, ···)X
p
1
1
X
p
2
2
X
p
3
3
.
where f is an arbitrary function of the dimensionless variables.
Formally, this results is described by the Buckingham’s Pi theorem, but we
will not go into details.
Example (Simple pendulum).
m
`
d
We want to find the period P . We know that P could depend on
mass m: [m] = M
length `: [`] = L
gravity g: [g] = LT
2
initial displacement d: [d] = L
and of course [P ] = T .
We observe that
m, `, g
have independent dimensions, and with the fourth,
we can form the dimensionless group
d/`
. So the relationship must be of the
form
P = f
d
l
m
p
1
`
p
2
g
p
3
,
where f is a dimensionless function. For the dimensions to balance,
T = M
p
1
L
p
2
L
p
3
T
2p
3
.
So p
1
= 0, p
2
= p
3
= 1/2. Then
P = f
d
`
s
`
g
.
While we cannot find the exact formula, using dimensional analysis, we know
that if both ` and d are quadrupled, then P will double.
3 Forces
Force is a central concept in Newtonian mechanics. As described by Newton’s
laws of motion, forces are what causes objects to accelerate, according to the
formula
F
=
ma
. To completely specify the dynamics of a system, one only
needs to know what forces act on what bodies.
However, unlike in Star Wars, the force is given a secondary importance in
modern treatments of mechanics. Instead, the potential is what is considered to
be fundamental, with force being a derived concept. In quantum mechanics, we
cannot even meaningfully define forces.
Yet, there are certain systems that must be described with forces instead
of potentials, the most important of which is a system that involves friction of
some sort.
3.1 Force and potential energy in one dimension
To define the potential, consider a particle of mass
m
moving in a straight line
with position
x
(
t
). Suppose
F
=
F
(
x
), i.e. it depends on position only. We
define the potential energy as follows:
Definition
(Potential energy)
.
Given a force field
F
=
F
(
x
), we define the
potential energy to be a function V (x) such that
F =
dV
dx
.
or
V =
Z
F dx.
V
is defined up to a constant term. We usually pick a constant of integration
such that the potential drops to 0 at infinity.
Using the potential, we can write the equation of motion as
m¨x =
dV
dx
.
There is a reason why we call this the potential energy. We usually consider
it to be an energy of some sort. In particular, we define the total energy of a
system to be
Definition (Total energy). The total energy of a system is given by
E = T + V,
where V is the potential energy and T =
1
2
m ˙x
2
is the kinetic energy.
If the force of a system is derived from a potential, we can show that energy
is conserved.
Proposition. Suppose the equation of a particle satisfies
m¨x =
dV
dx
.
Then the total energy
E = T + V =
1
2
m ˙x
2
+ V (x)
is conserved, i.e.
˙
E = 0.
Proof.
dE
dt
= m ˙x¨x +
dV
dx
˙x
= ˙x
m¨x +
dV
dx
= 0
Example. Consider the harmonic oscillator, whose potential is given by
V =
1
2
kx
2
.
Then the equation of motion is
m¨x = kx.
This is the case of, say, Hooke’s Law for a spring.
The general solution of this is
x(t) = A cos(ωt) + B sin(ωt)
with ω =
p
k/m.
A and B are arbitrary constants, and are related to the initial position and
velocity by x(0) = A, ˙x(0) = ωB.
For a general potential energy
V
(
x
), conservation of energy allows us to solve
the problem formally:
E =
1
2
m ˙x
2
+ V (x)
Since E is a constant, from this equation, we have
dx
dt
= ±
r
2
m
(E V (x))
t t
0
= ±
Z
dx
q
2
m
(E V (x))
.
To find
x
(
t
), we need to do the integral and then solve for
x
. This is usually
not possible by analytical methods, but we can approximate the solution by
numerical methods.
3.2 Motion in a potential
Given an arbitrary potential
V
(
x
), it is often difficult to completely solve the
equations of motion. However, just by looking at the graph of the potential, we
can usually get a qualitative understanding of the dynamics.
Example.
Consider
V
(
x
) =
m
(
x
3
3
x
). Note that this can be dimensionally
consistent even though we add up
x
3
and
3
x
, if we declare “3” to have dimension
L
2
.
We plot this as follows:
x
V
O
1 212
Suppose we release the particle from rest at
x
=
x
0
. Then
E
=
V
(
x
0
). We can
consider what happens to the particle for different values of x
0
.
x
0
= ±1: This is an equilibrium and the Particle stays there for all t.
1
< x
0
<
2: The particle does not have enough energy to escape the well.
So it oscillates back and forth in potential well.
x
0
< 1: The particle falls to x = −∞.
x
0
>
2: The particle has enough energy to overshoot the well and continues
to x = −∞.
x
0
= 2: This is a special case. Obviously, the particle goes towards
x
=
1.
But how long does it take, and what happens next? Here we have
E
= 2
m
.
We noted previously
t t
0
=
Z
dx
q
2
m
(E V (x))
.
Let x = 1 + ε(t). Then
2
m
(E V (x)) = 4 2(1 + ε)
3
+ 6(1 + ε)
= 6ε
2
2ε
3
.
So
t t
0
=
Z
ε
3
dε
0
6ε
2
2ε
3
We reach
x
=
1 when
ε
0. But for small
ε
, the integrand is approx-
imately
1
, which integrates to
ln ε −∞
as
ε
0. So
ε
= 0 is
achieved after infinite time, i.e. it takes infinite time to reach
ε
= 0, or
x = 1.
Equilibrium points
In reality, most of the time, particles are not flying around wildly doing crazy
stuff. Instead, they stay at (or near) some stable point, and only move very little
in a predictable manner. We call these points equilibrium points.
Definition
(Equilibrium point)
.
A particle is in equilibrium if it has no tendency
to move away. It will stay there for all time. Since
m¨x
=
V
0
(
x
), the equilibrium
points are the stationary points of the potential energy, i.e.
V
0
(x
0
) = 0.
Consider motion near an equilibrium point. We assume that the motion is
small and we can approximate
V
by a second-order Taylor expansion. Then we
can write V as
V (x) V (x
0
) +
1
2
V
00
(x
0
)(x x
0
)
2
.
Then the equation of motion is
m¨x = V
00
(x
0
)(x x
0
).
If
V
00
(
x
0
)
>
0, then this is of the form of the harmonic oscillator.
V
has a
local minimum at
x
0
, and we say the equilibrium point is stable. The particle
oscillates with angular frequency
ω =
r
V
00
(x
0
)
m
.
If
V
00
(
x
0
)
<
0, then
V
has a local maximum at
x
0
. In this case, the equilibrium
point is unstable, and the solution to the equation is
x x
0
Ae
γt
+ Be
γt
for
γ =
r
V
00
(x
0
)
m
.
For almost all initial conditions,
A 6
= 0 and the particle will diverge from the
equilibrium point, leading to a breakdown of the approximation.
If V
00
(x
0
) = 0, then further work is required to determine the outcome.
Example. Consider the simple pendulum.
m
`
d
θ
Suppose that the pendulum makes an angle
θ
with the vertical. Then the energy
is
E = T + V =
1
2
m`
2
˙
θ
2
mg` cos θ.
Therefore
V cos θ
. We have a stable equilibrium at
θ
= 0, and unstable
equilibrium at θ = π.
θ
V
π
π
mg`
mg`
If E > mg`, then
˙
θ never vanishes and the pendulum makes full circles.
If 0
< E < mg`
, then
˙
θ
vanishes at
θ
=
±θ
0
for some 0
< θ
0
< π
i.e.
E
=
mg` cos θ
0
. The pendulum oscillates back and forth. It takes a quarter
of a period to reach from
θ
= 0 to
θ
=
θ
0
. Using the previous general solution,
oscillation period P is given by
P
4
=
Z
θ
0
0
=
dθ
q
2E
m`
2
+
2g
`
cos θ
.
Since we know that E = mg` cos θ
0
, we know that
P
4
=
s
`
g
Z
θ
0
0
dδ
2 cos θ 2 cos θ
0
.
The integral is difficult to evaluate in general, but for small
θ
0
, we can use
cos θ 1
1
2
θ
2
. So
P 4
s
`
g
Z
θ
0
0
dθ
p
θ
2
0
θ
2
= 2π
s
`
g
and is independent of the amplitude
θ
0
. This is of course the result for the
harmonic oscillator.
Force and potential energy in three dimensions
Everything looks nice so far. However, in real life, the world has (at least) three
(spatial) dimensions. To work with multiple dimensions, we will have to promote
our quantities into vectors.
Consider a particle of mass
m
moving in 3D. The equation of motion is now
a vector equation
m
¨
r = F.
We’ll define the familiar quantities we’ve had.
Definition
(Kinetic energy)
.
We define the kinetic energy of the particle to be
T =
1
2
m|v|
2
=
1
2
m
˙
r ·
˙
r.
If we want to know how it varies with time, we obtain
dT
dt
= m
¨
r ·
˙
r = F ·
˙
r = F · v.
This is the power.
Definition
(Power)
.
The power is the rate at which work is done on a particle
by a force. It is given by
P = F · v.
Definition
(Work done)
.
The work done on a particle by a force is the change
in kinetic energy caused by the force. The work done on a particle moving from
r
1
= r(t
1
) to r
2
= r(t
2
) along a trajectory C is the line integral
W =
Z
C
F · dr =
Z
t
2
t
1
F ·
˙
r dt =
Z
t
2
t
1
P dt.
Usually, we are interested in forces that conserve energy. These are forces
which can be given a potential, and are known as conservative forces.
Definition
(Conservative force and potential energy)
.
A conservative force is a
force field F(r) that can be written in the form
F = −∇V.
V is the potential energy function.
Proposition. If F is conservative, then the energy
E = T + V
=
1
2
m|v|
2
+ V (r)
is conserved. For any particle moving under this force, the work done is equal to
the change in potential energy, and is independent of the path taken between
the end points. In particular, if we travel around a closed loop, no work is done.
Proof.
dE
dt
=
d
dt
1
2
m
˙
r ·
˙
r + V
= m
¨
r ·
˙
r +
V
x
i
dx
i
dt
= (m
¨
r + V ) ·
˙
r
= (m
¨
r F) ·
˙
r
= 0
So the energy is conserved. In this case, the work done is
W =
Z
C
F · dr =
Z
C
(V ) · dr = V (r
1
) V (r
2
).
3.3 Central forces
While in theory the potential can take any form it likes, most of the time, our
system has spherical symmetry. In this case, the potential depends only on the
distance from the origin.
Definition
(Central force)
.
A central force is a force with a potential
V
(
r
) that
depends only on the distance from the origin,
r
=
|r|
. Note that a central force
can be either attractive or repulsive.
When dealing with central forces, the following formula is often helpful:
Proposition. r =
ˆ
r.
Intuitively, this is because the direction in which
r
increases most rapidly is
r, and the rate of increase is clearly 1. This can also be proved algebraically:
Proof. We know that
r
2
= x
2
1
+ x
2
2
+ x
2
3
.
Then
2r
r
x
i
= 2x
i
.
So
r
x
i
=
x
i
r
= (
ˆ
r)
i
.
Proposition. Let F = −∇V (r) be a central force. Then
F = −∇V =
dV
dr
ˆ
r,
where
ˆ
r
=
r/r
is the unit vector in the radial direction pointing away from the
origin.
Proof. Using the proof above,
(V )
i
=
V
x
i
=
dV
dr
r
x
i
=
dV
dr
(
ˆ
r)
i
Since central forces have spherical symmetry, they give rise to an additional
conserved quantity called angular momentum.
Definition (Angular momentum). The angular momentum of a particle is
L = r × p = mr ×
˙
r.
Proposition. Angular momentum is conserved by a central force.
Proof.
dL
dt
= m
˙
r ×
˙
r + mr ×
¨
r = 0 + r × F = 0.
where the last equality comes from the fact that
F
is parallel to
r
for a central
force.
In general, for a non-central force, the rate of change of angular momentum
is the torque.
Definition
(Torque)
.
The torque
G
of a particle is the rate of change of angular
momentum.
G =
dL
dt
= r × F.
Note that
L
and
G
depends on the choice of origin. For a central force, only
the angular momentum about the center of the force is conserved.
3.4 Gravity
We’ll now study an important central force gravity. This law was discovered
by Newton and was able to explain the orbits of various planets. However, we
will only study the force and potential aspects of it, and postpone the study of
orbits for a later time.
Law
(Newton’s law of gravitation)
.
If a particle of mass
M
is fixed at a origin,
then a second particle of mass m experiences a potential energy
V (r) =
GMm
r
,
where G 6.67 × 10
11
m
3
kg
1
s
2
is the gravitational constant.
The gravitational force experienced is then
F = −∇V =
GMm
r
2
ˆ
r.
Since the force is negative, particles are attracted to the origin.
The potential energy is a function of the masses of both the fixed mass
M
and the second particle
m
. However, it is useful what the fixed mass
M
does
with reference to the second particle.
Definition
(Gravitaional potential and field)
.
The gravitational potential is the
gravitational potential energy per unit mass. It is
Φ
g
(r) =
GM
r
.
Note that potential is confusingly different from potential energy.
If we have a second particle, the potential energy is given by V = mΦ
g
.
The gravitational field is the force per unit mass,
g = −∇Φ
g
=
GM
r
2
ˆ
r.
If we have many fixed masses
M
i
at points
r
i
, we can add up their gravitational
potential directly. Then the total gravitational potential is given by
Φ
g
(r) =
X
i
GM
i
|r r
i
|
.
Again, V = mΦ
g
for a particle of mass m.
An important (mathematical) result about gravitational fields is that we can
treat spherical objects as point particles. In particular,
Proposition.
The external gravitational potential of a spherically symmetric
object of mass
M
is the same as that of a point particle with the same mass at
the center of the object, i.e.
Φ
g
(r) =
GM
r
.
The proof can be found in the Vector Calculus course.
Example.
If you live on a spherical planet of mass
M
and radius
R
, and can
move only a small distance z R above the surface, then
V (r) = V (R + z)
=
GMm
R + z
=
GMm
R
1
z
R
+ ···
const. +
GMm
R
2
z
= const. + mgz,
where
g
=
GM/R
2
9.8 m s
2
for Earth. Usually we are lazy and just say that
the potential is mgz.
Example.
How fast do we need to jump to escape the gravitational pull of the
Earth? If we jump upwards with speed v from the surface, then
E = T + V =
1
2
mv
2
GMm
R
.
After escape, we must have
T
0 and
V
= 0. Since energy is conserved, we
must have E 0 from the very beginning. i.e.
v > v
esc
=
r
2GM
R
.
Inertial and gravitational mass
A careful reader would observe that “mass” appears in two unrelated equations:
F = m
i
¨
r
and
F =
GM
g
m
g
r
2
ˆ
r,
and they play totally different roles. The first is the inertial mass, which
measures the resistance to motion, while the second is the gravitational mass,
which measures its response to gravitational forces.
Conceptually, these are quite different. There is no a priori reason why these
two should be equal. However, experiment shows that they are indeed equivalent
to each other, i.e. m
i
= m
g
, with an accuracy of 10
12
or better.
This (philosophical) problem was only resolved when Einstein introduced his
general theory of relativity, which says that gravity is actually a fictitious force,
which means that the acceleration of the particle is independent of its mass.
We can further distinct the gravitational mass by “passive” and “active”, i.e.
the amount of gravitational field generated by a particle (
M
), and the amount
of gravitational force received by a particle (
m
), but they are still equal, and we
end up calling all of them “mass”.
3.5 Electromagnetism
Next we will study the laws of electromagnetism. We will only provide a
very rudimentary introduction to electromagnetism. Electromagnetism will be
examined more in-depth in the IB Electromagnetism and II Electrodynamics
courses.
As the name suggests, electromagnetism comprises two parts electricity
and magnetism. Similar to gravity, we generally imagine electromagnetism
working as follows: charges generate fields, and fields cause charges to move.
A common charged particle is the electron, which is currently believed to
be a fundamental particle. It has charge
q
e
=
1.6 × 10
19
C
. Other particles’
charges are always integer multiples of q
e
(unless you are a quark).
In electromagnetism, there are two fields the electric field
E
(
r, t
) and the
magnetic field
B
(
r, t
). Their effects on charged particles is described by the
Lorentz force law.
Law
(Lorentz force law)
.
The electromagnetic force experienced by a particle
with electric charge q is
F = q(E + v × B).
This is the first time where we see a force that depends on the velocity of
the particle. In all other forces we’ve seen, the force depends only on the field
which implicitly depends on the position only. This is weird, and seems to
violate Galilean relativity, since velocity is a relative concept that depends on
the frame of reference. It turns out that weird things happen to the
B
and
E
fields when you change the frame of reference. You will learn about these in the
IB Electromagnetism course (if you take it).
As a result, the magnetic force is not a conservative force, and it cannot be
given a (regular) potential. On the other hand, assuming that the fields are
time-independent, the electric field is conservative. We write the potential as
Φ
e
(r), and E = −∇Φ
e
.
Definition
(Electrostatic potential)
.
The electrostatic potential is a function
Φ
e
(r) such that
E = −∇Φ
e
.
While the magnetic force is not conservative in the traditional sense, it always
acts perpendicularly to the velocity. Hence it does no work. So overall, energy
is conserved under the action of the electromagnetic force.
Proposition. For time independent E(r) and B(r), the energy
E = T + V =
1
2
m|v|
2
+ qΦ
e
is conserved.
Proof.
dE
dt
= m
¨
r ·
˙
r + q(Φ
e
) ·
˙
r
= (m
¨
r qE) ·
˙
r
= (q
˙
r × B) ·
˙
r
= 0
Motion in a uniform magnetic field
Consider the particular case where there is no electric field, i.e.
E
=
0
, and that
the magnetic field is uniform throughout space. We choose our axes such that
B = (0, 0, B) is constant.
According to the Lorentz force law,
F
=
q
(
E
+
v × B
) =
qv × B
. Since
the force is always perpendicular to the velocity, we expect this to act as a
centripetal force to make the particle travel in circles.
Indeed, writing out the components of the equation of motion, we obtain
m¨x = qB ˙y (1)
m¨y = qB ˙x (2)
m¨z = 0 (3)
From (3), we see that there is uniform motion parallel to
B
, which is not
interesting. We will look at the x and y components.
There are many ways to solve this system of equations. Here we solve it
using complex numbers.
Let ζ = x + iy. Then (1) + (2)i gives
m
¨
ζ = iqB
˙
ζ.
Then the solution is
ζ = αe
t
+ β,
where
ω
=
qB/m
is the gyrofrequency, and
α
and
β
are complex integration
constants. We see that the particle goes in circles, with center β and radius α.
We can choose coordinates such that, at
t
= 0,
r
= 0 and
˙
r
= (0
, v, w
),
i.e. ζ = 0 and
˙
ζ = iv, and z = 0 and ˙z = w.
The solution is then
ζ = R(1 e
t
).
with R = v = (mv)/(qB) is the gyroradius or Larmor radius. Alternatively,
x = R(1 cos ωt)
y = R sin ωt
z = wt.
This is circular motion in the plane perpendicular to B:
(x R)
2
+ y
2
= R
2
,
combined with uniform motion parallel to B, i.e. a helix.
Alternatively, we can solve this with vector operations. Start with
m
¨
r = q
˙
r × B
Let B = Bn with |n| = 1. Then
¨
r = ω
˙
r × n,
with our gyrofrequency
ω
=
qB/m
. We integrate once, assuming
r
(0) =
0
and
˙
r(0) = v
0
.
˙
r = ωr × n + v
0
. ()
Now we project () parallel to and perpendicular to B.
First we dot () with n:
˙
r · n = v
0
· n = w = const.
We integrate again to obtain
r · n = wt.
This is the part parallel to B.
To resolve perpendicularly, write r = (r · n)n + r
, with r
· n = 0.
The perpendicular component of () gives
˙
r
= wr
× n + v
0
(v
0
· n)n.
We solve this by differentiating again to obtain
¨
r
= ω
˙
r
× n = ω
2
r
+ ωv
0
× n,
which we can solve using particular integrals and complementary functions.
Point charges
So far we’ve discussed the effects of the fields on particles. But how can we
create fields in the first place? We’ll only look at the simplest case, where a
point charge generates an electric field.
Law
(Columb’s law)
.
A particle of charge
Q
, fixed at the origin, produces an
electrostatic potential
Φ
e
=
Q
4πε
0
r
,
where ε
0
8.85 × 10
12
m
3
kg
1
s
2
C
2
.
The corresponding electric field is
E = −∇Φ
e
=
Q
4πε
0
ˆ
r
r
2
.
The resulting force on a particle of charge q is
F = qE =
Qq
4πε
0
ˆ
r
r
2
.
Definition
(Electric constant)
. ε
0
is the electric constant or vacuum permittivity
or permittivity of free space.
The form of equations here are closely analogous to those of gravity. However,
there is an important difference: charges can be positive or negative. Thus
electrostatic forces can be either attractive or repulsive, whereas gravity is always
attractive.
3.6 Friction
At an atomic level, energy is always conserved. However, in many everyday
processes, this does not appear to be the case. This is because friction tends to
take kinetic energy away from objects.
In general, we can divide friction into dry friction and fluid friction.
Dry friction
When solid objects are in contact, a normal reaction force
N
(perpendicular
to the contact surface) prevents them from interpenetrating, while a frictional
force
F
(tangential to the surface) resists relative tangential motion (sliding or
slipping).
N
F
mg
If the tangential force is small, it is insufficient to overcome friction and no
sliding occurs. We have static friction of
|F| µ
s
|N|,
where µ
s
is the coefficient of static friction.
When the external force on the object exceeds
µ
s
|N|
, sliding starts, and we
have a kinetic friction of
|F| = µ
k
|N|,
where µ
k
is the coefficient of kinetic friction.
These coefficients are measures of roughness and depend on the two surfaces
involved. For example, Teflon on Teflon has coefficient of around 0.04, while
rubber on asphalt has about 0.8, while a hypothetical perfectly smooth surface
has coefficient 0. Usually, µ
s
> µ
k
> 0.
Fluid drag
When a solid object moves through a fluid (i.e. liquid or gas), it experiences a
drag force.
There are two important regimes.
(i)
Linear drag: for small things in viscous fluids moving slowly, e.g. a single
cell organism in water, the friction is proportional to the velocity, i.e.
F = k
1
v.
where
v
is the velocity of the object relative to the fluid, and
k
1
>
0 is a
constant. This
k
1
depends on the shape of the object. For example, for a
sphere of radius R, Stoke’s Law gives
k
1
= 6πµR,
where µ is the viscosity of the fluid.
(ii)
Quadratic drag: for large objects moving rapidly in less viscous fluid, e.g.
cars or tennis balls in air, the friction is proportional to the square of the
velocity, i.e.
F = k
2
|v|
2
ˆ
v.
In either case, the object loses energy. The power exerted by the drag force is
F · v =
(
k
1
|v|
2
k
2
|v|
3
Example.
Consider a projectile moving in a uniform gravitational field and
experiencing a linear drag force.
At t = 0, we throw the projectile with velocity u from x = 0.
The equation of motion is
m
dv
dt
= mg kv.
We first solve for v, and then deduce x.
We use an integrating factor exp(
k
m
t) to obtain
d
dt
e
kt/m
v
= e
kt/m
g
e
kt/m
v =
m
k
e
kt/m
g + c
v =
m
k
g + ce
kt/m
Since v = u at t = 0, we get c = u
m
k
g. So
v =
˙
x =
m
k
g +
u
m
k
g
e
kt/m
.
Integrating once gives
x =
m
k
gt
m
k
u
m
k
g
e
kt/m
+ d.
Since x = 0 at t = 0. So
d =
m
k
u
m
k
g
.
So
x =
m
k
gt +
m
k
u
m
k
g
(1 e
kt/m
).
In component form, let x = (x, y), u = (u cos θ, u sin θ), g = (0, g). So
x =
mu
k
cos θ(1 e
kt/m
)
y =
mgt
k
+
m
k
u sin θ +
mg
k
(1 e
kt/m
).
We can characterize the strength of the drag force by the dimensionless constant
ku/(mg), with a larger constant corresponding to a larger drag force.
Effect of damping on small oscillations
We’ve previously seen that particles near a potential minimum oscillate indefi-
nitely. However, if there is friction in the system, the oscillation will damp out
and energy is continually lost. Eventually, the system comes to rest at the stable
equilibrium.
Example.
If a linear drag force is added to a harmonic oscillator, then the
equation of motion becomes
m
¨
x =
2
x k
˙
x,
where
ω
is the angular frequency of the oscillator in the absence of damping.
Rewrite as
¨
x + 2γ
˙
x + ω
2
x = 0,
where γ = k/2m > 0. Solutions are x = e
λt
, where
λ
2
+ 2γλ + ω
2
= 0,
or
λ = γ ±
p
γ
2
ω
2
.
If
γ > ω
, then the roots are real and negative. So we have exponential decay.
We call this an overdamped oscillator.
If 0
< γ < ω
, then the roots are complex with
Re
(
λ
) =
γ
. So we have
decaying oscillations. We call this an underdamped oscillator.
For details, refer to Differential Equations.
4 Orbits
The goal of this chapter is to study the orbit of a particle in a central force,
m
¨
r = −∇V (r).
While the universe is in three dimensions, the orbit is confined to a plane. This is
since the angular momentum
L
=
mr×
˙
r
is a constant vector, as we’ve previously
shown. Furthermore
L · r
= 0. Therefore, the motion takes place in a plane
passing through the origin, and perpendicular to L.
4.1 Polar coordinates in the plane
We choose our axes such that the orbital plane is
z
= 0. To describe the orbit,
we introduce polar coordinates (r, θ):
x = r cos θ, y = r sin θ.
Our object is to separate the motion of the particle into radial and angular
components. We do so by defining unit vectors in the directions of increasing
r
and increasing θ:
ˆ
r =
cos θ
sin θ
,
ˆ
θ =
sin θ
cos θ
.
x
y
r
ˆ
r
ˆ
θ
θ
These two unit vectors form an orthonormal basis. However, they are not basis
vectors in the normal sense. The directions of these basis vectors depend on
time. In particular, we have
Proposition.
d
ˆ
r
dθ
=
sin θ
cos θ
=
ˆ
θ
d
ˆ
θ
dθ
=
cos θ
sin θ
=
ˆ
r.
Often, we want the derivative with respect to time, instead of
θ
. By the
chain rule, we have
d
ˆ
r
dt
=
˙
θ
ˆ
θ,
d
ˆ
θ
dt
=
˙
θ
ˆ
r.
We can now express the position, velocity and acceleration in this new polar
basis. The position is given by
r = r
ˆ
r.
Taking the derivative gives the velocity as
˙
r = ˙r
ˆ
r + r
˙
θ
ˆ
θ.
The acceleration is then
¨
r = ¨r
ˆ
r + ˙r
˙
θ
ˆ
θ + ˙r
˙
θ
ˆ
θ + r
¨
θ
ˆ
θ r
˙
θ
2
ˆ
r
= (¨r r
˙
θ
2
)
ˆ
r + (r
¨
θ + 2 ˙r
˙
θ)
ˆ
θ.
Definition
(Radial and angular velocity)
. ˙r
is the radial velocity, and
˙
θ
is the
angular velocity.
Example
(Uniform motion in a circle)
.
If we are moving in a circle, then
˙r
= 0
and
˙
θ = ω = constant. So
˙
r =
ˆ
θ.
The speed is given by
v = |
˙
r| = r|ω| = const
and the acceleration is
¨
r =
2
ˆ
r.
Hence in order to make a particle of mass
m
move uniformly in a circle, we must
supply a centripetal force mv
2
/r towards the center.
4.2 Motion in a central force field
Now let’s put in our central force. Since V = V (r), we have
F = −∇V =
dV
dr
ˆ
r.
So Newton’s 2nd law in polar coordinates is
m(¨r r
˙
θ
2
)
ˆ
r + m(r
¨
θ + 2 ˙r
˙
θ)
ˆ
θ =
dV
dr
ˆ
r.
The θ component of this equation is
m(r
¨
θ 2 ˙r
˙
θ) = 0.
We can rewrite it as
1
r
d
dt
(mr
2
˙
θ) = 0.
Let
L
=
mr
2
˙
θ
. This is the
z
component (and the only component) of the
conserved angular momentum L:
L = mr ×
˙
r
= mr
ˆ
r × ( ˙r
ˆ
r + r
˙
θ
ˆ
θ)
= mr
2
˙
θ
ˆ
r ×
ˆ
θ
= mr
2
˙
θ
ˆ
z.
So the angular component tells us that
L
is constant, which is the conservation
of angular momentum.
However, a more convenient quantity is the angular momentum per unit
mass:
Notation
(Angular momentum per unit mass)
.
The angular momentum per
unit mass is
h =
L
m
= r
2
˙
θ = const.
Now the radial (r) component of the equation of motion is
m(¨r r
˙
θ
2
) =
dV
dr
.
We eliminate
˙
θ using r
2
˙
θ = h to obtain
m¨r =
dV
dr
+
mh
2
r
3
=
dV
eff
dr
,
where
V
eff
(r) = V (r) +
mh
2
2r
2
.
We have now reduced the problem to 1D motion in an (effective) potential as
studied previously.
The total energy of the particle is
E =
1
2
m|
˙
r|
2
+ V (r)
=
1
2
m( ˙r
2
+ r
2
˙
θ
2
) + V (r)
(since
˙
r = ˙r
ˆ
r + r
˙
θ
ˆ
θ, and
ˆ
r and
ˆ
θ are orthogonal)
=
1
2
m ˙r
2
+
mh
2
2r
2
+ V (r)
=
1
2
m ˙r
2
+ V
eff
(r).
Example.
Consider an attractive force following the inverse-square law (e.g.
gravity). Here
V =
mk
r
,
for some constant k. So
V
eff
=
mk
r
+
mh
2
2r
2
.
We have two terms of opposite signs and different dependencies on
r
. For small
r
,
the second term dominates and
V
eff
is large. For large
r
, the first term dominates.
Then V
eff
asymptotically approaches 0 from below.
r
V
eff
E
min
r
The minimum of V
eff
is at
r
=
h
2
k
, E
min
=
mk
2
2h
2
.
We have a few possible types of motion:
If
E
=
E
min
, then
r
remains at
r
and
˙
θh/r
2
is constant. So we have a
uniform motion in a circle.
If
E
min
< E <
0, then
r
oscillates and
˙r
=
h/r
2
does also. This is a
non-circular, bounded orbit.
We’ll now introduce a lot of (pointless) definitions:
Definition
(Periapsis, apoapsis and apsides)
.
The points of minimum and
maximum
r
in such an orbit are called the periapsis and apoapsis. They
are collectively known as the apsides.
Definition
(Perihelion and aphelion)
.
For an orbit around the Sun, the
periapsis and apoapsis are known as the perihelion and aphelion.
In particular
Definition
(Perigee and apogee)
.
The perihelion and aphelion of the
Earth are known as the perigee and apogee.
If
E
0, then
r
comes in from
, reaches a minimum, and returns to
infinity. This is an unbounded orbit.
We will later show that in the case of motion in an inverse square force, the
trajectories are conic sections (circles, ellipses, parabolae and hyperbolae).
Stability of circular orbits
We’ll now look at circular orbits, since circles are nice. Consider a general
potential energy V (r). We have to answer two questions:
Do circular orbits exist?
If they do, are they stable?
The conditions for existence and stability are rather straightforward. For a
circular orbit,
r
=
r
= const for some value of
h 6
= 0 (if
h
= 0, then the object
is just standing still!). Since ¨r = 0 for constant r, we require
V
0
eff
(r
) = 0.
The orbit is stable if r
is a minimum of V
eff
, i.e.
V
00
eff
(r
) > 0.
In terms of V (r), circular orbit requires
V
0
(r
) =
mh
2
r
3
and stability further requires
V
00
(r
) +
3mh
2
r
4
= V
00
(r
) +
3
r
V
0
(r
) > 0.
In terms of the radial force F (r) = V
0
(r), the orbit is stable if
F
0
(r
) +
3
r
F (r
) < 0.
Example. Consider a central force with
V (r) =
mk
r
p
for some k, p > 0. Then
V
00
(r) +
3
r
V
0
(r) =
p(p + 1) + 3p
mk
r
p+2
= p(2 p)
mk
r
p+2
.
So circular orbits are stable for
p <
2. This is illustrated by the graphs of
V
eff
(
r
)
for p = 1 and p = 3.
V
eff
p = 1
p = 3
4.3 Equation of the shape of the orbit
In general, we could determine r(t) by integrating the energy equation
E =
1
2
m ˙r
2
+ V
eff
(r)
t = ±
r
m
2
Z
dr
p
E V
eff
(r)
However, this is usually not practical, because we can’t do the integral. Instead,
it is usually much easier to find the shape r(θ) of the orbit.
Still, solving for
r
(
θ
) is also not easy. We will need a magic trick we
introduce the new variable
Notation.
u =
1
r
.
Then
˙r =
dr
dθ
˙
θ =
dr
dθ
h
r
2
= h
du
dθ
,
and
¨r =
d
dt
h
du
dθ
= h
d
2
u
dθ
2
˙
θ = h
d
2
u
dθ
2
h
r
2
= h
2
u
2
d
2
u
dθ
2
.
This doesn’t look very linear with
u
2
, but it will help linearizing the equation
when we put in other factors.
The radial equation of motion
m¨r
mh
2
r
3
= F (r)
then becomes
Proposition (Binet’s equation).
mh
2
u
2
d
2
u
dθ
2
+ u
= F
1
u
.
This still looks rather complicated, but observe that for an inverse square
force, F (1/u) is proportional to u
2
, and then the equation is linear!
In general, given an arbitrary F (r), we aim to solve this second order ODE
for u(θ). If needed, we can then work out the time-dependence via
˙
θ = hu
2
.
4.4 The Kepler problem
The Kepler problem is the study of the orbits of two objects interacting via
a central force that obeys the inverse square law. The goal is to classify the
possible orbits and study their properties. One of the most important examples
of the Kepler problem is the orbit of celestial objects, as studied by Kepler
himself.
Shapes of orbits
For a planet orbiting the sun, the potential and force are given by
V (r) =
mk
r
, F (r) =
mk
r
2
with
k
=
GM
(for the Coulomb attraction of opposite charges, we have the same
equation with k =
Qq
4πε
0
m
).
Binet’s equation then becomes linear, and
d
2
u
dθ
2
+ u =
k
h
2
.
We write the general solution as
u =
k
h
2
+ A cos(θ θ
0
),
where A 0 and θ
0
are arbitrary constants.
If
A
= 0, then
u
is constant, and the orbit is circular. Otherwise,
u
reaches a
maximum at
θ
=
θ
0
. This is the periapsis. We now re-define polar coordinates
such that the periapsis is at θ = 0. Then
Proposition. The orbit of a planet around the sun is given by
r =
`
1 + e cos θ
, ()
with
`
=
h
2
/k
and
e
=
Ah
2
/k
. This is the polar equation of a conic, with a
focus (the sun) at the origin.
Definition
(Eccentricity)
.
The dimensionless parameter
e
0 in the equation
of orbit is the eccentricity and determines the shape of the orbit.
We can rewrite (
) in Cartesian coordinates with
x
=
r cos θ
and
y
=
r sin θ
.
Then we obtain
(1 e
2
)x
2
+ 2e`x + y
2
= `
2
. ()
There are three different possibilities:
Ellipse: (0 e < 1). r is bounded by
`
1 + e
r
`
1 e
.
() can be put into the equation of an ellipse centered on (ea, 0),
(x + ea)
2
a
2
+
y
2
b
2
= 1,
where a =
`
1 e
2
and b =
`
1 e
2
a.
ae
b
a
`
O
a
and
b
are the semi-major and semi-minor axis.
`
is the semi-latus rectum.
One focus of the ellipse is at the origin. If
e
= 0, then
a
=
b
=
`
and the
ellipse is a circle.
Hyperbola: (
e >
1). For
e >
1,
r
as
θ ±α
, where
α
=
cos
1
(1
/e
)
(
π/
2
, π
). Then (
) can be put into the equation of a hyperbola centered on
(ea, 0),
(x ea)
2
a
2
y
2
b
2
= 1,
with a =
`
e
2
1
, b =
`
e
2
1
.
a
O
`
b
This corresponds to an unbound orbit that is deflected (scattered) by an
attractive force.
b
is both the semi-minor axis and the impact parameter. It is the distance
by which the planet would miss the object if there were no attractive force.
The asymptote is y =
b
a
(x ea), or
x
p
e
2
1 y = eb.
Alternatively, we can write the equation of the asymptote as
(x, y) ·
e
2
1
e
,
1
e
!
= b
or r · n = b, the equation of a line at a distance b from the origin.
Parabola: (e = 1). Then () becomes
r =
`
1 + cos θ
.
We see that
r
as
θ ±π
. (
) becomes the equation of a parabola,
y
2
= `(` 2x). The trajectory is similar to that of a hyperbola.
Energy and eccentricity
We can figure out which path a planet follows by considering its energy.
E =
1
2
m( ˙r
2
+ r
2
˙
θ
2
)
mk
r
=
1
2
mh
2
du
dθ
2
+ u
2
!
mku
Substitute u =
1
`
(1 + e cos θ) and ` =
h
2
k
, and it simplifies to
E =
mk
2`
(e
2
1),
which is independent of θ, as it must be.
Orbits are bounded for
e <
1. This corresponds to the case
E <
0. Un-
bounded orbits have
e >
1 and thus
E >
0. A parabolic orbit has
e
= 1,
E
= 0,
and is “marginally bound”.
Note that the condition
E >
0 is equivalent to
|
˙
r| >
q
2GM
r
=
v
esc
, which
means you have enough kinetic energy to escape orbit.
Kepler’s laws of planetary motion
When Kepler first studied the laws of planetary motion, he took a telescope,
observed actual planets, and came up with his famous three laws of motion. We
are now going to derive the laws with pen and paper instead.
Law
(Kepler’s first law)
.
The orbit of each planet is an ellipse with the Sun at
one focus.
Law
(Kepler’s second law)
.
The line between the planet and the sun sweeps
out equal areas in equal times.
Law
(Kepler’s third law)
.
The square of the orbital period is proportional to
the cube of the semi-major axis, or
P
2
a
3
.
We have already shown that Law 1 follows from Newtonian dynamics and
the inverse-square law of gravity. In the solar system, planets generally have very
low eccentricity (i.e. very close to circular motion), but asteroids and comets
can have very eccentric orbits. In other solar systems, even planets have have
highly eccentric orbits. As we’ve previously shown, it is also possible for the
object to have a parabolic or hyperbolic orbit. However, we tend not to call
these “planets”.
Law 2 follows simply from the conservation of angular momentum: The area
swept out by moving dθ is dA =
1
2
r
2
dθ (area of sector of circle). So
dA
dt
=
1
2
r
2
˙
θ =
h
2
= const.
and is true for any central force.
Law 3 follows from this: the total area of the ellipse is
A
=
πab
=
h
2
P
(by
the second law). But b
2
= a
2
(1 e
2
) and h
2
= k` = ka(1 e
2
). So
P
2
=
(2π)
2
a
4
(1 e
2
)
ka(1 e
2
)
=
(2π)
2
a
3
k
.
Note that the third law is very easy to prove directly for circular orbits. Since
the radius is constant, ¨r = 0. So the equations of motion give
r
˙
θ
2
=
k
r
2
So
r
3
˙
θ
2
= k
Since
˙
θ P
1
, the result follows.
4.5 Rutherford scattering
Finally, we will consider the case where the force is repulsive instead of attractive.
An important example is the Rutherford gold foil experiment, where Rutherford
bombarded atoms with alpha particles, and the alpha particles are repelled by
the nucleus of the atom.
Under a repulsive force, the potential and force are given by
V (r) = +
mk
r
, F (r) = +
mk
r
2
.
For Coulomb repulsion of like charges,
k =
Qq
4πε
0
m
> 0.
The solution is now
u =
k
h
2
+ A cos(θ θ
0
).
wlog, we can take A 0, θ
0
= 0. Then
r =
`
e cos θ 1
with
` =
h
2
k
, e =
Ah
2
k
.
We know that
r
and
`
are positive. So we must have
e
1. Then
r
as
θ ±α, where α = cos
1
(1/e).
The orbit is a hyperbola, again given by
(x ea)
2
a
2
y
2
b
2
= 1,
with
a
=
`
e
2
1
and
b
=
`
e
2
1
. However, this time, the trajectory is the other
branch of the hyperbola.
O
b
2α
β
It seems as if the particle is deflected by O.
We can characterize the path of the particle by the impact parameter
b
and
the incident speed
v
(i.e. the speed when far away from the origin). We know
that the angular momentum per unit mass is
h
=
bv
(velocity
×
perpendicular
distance to O).
How does the scattering angle
β
=
π
2
α
depend on the impact parameter
b and the incident speed v?
Recall that the angle α is given by α = cos
1
(1/e). So we obtain
1
e
= cos α = cos
π
2
β
2
= sin
β
2
,
So
b =
`
e
2
1
=
(bv)
2
k
tan
β
2
.
So
β = 2 tan
1
k
bv
2
.
We see that if we have a small impact parameter, i.e.
b k/v
2
, then we can
have a scattering angle approaching π.
5 Rotating frames
Recall that Newton’s laws hold only in inertial frames. However, sometimes, our
frames are not inertial. In this chapter, we are going to study a particular kind
of non-inertial frame a rotating frame. An important rotating frame is the
Earth itself, but there are also other examples such as merry-go-rounds.
5.1 Motion in rotating frames
Now suppose that
S
is an inertial frame, and
S
0
is rotating about the
z
axis
with angular velocity ω =
˙
θ with respect to S.
Definition
(Angular velocity vector)
.
The angular velocity vector of a rotating
frame is ω = ω
ˆ
z, where
ˆ
z is the axis of rotation and ω is the angular speed.
First we wish to relate the basis vectors
{e
i
}
and
{e
0
i
}
of
S
and
S
0
respectively.
Consider a particle at rest in S
0
. From the perspective of S, its velocity is
dr
dt
S
= ω × r,
where
ω
=
ω
ˆ
z
is the angular velocity vector (aligned with the rotation axis).
This formula also applies to the basis vectors of S
0
.
de
0
i
dt
S
= ω × e
0
i
.
Now given a general time-dependent vector
a
, we can express it in the
{e
0
i
}
basis
as follows:
a =
X
a
0
i
(t)e
0
i
.
From the perspective of
S
0
,
e
0
i
is constant and the time derivative of
a
is given
by
da
dt
S
0
=
X
da
0
i
dt
e
0
i
.
In
S
, however,
e
0
i
is not constant. So we apply the product rule to obtain the
time derivative of a:
da
dt
S
=
X
da
i
dt
e
0
i
+
X
a
0
i
ω × e
0
i
=
da
dt
S
0
+ ω × a.
This key identity applies to all vectors and can be written as an operator equation:
Proposition.
If
S
is an inertial frame, and
S
0
is rotating relative to
S
with
angular velocity ω, then
d
dt
S
=
d
dt
S
0
+ ω × .
Applied to the position vector r(t) of a particle, it gives
dr
dt
S
=
dr
dt
S
0
+ ω × r.
We can interpret this as saying that the difference in velocity measured in
the two frames is the relative velocity of the frames.
We apply this formula a second time, and allow
ω
to depend on time. Then
we have
d
2
r
dt
2
S
=

d
dt
S
0
+ ω×

dr
dt
S
0
+ ω × r
.
=
d
2
r
dt
2
S
0
+ 2ω ×
dr
dt
S
0
+
˙
ω × r + ω × (ω ×r)
Since S is inertial, Newton’s Second Law is
m
d
2
r
dt
2
S
= F.
So
Proposition.
m
d
2
r
dt
2
S
0
= F 2mω ×
dr
dt
S
0
m
˙
ω × r mω × (ω ×r).
Definition
(Fictious forces)
.
The additional terms on the RHS of the equation
of motion in rotating frames are fictitious forces, and are needed to explain the
motion observed in S
0
. They are
Coriolis force: 2mω ×
dr
dt
S
0
.
Euler force: m
˙
ω × r
Centrifugal force: mω ×(ω × r).
In most cases, ω is constant and can neglect the Euler force.
5.2 The centrifugal force
What exactly does the centrifugal force do? Let ω = ω
ˆ
ω, where |
ˆ
ω| = 1. Then
mω × (ω × r) = m
(ω · r)ω (ω · ω)r
=
2
r
,
where
r
=
r
(
r·
ˆ
ω
)
ˆ
ω
is the projection of the position on the plane perpendicular
to
ω
. So the centrifugal force is directed away from the axis of rotation, and its
magnitude is
2
times the distance form the axis.
ω
r
r
Note that
r
· r
= r · r (r ·
ˆ
ω)
2
(|r
|
2
) = 2r 2(r ·
ˆ
ω)
ˆ
ω = 2r
.
So
mω × (ω × r) = −∇
1
2
2
|r
|
2
= −∇
1
2
m|ω × r|
2
.
Thus the centrifugal force is a conservative (fictitious) force.
On a rotating planet, the gravitational and centrifugal forces per unit mass
combine to make the effective gravity,
g
eff
= g + ω
2
r
.
This gravity will not be vertically downwards. Consider a point
P
at latitude
λ
on the surface of a spherical planet of radius R.
We construct orthogonal axes:
O
ω
P
ˆ
z
ˆ
y
λ
with
ˆ
x into the page. So
ˆ
z is up,
ˆ
y is North, and
ˆ
x is East.
At P , we have
r = R
ˆ
z
g = g
ˆ
z
ω = ω(cos λ
ˆ
y + sin λ
ˆ
z)
So
g
eff
= g + ω
2
r
= g
ˆ
z + ω
2
R cos λ(cos λ
ˆ
z sin λ
ˆ
y)
= ω
2
R cos λ sin λ
ˆ
y (g ω
2
R cos
2
λ)
ˆ
z.
So the angle α between g and g
eff
is given by
tan α =
ω
2
R cos λ sin λ
g ω
2
R cos
2
λ
.
This is 0 at the equator and the poles, and greatest when you are halfway
between. However, this is still tiny on Earth, and does not affect our daily life.
5.3 The Coriolis force
The Coriolis force is a more subtle force. Writing
v
=
dr
dt
S
0
, we can write the
force as
F = 2mω × v.
Note that this has the same form as the Lorentz force caused by a magnetic
field, and is velocity-dependent. However, unlike the effects of a magnetic field,
particles do not go around in circles in a rotating frame, since we also have the
centrifugal force in play.
Since this force is always perpendicular to the velocity, it does no work.
Consider motion parallel to the Earth’s surface. We only care about the
effect of the Coriolis force on the horizontal trajectory, and ignore the vertical
component that is tiny compared to gravity.
So we only take the vertical component of
ω
,
ω sin λ
ˆ
z
. The horizontal velocity
v = v
x
ˆ
x + v
y
ˆ
y generates a horizontal Coriolis force:
2 sin λ
ˆ
z × v = 2 sin λ(v
y
ˆ
x v
x
ˆ
y).
In the Northern hemisphere (0
< λ < π/
2), this causes a deflection towards
the right. In the Southern Hemisphere, the deflection is to the left. The effect
vanishes at the equator.
Note that only the horizontal effect of horizontal motion vanishes at the
equator. The vertical effects or those caused by vertical motion still exist.
Example.
Suppose a ball is dropped from a tower of height
h
at the equator.
Where does it land?
In the rotating frame,
¨
r = g 2ω ×
˙
r ω × (ω × r).
We work to first order in ω. Then
¨
r = g 2ω ×
˙
r + O(ω
2
).
Integrate wrt t to obtain
˙
r = gt 2ω × (r r
0
) + O(ω
2
),
where
r
0
is the initial position. We substitute into the original equation to obtain
¨
r = g 2ω × gt + O(ω
2
).
(where some new
ω
2
terms are thrown into
O
(
ω
2
)). We integrate twice to obtain
r = r
0
+
1
2
gt
2
1
3
ω × gt
3
+ O(ω
2
).
In components, we have g = (0, 0, g), ω = (0, ω, 0) and r
0
= (0, 0, R + h). So
r =
1
3
ωgt
3
, 0, R + h
1
2
gt
2
+ O(ω
2
).
So the particle hits the ground at
t
=
p
2h/g
, and its eastward displacement is
1
3
wg
2h
g
3/2
.
This can be understood in terms of angular momentum conservation in the
non-rotating frame. At the beginning, the particle has the same angular velocity
with the Earth. As it falls towards the Earth, to maintain the same angular
momentum, the angular velocity has to increase to compensate for the decreased
radius. So it spins faster than the Earth and drifts towards the East, relative to
the Earth.
Example.
Consider a pendulum that is free to swing in any plane, e.g. a weight
on a string. At the North pole, it will swing in a plane that is fixed in an inertial
frame, while the Earth rotates beneath it. From the perspective of the rotating
frame, the plane of the pendulum rotates backwards. This can be explained as a
result of the Coriolis force.
In general, at latitude λ, the plane rotates rightwards with period
1 day
sin λ
.
6 Systems of particles
Now suppose we have
N
interacting particles. We adopt the following notation:
particle
i
has mass
m
i
, position
r
i
, and momentum
p
i
=
m
i
˙
r
i
. Note that the
subscript denotes which particle it is referring to, not vector components.
Newton’s Second Law for particle i is
m
i
¨
r
i
=
˙
p
i
= F
i
,
where F
i
is the total force acting on particle i. We can write F
i
as
F
i
= F
ext
i
+
N
X
j=1
F
ij
,
where
F
ij
is the force on particle
i
by particle
j
, and
F
ext
i
is the external force
on i, which comes from particles outside the system.
Since a particle cannot exert a force on itself, we have
F
ii
=
0
. Also, Newton’s
third law requires that
F
ij
= F
ji
.
For example, if the particles interact only via gravity, then we have
F
ij
=
Gm
i
m
j
(r
i
r
j
)
|r
i
r
j
|
3
= F
ji
.
6.1 Motion of the center of mass
Sometimes, we are interested in the collection of particles as a whole. For
example, if we treat a cat as a collection of particles, we are more interested in
how the cat as a whole falls, instead of tracking the individual particles of the
cat.
Hence, we define some aggregate quantities of the system such as the total
mass and investigate how these quantities relate.
Definition (Total mass). The total mass of the system is M =
P
m
i
.
Definition (Center of mass). The center of mass is located at
R =
1
M
N
X
i=1
m
i
r
i
.
This is the mass-weighted average position.
Definition (Total linear momentum). The total linear momentum is
P =
N
X
i=1
p
i
=
N
X
i=1
m
i
˙
r
i
= M
˙
R.
Note that this is equivalent to the momentum of a single particle of mass
M
at
the center of mass.
Definition (Total external force). The total external force is
F =
N
X
i=1
F
ext
i
.
We can now obtain the equation of motion of the center of mass:
Proposition.
M
¨
R = F.
Proof.
M
¨
R =
˙
P
=
N
X
i=1
˙
p
i
=
N
X
i=1
F
ext
i
+
N
X
i=1
N
X
j=1
F
ij
= F +
1
2
X
i
X
j
(F
ij
+ F
ji
)
= F
This means that if we don’t care about the internal structure, we can treat
the system as a point particle of mass
M
at the center of mass
R
. This is why
Newton’s Laws apply to macroscopic objects even though they are not individual
particles.
Law
(Conservation of momentum)
.
If there is no external force, i.e.
F
=
0
, then
˙
P = 0. So the total momentum is conserved.
If there is no external force, then the center of mass moves uniformly in a
straight line. In this case, we can pick a particularly nice frame of reference,
known as the center of mass frame.
Definition
(Center of mass frame)
.
The center of mass frame is an inertial
frame in which R = 0 for all time.
Doing calculations in the center of mass frame is usually much more convenient
than using other frames,
After doing linear motion, we can now look at angular motion.
Definition
(Total angular momentum)
.
The total angular momentum of the
system about the origin is
L =
X
i
r
i
× p
i
.
How does the total angular momentum change with time? Here we have to
assume a stronger version of Newton’s Third law, saying that
F
ij
= F
ji
and is parallel to (r
i
r
j
).
This is true, at least, for gravitational and electrostatic forces.
Then we have
˙
L =
X
i
r
i
×
˙
p
i
+
˙
r
i
× p
i
=
X
i
r
i
×
F
ext
i
+
X
j
F
ij
+ m(
˙
r
i
×
˙
r
i
)
=
X
i
r
i
× F
ext
i
+
X
i
X
j
r
i
× F
ij
=
X
i
G
ext
i
+
1
2
X
i
X
j
(r
i
× F
ij
+ r
j
× F
ji
)
= G +
1
2
X
i
X
j
(r
i
r
j
) × F
ij
= G,
where
Definition (Total external torque). The total external torque is
G =
X
i
r
i
× F
ext
i
.
So the total angular momentum is conserved if
G
=
0
, ie the total external
torque vanishes.
6.2 Motion relative to the center of mass
So far, we have shown that externally, a multi-particle system behaves as if it
were a point particle at the center of mass. But internally, what happens to the
individual particles themselves?
We write
r
i
=
R
+
r
c
i
, where
r
c
i
is the position of particle
i
relative to the
center of mass.
We first obtain two useful equalities:
X
i
m
i
r
c
i
=
X
m
i
r
i
X
m
i
R = MR MR = 0.
Differentiating gives
X
i
m
i
˙
r
c
i
= 0.
Using these equalities, we can express the angular momentum and kinetic energy
in terms of R and r
c
i
only:
L =
X
i
m
i
(R + r
c
i
) × (
˙
R +
˙
r
c
i
)
=
X
i
m
i
R ×
˙
R + R ×
X
i
m
i
˙
r
c
i
+
X
i
m
i
r
c
i
×
˙
R +
X
i
m
i
r
c
i
×
˙
r
c
i
= MR ×
˙
R +
X
i
m
i
r
c
i
×
˙
r
c
i
T =
1
2
X
i
m
i
|
˙
r
i
|
2
=
1
2
X
i
m
I
(
˙
R +
˙
r
i
c
) · (
˙
R +
˙
r
c
i
)
=
1
2
X
i
m
i
˙
R ·
˙
R +
˙
R ·
X
i
m
i
˙
r
c
i
+
1
2
X
i
m
i
˙
r
c
i
·
˙
r
c
i
=
1
2
M|
˙
R|
2
+
1
2
X
i
m
i
|
˙
r
c
i
|
2
We see that each item is composed of two parts that of the center of mass
and that of motion relative to center of mass.
If the forces are conservative in the sense that
F
ext
i
= −∇
i
V
i
(r
i
),
and
F
ij
= −∇
i
V
ij
(r
i
r
j
),
where
i
is the gradient with respect to
r
i
, then energy is conserved in the from
E = T +
X
i
V
i
(r
i
) +
1
2
X
i
X
j
V
ij
(r
i
r
j
) = const.
6.3 The two-body problem
The two-body problem is to determine the motion of two bodies interacting only
via gravitational forces.
The center of mass is at
R =
1
M
(m
1
r
1
+ m
2
r
2
),
where M = m
1
+ m
2
.
The magic trick to solving the two-body problem is to define the separation
vector (or relative position vector)
r = r
1
r
2
.
Then we write everything in terms of R and r.
r
1
= R +
m
2
M
r, r
2
= R
m
1
M
r.
r
2
r
1
R
r
Since the external force
F
=
0
, we have
¨
R
=
0
, i.e. the center of mass moves
uniformly.
Meanwhile,
¨
r =
¨
r
1
¨
r
2
=
1
m
1
F
12
1
m
2
F
21
=
1
m
1
+
1
m
2
F
12
We can write this as
µ
¨
r = F
12
(r),
where
µ =
m
1
m
2
m
1
+ m
2
is the reduced mass. This is the same as the equation of motion for one particle of
mass
µ
with position vector
r
relative to a fixed origin as studied previously.
For example, with gravity,
µ
¨
r =
Gm
1
m
2
ˆ
r
|r|
2
.
So
¨
r =
GM
ˆ
r
|r|
2
.
For example, give a planet orbiting the Sun, both the planet and the sun moves
in ellipses about their center of mass. The orbital period depends on the total
mass.
It can be shown that
L = MR ×
˙
R + µr ×
˙
r
T =
1
2
M|
˙
R|
2
+
1
2
µ|
˙
r|
2
by expressing r
1
and r
2
in terms of r and R.
6.4 Variable-mass problem
All systems we’ve studied so far have fixed mass. However, in real life, many
objects have changing mass, such as rockets, fireworks, falling raindrops and
rolling snowballs.
Again, we will use Newton’s second law, which states that
dp
dt
= F, with p = m
˙
r.
We will consider a rocket moving in one dimension with mass
m
(
t
) and velocity
v
(
t
). The rocket propels itself forwards by burning fuel and ejecting the exhaust
at velocity u relative to the rocket.
At time t, the rocket looks like this:
m(t)
v(t)
At time
t
+
δt
, it ejects exhaust of mass
m
(
t
)
m
(
t
+
δt
) with velocity
v
(
t
)
u + O(δt).
m(t)
v(t)
m
v(t) u
The change in total momentum of the system (rocket + exhaust) is
δp = m(t + δt)v(t + δt) + [m(t) m(t + δt)][v(t) u(t) + O(δt)] m(t)v(t)
= (m + ˙t + O(δt
2
))(v + ˙vδt + O(δt
2
)) ˙t(v u) + O(δt
2
) mv
= ( ˙mv + m ˙v ˙mv + ˙mu)δt + O(δt
2
)
= (m ˙v + ˙mu)δt + O(δt
2
).
Newton’s second law gives
lim
δ0
δp
δt
= F
where F is the external force on the rocket. So we obtain
Proposition (Rocket equation).
m
dv
dt
+ u
dm
dt
= F.
Example.
Suppose that we travel in space with
F
= 0. Assume also that
u
is
constant. Then we have
m
dv
dt
= u
dm
dt
.
So
v = v
0
+ u log
m
0
m(t)
,
Note that we are expressing things in terms of the mass remaining
m
, not time
t
.
Note also that the velocity does not depend on the rate at which mass is
ejected, only the velocity at which it is ejected. Of course, if we expressed
v
as a
function of time, then the velocity at a specific time does depend on the rate at
which mass is ejected.
Example.
Consider a falling raindrop of mass
m
(
t
), gathering mass from a
stationary cloud. In this case, u = v. So
m
dv
dt
+ v
dm
dt
=
d
dt
(mv) = mg,
with
v
measured downwards. To obtain a solution of this, we will need a model
to determine the rate at which the raindrop gathers mass.
7 Rigid bodies
This chapter is somewhat similar to the previous chapter. We again have a
lot of particles and we study their motion. However, instead of having forces
between the individual particles, this time the particles are constrained such
that their relative positions are fixed. This corresponds to a solid object that
cannot deform. We call these rigid bodies.
Definition
(Rigid body)
.
A rigid body is an extended object, consisting of
N
particles that are constrained such that the distance between any pair of
particles, |r
i
r
j
|, is fixed.
The possible motions of a rigid body are the continuous isometries of Euclidean
space, i.e. translations and rotations. However, as we have previously shown,
pure translations of rigid bodies are uninteresting they simply correspond to
the center of mass moving under an external force. Hence we will first study
rotations.
Later, we will combine rotational and translational effects and see what
happens.
7.1 Angular velocity
We’ll first consider the cases where there is just one particle, moving in a circle
of radius s about the z axis. Its position and velocity vectors are
r = (s cos θ, s sin θ, z)
˙
r = (s
˙
θ sin θ, s
˙
θ cos θ, 0).
We can write
˙
r = ω × r,
where
ω =
˙
θ
ˆ
z
is the angular velocity vector.
In general, we write
ω =
˙
θ
ˆ
n = ω
ˆ
n,
where
ˆ
n is a unit vector parallel to the rotation axis.
The kinetic energy of this particle is thus
T =
1
2
m|
˙
r|
2
=
1
2
ms
2
˙
θ
2
=
1
2
Iω
2
where
I
=
ms
2
is the moment of inertia. This is the counterpart of “mass” in
rotational motion.
Definition (Moment of inertia). The moment of inertia of a particle is
I = ms
2
= m|
ˆ
n × r|
2
,
where s is the distance of the particle from the axis of rotation.
7.2 Moment of inertia
In general, consider a rigid body in which all
N
particles rotate about the same
axis with the same angular velocity:
˙
r
i
= ω × r
i
.
This ensures that
d
dt
|r
i
r
j
|
2
= 2(
˙
r
i
˙
r
j
) · (r
i
r
j
) = 2
ω × (r
i
r
j
)
· (r
i
r
j
) = 0,
as required for a rigid body.
Similar to what we had above, the rotational kinetic energy is
T =
1
2
N
X
i=1
m
i
|
˙
r
i
|
2
=
1
2
Iω
2
,
where
Definition
(Moment of inertia)
.
The moment of inertia of a rigid body about
the rotation axis
ˆ
n is
I =
N
X
i=1
m
i
s
2
i
=
N
X
i=1
m
i
|
ˆ
n × r
i
|
2
.
Again, we define the angular momentum of the system:
Definition. The angular momentum is
L =
X
i
m
i
r
i
×
˙
r
i
=
X
i
m
i
r
i
× (ω × r
i
).
Note that our definitions for angular motion are analogous to those for linear
motion. The moment of inertia
I
is defined such that
T
=
1
2
Iω
2
. Ideally, we
would want the momentum to be
L
=
Iω
. However, this not true. In fact,
L
need not be parallel to ω.
What is true, is that the component of
L
parallel to
ω
is equal to
Iω
. Write
ω = ω
ˆ
n. Then we have
L ·
ˆ
n = ω
X
i
m
i
ˆ
n · (r
i
× (
ˆ
n × r
i
))
= ω
X
i
m(
ˆ
n × r
i
) · (
ˆ
n × r
i
)
= Iω.
What does L itself look like? Using vector identities, we have
L =
X
i
m
i
(r
i
· r
i
)ω (r
i
· ω)r
i
Note that this is a linear function of ω. So we can write
L = Iω,
where we abuse notation to use
I
for the inertia tensor. This is represented by a
symmetric matrix with components
I
jk
=
X
i
m
i
(|r
i
|
2
δ
jk
(r
i
)
j
(r
i
)
k
),
where i refers to the index of the particle, and j, k are dummy suffixes.
If the body rotates about a principal axis, i.e. one of the three orthogonal
eigenvectors of
I
, then
L
will be parallel to
ω
. Usually, the principal axes lie on
the axes of rotational symmetry of the body.
7.3 Calculating the moment of inertia
For a solid body, we usually want to think of it as a continuous substance with
a mass density, instead of individual point particles. So we replace the sum of
particles by a volume integral weighted by the mass density ρ(r).
Definition (Mass, center of mass and moment of inertia). The mass is
M =
Z
ρ dV.
The center of mass is
R =
1
M
Z
ρr dV
The moment of inertia is
I =
Z
ρs
2
dV =
Z
ρ|
ˆ
n × r|
2
dV.
In theory, we can study inhomogeneous bodies with varying
ρ
, but usually
we mainly consider homogeneous ones with constant ρ throughout.
Example
(Thin circular ring)
.
Suppose the ring has mass
M
and radius
a
, and
a rotation axis through the center, perpendicular to the plane of the ring.
a
Then the moment of inertia is
I = M a
2
.
Example
(Thin rod)
.
Suppose a rod has mass
M
and length
`
. It rotates
through one end, perpendicular to the rod.
`
The mass per unit length is M/`. So the moment of inertia is
I =
Z
`
0
M
`
x
2
dx =
1
3
M`
2
.
Example
(Thin disc)
.
Consider a disc of mass
M
and radius
a
, with a rotation
axis through the center, perpendicular to the plane of the disc.
a
Then
I =
Z
2π
0
Z
a
0
M
πa
2
|{z}
mass per unit length
r
2
|{z}
s
2
r dr dθ
| {z }
area element
=
M
πa
2
Z
a
0
r
3
dr
Z
2π
0
dθ
=
M
πa
2
1
4
a
4
(2π)
=
1
2
Ma
2
.
Now suppose that the rotation axis is in the plane of the disc instead (also
rotating through the center). Then
I =
Z
2π
0
Z
a
0
M
πa
2
|{z}
mass per unit length
(r sin θ)
2
| {z }
s
2
r dr dθ
| {z }
area element
=
M
πa
2
Z
a
0
r
3
dr
Z
2π
0
sin
2
θ dθ
=
M
πa
2
1
4
a
4
π
=
1
4
Ma
2
.
Example.
Consider a solid sphere with mass
M
, radius
a
, with a rotation axis
though the center.
a
Using spherical polar coordinates (r, θ, φ) based on the rotation axis,
I =
Z
2π
0
Z
π
0
Z
a
0
M
4
3
πa
3
|{z}
ρ
(r sin θ)
2
| {z }
s
2
r
2
sin θ dr dθ dφ
| {z }
volume element
=
M
4
3
πa
3
Z
a
0
r
4
dr
Z
π
0
(1 cos
2
) sin θ dθ
Z
2π
0
dφ
=
M
4
3
πa
3
·
1
5
a
5
·
4
3
· 2π
=
2
5
Ma
2
.
Usually, finding the moment of inertia involves doing complicated integrals.
We will now come up with two theorems that help us find moments of inertia.
Theorem (Perpendicular axis theorem). For a two-dimensional object (a lam-
ina), and three perpendicular axes
x, y, z
through the same spot, with
z
normal
to the plane,
I
z
= I
x
+ I
y
,
where I
z
is the moment of inertia about the z axis.
x
y
z
Note that this does not apply to 3D objects! For example, in a sphere,
I
x
= I
y
= I
z
.
Proof. Let ρ be the mass per unit volume. Then
I
x
=
Z
ρy
2
dA
I
y
=
Z
ρx
2
dA
I
z
=
Z
ρ(x
2
+ y
2
) dA = I
x
+ I
y
.
Example. For a disc, I
x
= I
y
by symmetry. So I
z
= 2I
x
.
Theorem
(Parallel axis theorem)
.
If a rigid body of mass
M
has moment of
inertia
I
C
about an axis passing through the center of mass, then its moment of
inertia about a parallel axis a distance d away is
I = I
C
+ Md
2
.
CM
d
Proof.
With a convenient choice of Cartesian coordinates such that the center of
mass is at the origin and the two rotation axes are x = y = 0 and x = d, y = 0,
I
C
=
Z
ρ(x
2
+ y
2
) dV,
and
Z
ρr dV = 0.
So
I =
Z
ρ((x d)
2
+ y
2
) dV
=
Z
ρ(x
2
+ y
2
) dV 2d
Z
ρx dV +
Z
d
2
ρ dV
= I
c
+ 0 + M d
2
= I
c
+ Md
2
.
Example.
Take a disc of mass
M
and radius
a
, and rotation axis through a
point on the circumference, perpendicular to the plane of the disc. Then
a
I = I
c
+ Ma
2
=
1
2
Ma
2
+ Ma
2
=
3
2
Ma
2
.
7.4 Motion of a rigid body
The general motion of a rigid body can be described as a translation of its centre
of mass, following a trajectory
R
(
t
), together with a rotation about an axis
through the center of mass. As before, we write
r
i
= R + r
c
i
.
Then
˙
r
i
=
˙
R +
˙
r
c
i
.
Using this, we can break down the velocity and kinetic energy into translational
and rotational parts.
If the body rotates with angular velocity ω about the center of mass, then
˙
r
c
i
= ω × r
c
i
.
Since r
c
i
= r
i
R, we have
˙
r
i
=
˙
R + ω × r
c
i
=
˙
R + ω × (r
i
R).
On the other hand, the kinetic energy, as calculated in previous lectures, is
T =
1
2
M|
˙
R|
2
+
1
2
X
i
m
i
|
˙
r
c
i
|
2
=
1
2
M|
˙
R|
2
| {z }
translational KE
+
1
2
I
c
ω
2
| {z }
rotational KE
.
Sometimes we do not want to use the center of mass as the center. For example,
if an item is held at the edge and spun around, we’d like to study the motion
about the point at which the item is held, and not the center of mass.
So consider any point
Q
, with position vector
Q
(
t
) that is not the center of
mass but moves with the rigid body, i.e.
˙
Q =
˙
R + ω × (Q R).
Usually this is a point inside the object itself, but we do not assume that in our
calculation.
Then we can write
˙
r
i
=
˙
R + ω × (r
i
R)
=
˙
Q ω × (Q R) + ω × (r
i
R)
=
˙
Q + ω × (r
i
Q).
Therefore the motion can be considered as a translation of
Q
(with different
velocity than the center of mass), together with rotation about
Q
(with the same
angular velocity ω).
Equations of motion
As shown previously, the linear and angular momenta evolve according to
˙
P = F (total external force)
˙
L = G (total external torque)
These two equations determine the translational and rotational motion of a rigid
body.
L
and
G
depend on the choice of origin, which could be any point that is
fixed in an inertial frame. More surprisingly, it can also be applied to the center
of mass, even if this is accelerated: If
m
i
¨
r
i
= F
i
,
then
m
i
¨
r
c
i
= F
i
+ m
i
¨
R.
So there is a fictitious force
m
i
¨
R
in the center-of-mass frame. But the total
torque of the fictitious forces about the center of mass is
X
i
r
c
i
×
m
i
¨
R
=
X
m
i
r
c
i
×
¨
R = 0 ×
˙
R = 0.
So we can still apply the above two equations.
In summary, the laws of motion apply in any inertial frame, or the center of
mass (possibly non-inertial) frame.
Motion in a uniform gravitational field
In a uniform gravitational field
g
, the total gravitational force and torque are
the same as those that would act on a single particle of mass
M
located at the
center of mass (which is also the center of gravity):
F =
X
i
F
ext
i
=
X
i
m
i
g = Mg,
and
G =
X
i
G
ext
i
=
X
i
r
i
× (m
i
g) =
X
m
i
r
i
× g = M R × g.
In particular, the gravitational torque about the center of mass vanishes:
G
c
=
0
.
We obtain a similar result for gravitational potential energy.
The gravitational potential in a uniform g is
Φ
g
= r · g.
(since g = −∇Φ
g
by definition)
So
V
ext
=
X
i
V
ext
i
=
X
i
m
i
(r
i
· g)
= M(R · g).
Example
(Thrown stick)
.
Suppose we throw a symmetrical stick. So the center
of mass is the actual center. Then the center of the stick moves in a parabola.
Meanwhile, the stick rotates with constant angular velocity about its center due
to the absence of torque.
Example. Swinging bar.
`
Mg
This is an example of a compound pendulum.
Consider the bar to be rotating about the pivot (and not translating). Its
angular momentum is
L
=
I
˙
θ
with
I
=
1
3
M`
2
. The gravitational torque about
the pivot is
G = Mg
`
2
sin θ.
The equation of motion is
˙
L = G.
So
I
¨
θ = M g
`
2
sin θ,
or
¨
θ =
3g
2`
sin θ.
which is exactly equivalent to a simple pendulum of length 2
`/
3, with angular
frequency
q
3g
2`
.
This can also be obtained from an energy argument:
E = T + V =
1
2
I
˙
θ
2
Mg
`
2
cos θ.
We differentiate to obtain
dE
dt
=
˙
θ(I
¨
θ + Mg
`
2
sin θ) = 0.
So
I
¨
θ = M g
`
2
sin θ.
Sliding versus rolling
Consider a cylinder or sphere of radius
a
, moving along a stationary horizontal
surface.
a
v
ω
In general, the motion consists of a translation of the center of mass (with
velocity v) plus a rotation about the center of mass (with angular velocity ω).
The horizontal velocity at the point of contact is
v
slip
= v .
For a pure sliding motion,
v 6
= 0 and
ω
= 0, in which case
v 6
= 0: the point
of contact moves relative to the surface and kinetic friction may occur.
For a pure rolling motion,
v 6
= 0 and
ω 6
= 0 such that
v
= 0: the point
of contact is stationary. This is the no-slip condition.
The rolling body can alternatively be considered to be rotating instanta-
neously about the point of contact (with angular velocity
ω
) and not translating.
Example (Rolling down hill).
a
v
ω
α
Consider a cylinder or sphere of mass
M
and radius
a
rolling down a rough plane
inclined at angle α. The no-slip (rolling) condition is
v = 0.
The kinetic energy is
T =
1
2
Mv
2
+
1
2
Iω
2
=
1
2
M +
I
a
2
v
2
.
The total energy is
E =
1
2
M +
I
a
2
˙x
2
Mgx sin α,
where
x
is the distance down slope. While there is a frictional force, the
instantaneous velocity is 0, and no work is done. So energy is conserved, and we
have
dE
dt
= ˙x

M +
I
a
2
¨x M g sin α
= 0.
So
M +
I
a
2
¨x = Mg sin α.
For example, if we have a uniform solid cylinder,
I =
1
2
Ma
2
(as for a disc)
and so
¨x =
2
3
g sin α.
For a thin cylindrical shell,
I = M a
2
.
So
¨x =
1
2
g sin α.
Alternatively, we may do it in terms of forces and torques,
v
α
N
F
Mg
The equations of motion are
M ˙v = Mg sin α F
and
I ˙ω = aF.
While rolling,
˙v a ˙ω = 0.
So
M ˙v = Mg sin α
I
a
2
˙v,
leading to the same result.
Note that even though there is a frictional force, it does no work, since
v
slip
= 0. So energy is still conserved.
Example (Snooker ball).
a
N
Mg
F
ω
It is struck centrally so as to initiate translation, but not rotation. Sliding occurs
initially. Intuitively, we think it will start to roll, and we’ll see that’s the case.
The constant frictional force is
F = µ
k
N = µ
k
Mg,
which applies while v > 0.
The moment of inertia about the center of mass is
I =
2
5
Ma
2
.
The equations of motion are
M ˙v = F
I ˙ω = aF
Initially, v = v
0
and ω = 0. Then the solution is
v = v
0
µ
k
gt
ω =
5
2
µ
k
g
a
t
as long as v > 0. The slip velocity is
v
slip
= v = v
0
7
2
µ
k
gt = v
0
1
t
t
roll
,
where
t
roll
=
2v
0
7µ
k
g
.
This is valid up till
t
=
t
roll
. Then the slip velocity is 0, rolling begins and
friction ceases.
At this point,
v
=
=
5
7
v
0
. The energy is then
5
14
Mv
2
0
<
1
2
Mv
2
0
. So energy
is lost to friction.
8 Special relativity
When particles move Extremely Fast
TM
, Newtonian Dynamics becomes inaccu-
rate and is replaced by Einstein’s Special Theory of Relativity (1905).
Its effects are noticeable only when particles approach to the speed of light,
c = 299 792 458 m s
1
3 × 10
8
m s
1
This is really fast.
The Special Theory of Relativity rests on the following postulate:
The laws of physics are the same in all inertial frames
This is the principle of relativity familiar to Galileo. Galilean relativity mentioned
in the first chapter satisfies this postulate for dynamics. People then thought
that Galilean relativity is what the world obeys. However, it turns out that
there is a whole family of solutions that satisfy the postulate (for dynamics),
and Galilean relativity is just one of them.
This is not a problem (yet), since Galilean relativity seems so intuitive, and
we might as well take it to be the true one. However, it turns out that solving
Maxwell’s equations of electromagnetism gives an explicit value of the speed of
light,
c
. This is independent of the frame of reference. So the speed of light must
be the same in every inertial frame.
This is not compatible with Galilean relativity.
Consider the two inertial frames
S
and
S
0
, moving with relative velocity
v
.
Then if light has velocity
c
in
S
, then Galilean relativity predicts it has velocity
c v in S
0
, which is wrong.
Therefore, we need to find a different solution to the principle of relativity
that preserves the speed of light.
8.1 The Lorentz transformation
Consider again inertial frames
S
and
S
0
whose origins coincide at
t
=
t
0
= 0. For
now, neglect the
y
and
z
directions, and consider the relationship between (
x, t
)
and (x
0
, t
0
). The general form is
x
0
= f(x, t), t
0
= g(x, t),
for some functions f and g. This is not very helpful.
In any inertial frame, a free particle moves with constant velocity. So straight
lines in (
x, t
) must map into straight lines in (
x
0
, t
0
). Therefore the relationship
must be linear.
Given that the origins of
S
and
S
0
coincide at
t
=
t
0
= 0, and
S
0
moves with
velocity v relative to S, we know that the line x = vt must map into x
0
= 0.
Combining these two information, the transformation must be of the form
x
0
= γ(x vt), (1)
for some factor
γ
that may depend on
|v|
(not
v
itself. We can use symmetry
arguments to show that γ should take the same value for velocities v and v).
Note that Galilean transformation is compatible with this just take
γ
to
be always 1.
Now reverse the roles of the frames. From the perspective
S
0
,
S
moves with
velocity v. A similar argument leads to
x = γ(x
0
+ vt
0
), (2)
with the same factor
γ
, since
γ
only depends on
|v|
. Now consider a light ray
(or photon) passing through the origin
x
=
x
0
= 0 at
t
=
t
0
= 0. Its trajectory in
S is
x = ct.
We want a γ such that the trajectory in S
0
is
x
0
= ct
0
as well, so that the speed of light is the same in each frame. Substitute these
into (1) and (2)
ct
0
= γ(c v)t
ct = γ(c + v)t
0
Multiply the two equations together and divide by tt
0
to obtain
c
2
= γ
2
(c
2
v
2
).
So
γ =
r
c
2
c
2
v
2
=
1
p
1 (v/c)
2
.
Definition (Lorentz factor). The Lorentz factor is
γ =
1
p
1 (v/c)
2
.
Note that
γ 1 and is an increasing function of |v|.
When v c, then γ 1, and we recover the Galilean transformation.
When |v| c, then γ .
If
|v| c
, then
γ
is imaginary, which is physically impossible (or at least
weird).
If we take
c
, then
γ
= 1. So Galilean transformation is the transfor-
mation we will have if light is infinitely fast. Alternatively, in the world of
Special Relativity, the speed of light is “infinitely fast”.
v
c
γ
1
For the sense of scale, we have the following values of γ at different speeds:
γ = 2 when v = 0.866c.
γ = 10 when v = 0.9949.
γ = 20 when v = 0.999c.
We still have to solve for the relation between
t
and
t
0
. Eliminate
x
between
(1) and (2) to obtain
x = γ(γ(x vt) + vt
0
).
So
t
0
= γt (1 γ
2
)
γx
v
= γ
t
v
c
2
x
.
So we have
Law
(Principle of Special Relativity)
.
Let
S
and
S
0
be inertial frames, moving
at the relative velocity of v. Then
x
0
= γ(x vt)
t
0
= γ
t
v
c
2
x
,
where
γ =
1
p
1 (v/c)
2
.
This is the Lorentz transformations in the standard configuration (in one spatial
dimension).
The above is the form the Lorentz transformation is usually written, and is
convenient for actual calculations. However, this lacks symmetry between space
and time. To display the symmetry, one approach is to use units such that
c
= 1.
Then we have
x
0
= γ(x vt),
t
0
= γ(t vx).
Alternatively, if we want to keep our
c
’s, instead of comparing
x
and
t
, which
have different units, we can compare x and ct. Then we have
x
0
= γ
x
v
c
(ct)
,
ct
0
= γ
ct
v
c
x
.
Symmetries aside, to express
x, t
in terms of
x
0
, t
0
, we can invert this linear
mapping to find (after some algebra)
x = γ(x
0
+ vt
0
)
t = γ
t
0
+
v
c
2
x
0
Directions perpendicular to the relative motion of the frames are unaffected:
y
0
= y
z
0
= z
Now we check that the speed of light is really invariant:
For a light ray travelling in the x direction in S:
x = ct, y = 0, z = 0.
In S
0
, we have
x
0
t
0
=
γ(x vt)
γ(t vx/c
2
)
=
(c v)t
(1 v/c)t
= c,
as required.
For a light ray travelling in the Y direction in S,
x = 0, y = ct, z = 0.
In S
0
,
x
0
t
0
=
γ(x vt)
γ(t vx/c
2
)
= v,
and
y
0
t
0
=
y
γ(t vx/c
2
=
c
γ
,
and
z
0
= 0.
So the speed of light is
p
x
02
+ y
02
t
0
=
p
v
2
+ γ
2
c
2
= c,
as required.
More generally, the Lorentz transformation implies
c
2
t
02
r
02
= c
2
t
02
x
02
y
02
z
02
= c
2
γ
2
t
v
c
2
x
2
γ
2
(x vt)
2
y
2
z
2
= γ
2
1
v
2
c
2
(c
2
t
2
x
2
) y
2
z
2
= c
2
t
2
x
2
y
2
z
2
= c
2
t
2
r
2
.
We say that the quantity c
2
t
2
x
2
y
2
z
2
is Lorentz-invariant.
So if
r
t
= c, then
r
0
t
0
= c also.
8.2 Spacetime diagrams
It is often helpful to plot out what is happening on a diagram. We plot them on
a graph, where the position
x
is on the horizontal axis and the time
ct
is on the
vertical axis. We use ct instead of t so that the dimensions make sense.
x
ct
world line
P
Definition
(Spacetime)
.
The union of space and time in special relativity is
called Minkowski spacetime. Each point
P
represents an event, labelled by
coordinates (ct, x) (note the order!).
A particle traces out a world line in spacetime, which is straight if the particle
moves uniformly.
Light rays moving in the x direction have world lines inclined at 45
.
x
ct
light raylight ray
We can also draw the axes of
S
0
, moving in the
x
direction at velocity
v
relative to
S
. The
ct
0
axis corresponds to
x
0
= 0, i.e.
x
=
vt
. The
x
0
axis
corresponds to t
0
= 0, i.e. t = vx/c
2
.
x
ct
x
0
ct
0
Note that the
x
0
and
ct
0
axes are not orthogonal, but are symmetrical about
the diagonal (dashed line). So they agree on where the world line of a light ray
should lie on.
8.3 Relativistic physics
Now we can look at all sorts of relativistic weirdness!
Simultaneity
The first relativistic weirdness is that different frames disagree on whether two
evens are simultaneous
Definition
(Simultaneous events)
.
We say two events
P
1
and
P
2
are simultane-
ous in the frame S if t
1
= t
2
.
They are represented in the following spacetime diagram by horizontal dashed
lines.
However, events that are simultaneous in
S
0
have equal values of
t
0
, and so
lie on lines
ct
v
c
x = constant.
x
ct
P
1
P
2
x
0
ct
0
The lines of simultaneity of
S
0
and those of
S
are different, and events simulta-
neous in
S
need not be simultaneous in
S
0
. So simultaneity is relative.
S
thinks
P
1
and P
2
happened at the same time, while S
0
thinks P
2
happens first.
Note that this is genuine disagreement. It is not due to effects like, it takes
time for the light conveying the information to different observers. Our account
above already takes that into account (since the whole discussion does not involve
specific observers).
Causality
Although different people may disagree on the temporal order of events, the
consistent ordering of cause and effect can be ensured.
Since things can only travel at at most the speed of light,
P
cannot affect
R
if
R
happens a millisecond after
P
but is at millions of galaxies away. We can
draw a light cone that denotes the regions in which things can be influenced
by
P
. These are the regions of space-time light (or any other particle) can
possibly travel to.
P
can only influence events within its future light cone, and
be influenced by events within its past light cone.
x
ct
P
Q
R
All observers agree that
Q
occurs after
P
. Different observers may disagree on
the temporal ordering of
P
and
R
. However, since nothing can travel faster
than light,
P
and
R
cannot influence each other. Since everyone agrees on how
fast light travels, they also agree on the light cones, and hence causality. So
philosophers are happy.
Time dilation
Suppose we have a clock that is stationary in
S
0
(which travels at constant
velocity
v
with respect to inertial frame
S
) ticks at constant intervals
t
0
. What
is the interval between ticks in S?
Lorentz transformation gives
t = γ
t
0
+
v
c
2
x
0
.
Since x
0
= constant for the clock, we have
t = γt
0
> t
0
.
So the interval measured in S is greater! So moving clocks run slowly.
A non-mathematical explanation comes from Feynman (not lectured): Sup-
pose we have a very simple clock: We send a light beam towards a mirror, and
wait for it to reflect back. When the clock detects the reflected light, it ticks,
and then sends the next light beam.
Then the interval between two ticks is the distance 2
d
divided by the speed
of light.
d
From the point of view of an observer moving downwards, by the time light
reaches the right mirror, it would have moved down a bit. So S sees
d
a
However, the distance travelled by the light beam is now
p
(2d)
2
+ a
2
>
2
d
.
Since they agree on the speed of light, it must have taken longer for the clock to
receive the reflected light in S. So the interval between ticks are longer.
By the principle of relativity, all clocks must measure the same time dilation,
or else we can compare the two clocks and know if we are “moving”.
This is famously evidenced by muons. Their half-life is around 2 microseconds
(i.e. on average they decay to something else after around 2 microseconds). They
are created when cosmic rays bombard the atmosphere. However, even if they
travel at the speed of light, 2 microseconds only allows it to travel
600 m
, certainly
not sufficient to reach the surface of Earth. However, we observe lots of muons
on Earth. This is because muons are travelling so fast that their clocks run
really slowly.
The twin paradox
Consider two twins: Luke and Leia. Luke stays at home. Leia travels at a
constant speed
v
to a distant planet
P
, turns around, and returns at the same
speed.
In Luke’s frame of reference,
x
ct
Luke
cT
2cT
Leia: x = vt
A (Leia’s arrival)
R
Leia’s arrival (A) at P has coordinates
(ct, x) = (cT, vT ).
The time experienced by Leia on her outward journey is
T
0
= γ
T
v
c
2
T
=
T
γ
.
By Leia’s return
R
, Luke has aged by 2
T
, but Leia has aged by
2T
γ
<
2
T
. So
she is younger than Luke, because of time dilation.
The paradox is: From Leia’s perspective, Luke travelled away from her at
speed and the returned, so he should be younger than her!
Why is the problem not symmetric?
We can draw Leia’s initial frame of reference in dashed lines:
x
ct
Leia
A
Han
R
ct
0
x
0
X
Z
In Leia’s frame, by the time she arrives at
A
, she has experienced a time
T
0
=
T
γ
as shown above. This event is simultaneous with event
X
in Leia’s frame. Then
in Luke’s frame, the coordinates of X are
(ct, x) =
cT
0
γ
, 0
=
cT
γ
2
, 0
,
obtained through calculations similar to that above. So Leia thinks Luke has
aged less by a factor of 1
2
. At this stage, the problem is symmetric, and Luke
also thinks Leia has aged less by a factor of 1
2
.
Things change when Leia turns around and changes frame of reference. To
understand this better, suppose Leia meets a friend, Han, who is just leaving
P
at speed
v
. On his journey back, Han also thinks Luke ages
T
2
. But in his
frame of reference, his departure is simultaneous with Luke’s event
Z
, not
X
,
since he has different lines of simultaneity.
So the asymmetry between Luke and Leia occurs when Leia turns around.
At this point, she sees Luke age rapidly from X to Z.
Length contraction
A rod of length L
0
is stationary in S
0
. What is its length in S?
In
S
0
, then length of the rod is the distance between the two ends at the
same time. So we have
x
0
ct
0
L
0
L
0
In S, we have
x
ct
L
x
0
L
0
The lines
x
0
= 0 and
x
0
=
L
0
map into
x
=
vt
and
x
=
vt
+
L
0
. So the length
in
S
is
L
=
L
0
< L
0
. Therefore moving objects are contracted in the direction
of motion.
Definition
(Proper length)
.
The proper length is the length measured in an
object’s rest frame.
This is analogous to the fact that if you view a bar from an angle, it looks
shorter than if you view it from the front. In relativity, what causes the
contraction is not a spatial rotation, but a spacetime hyperbolic rotation.
Question: does a train of length 2
L
fit alongside a platform of length
L
if it
travels through the station at a speed v such that γ = 2?
For the system of observers on the platform, the train contracts to a length
2L/γ = L. So it fits.
But for the system of observers on the train, the platform contracts to length
L/γ = L/2, which is much too short!
This can be explained by the difference of lines of simultaneity, since length
is the distance between front and back at the same time.
x
ctback of train front of train
back of
platform
front of
platform
L
fits in S
doesn’t fit in S
0
Composition of velocities
A particle moves with constant velocity
u
0
in frame
S
0
, which moves with velocity
v relative to S. What is its velocity u in S?
The world line of the particle in S
0
is
x
0
= u
0
t
0
.
In S, using the inverse Lorentz transformation,
u =
x
t
=
γ(x
0
+ vt
0
)
γ(t
0
+ (v/c
2
)x
0
)
=
u
0
t
0
+ vt
0
t
0
+ (v/c
2
)u
0
t
0
=
u
0
+ v
1 + u
0
v/c
2
.
This is the formula for the relativistic composition of velocities.
The inverse transformation is found by swapping
u
and
u
0
, and swapping the
sign of v, i.e.
u
0
=
u v
1 uv/c
2
.
Note the following:
if
u
0
v c
2
, then the transformation reduces to the standard Galilean
addition of velocities u u
0
+ v.
u
is a monotonically increasing function of
u
for any constant
v
(with
|v| < c).
When
u
0
=
±c
,
u
=
u
0
for any
v
, i.e. the speed of light is constant in all
frames of reference.
Hence
|u
0
| < c
iff
|u| < c
. This means that we cannot reach the speed of
light by composition of velocities.
8.4 Geometry of spacetime
We’ll now look at the geometry of spacetime, and study the properties of vectors
in this spacetime. While spacetime has 4 dimensions, and each point can be
represented by 4 real numbers, this is not ordinary
R
4
. This can be seen when
changing coordinate systems, instead of rotating the axes like in
R
4
, we “squash”
the axes towards the diagonal, which is a hyperbolic rotation. In particular,
we will have a different notion of a dot product. We say that this space has
dimension d = 1 + 3.
The invariant interval
In regular Euclidean space, given a vector
x
, all coordinate systems agree on the
length |x|. In Minkowski space, they agree on something else.
Consider events
P
and
Q
with coordinates (
ct
1
, x
1
) and (
ct
2
, x
2
) separated
by t = t
2
t
1
and x = x
2
x
1
.
Definition
(Invariant interval)
.
The invariant interval or spacetime interval
between P and Q is defined as
s
2
= c
2
t
2
x
2
.
Note that this quantity
s
2
can be both positive or negative so
s
might be
imaginary!
Proposition. All inertial observers agree on the value of s
2
.
Proof.
c
2
t
02
x
02
= c
2
γ
2
t
v
c
2
x
2
γ
2
(∆x vt)
2
= γ
2
1
v
2
c
2
(c
2
t
2
x
2
)
= c
2
t
2
x
2
.
In three spatial dimensions,
s
2
= c
2
t
2
x
2
y
2
z
2
.
We take this as the “distance” between the two points. For two infinitesimally
separated events, we have
Definition (Line element). The line element is
ds
2
= c
2
dt
2
dx
2
dy
2
dz
2
.
Definition
(Timelike, spacelike and lightlike separation)
.
Events with
s
2
>
0
are timelike separated. It is possible to find inertial frames in which the two events
occur in the same position, and are purely separated by time. Timelike-separated
events lie within each other’s light cones and can influence one another.
Events with
s
2
<
0 are spacelike separated. It is possible to find inertial
frame in which the two events occur in the same time, and are purely separated
by space. Spacelike-separated events lie out of each other’s light cones and
cannot influence one another.
Events with
s
2
= 0 are lightlike or null separated. In all inertial frames, the
events lie on the boundary of each other’s light cones. e.g. different points in
the trajectory of a photon are lightlike separated, hence the name.
Note that s
2
= 0 does not imply that P and Q are the same event.
The Lorentz group
The coordinates of an event
P
in frame
S
can be written as a 4-vector (i.e.
4-component vector) X. We write
X =
ct
x
y
z
The invariant interval between the origin and
P
can be written as an inner
product
X · X = X
T
ηX = c
2
t
2
x
2
y
2
z
2
,
where
η =
1 0 0 0
0 1 0 0
0 0 1 0
0 0 0 1
.
4-vectors with
X · X >
0 are called timelike, and those
X · X <
0 are spacelike.
If X · X = 0, it is lightlike or null.
A Lorentz transformation is a linear transformation of the coordinates from
one frame S to another S
0
, represented by a 4 × 4 tensor (“matrix”):
X
0
= ΛX
Lorentz transformations can be defined as those that leave the inner product
invariant:
(X)(X
0
· X
0
= X · X),
which implies the matrix equation
Λ
T
ηΛ = η. ()
These also preserve X · Y if X and Y are both 4-vectors.
Two classes of solution to this equation are:
Λ =
1 0 0 0
0
0 R
0
,
where
R
is a 3
×
3 orthogonal matrix, which rotates (or reflects) space and leaves
time intact; and
Λ =
γ γβ 0 0
γβ γ 0 0
0 0 1 0
0 0 0 1
,
where
β
=
v
c
, and
γ
= 1
/
p
1 β
2
. Here we leave the
y
and
z
coordinates intact,
and apply a Lorentz boost along the x direction.
The set of all matrices satisfying equation (
) form the Lorentz group
O
(1
,
3).
It is generated by rotations and boosts, as defined above, which includes the
absurd spatial reflections and time reversal.
The subgroup with det Λ = +1 is the proper Lorentz group SO(1, 3).
The subgroup that preserves spatial orientation and the direction of time is
the restricted Lorentz group
SO
+
(1
,
3). Note that this is different from
SO
(1
,
3),
since if you do both spatial reflection and time reversal, the determinant of the
matrix is still positive. We want to eliminate those as well!
Rapidity
Focus on the upper left 2
×
2 matrix of Lorentz boosts in the
x
direction. Write
Λ[β] =
γ γβ
γβ γ
, γ =
1
p
1 β
2
.
Combining two boosts in the x direction, we have
Λ[β
1
]Λ[β
2
] =
γ
1
γ
1
β
1
γ
1
β
1
γ
1
γ
2
γ
2
β
2
γ
2
β
2
γ
2
= Λ
β
1
+ β
2
1 + β
1
β
2
after some messy algebra. This is just the velocity composition formula as before.
This result does not look nice. This suggests that we might be writing things
in the wrong way.
We can compare this with spatial rotation. Recall that
R(θ) =
cos θ sin θ
sin θ cos θ
with
R(θ
1
)R(θ
2
) = R(θ
1
+ θ
2
).
For Lorentz boosts, we can define
Definition (Rapidity). The rapidity of a Lorentz boot is φ such that
β = tanh φ, γ = cosh φ, γβ = sinh φ.
Then
Λ[β] =
cosh φ sinh φ
sinh φ cosh φ
= Λ(φ).
The rapidities add like rotation angles:
Λ(φ
1
)Λ(φ
2
) = Λ(φ
1
+ φ
2
).
This shows the close relation betweens spatial rotations and Lorentz boosts.
Lorentz boots are simply hyperbolic rotations in spacetime!
8.5 Relativistic kinematics
In Newtonian mechanics, we describe a particle by its position
x
(
t
), with its
velocity being u(t) =
dx
dt
.
In relativity, this is unsatisfactory. In special relativity, space and time can
be mixed together by Lorentz boosts, and we prefer not to single out time from
space. For example, when we write the 4-vector
X
, we put in both the time and
space components, and Lorentz transformations are 4
×
4 matrices that act on
X.
In the definition of velocity, however, we are differentiating space with respect
to time, which is rather weird. First of all, we need something to replace time.
Recall that we defined “proper length” as the length in the item in its rest frame.
Similarly, we can define the proper time.
Definition (Proper time). The proper time τ is defined such that
τ =
s
c
τ
is the time experienced by the particle, i.e. the time in the particles rest frame.
The world line of a particle can be parametrized using the proper time by
t(τ) and x(τ).
x
ct
τ
1
τ
2
Infinitesimal changes are related by
dτ =
ds
c
=
1
c
p
c
2
dt
2
|dx|
2
=
r
1
|u|
2
c
2
dt.
Thus
dt
dτ
= γ
u
with
γ
u
=
1
q
1
|u|
2
c
2
.
The total time experienced by the particle along a segment of its world line is
T =
Z
dτ =
Z
1
γ
u
dt.
We can then define the position 4-vector and 4-velocity.
Definition (Position 4-vector and 4-velocty). The position 4-vect or is
X(τ) =
ct(τ)
x(τ)
.
Its 4-velocity is defined as
U =
dX
dτ
=
c
dt
dτ
dx
dτ
=
dt
dτ
c
u
= γ
u
c
u
,
where u =
dx
dt
.
Another common notation is
X = (ct, x), U = γ
u
(c, u).
If frames
S
and
S
0
are related by
X
0
= Λ
X
, then the 4-velocity also transforms
as U
0
= ΛU.
Definition
(4-vector)
.
A 4-vector is a 4-component vectors that transforms in
this way under a Lorentz transformation, i.e. X
0
= ΛX.
When using suffix notation, the indices are written above (superscript) instead
of below (subscript). The indices are written with Greek letters which range
from 0 to 3. So we have
X
µ
instead of
X
i
, for
µ
= 0
,
1
,
2
,
3. If we write
X
µ
instead, it means a different thing. This will be explained more in-depth in the
electromagnetism course (and you’ll get more confused!).
U
is a 4-vector because
X
is a 4-vector and
τ
is a Lorentz invariant. Note
that dX/dt is not a 4-vector.
Note that this definition of 4-vector is analogous to that of a tensor things
that transform nicely according to our rules. Then
τ
would be a scalar, i.e.
rank-0 tensor, while t is just a number, not a scalar.
For any 4-vector
U
, the inner product
U · U
=
U
0
· U
0
is Lorentz invariant,
i.e. the same in all inertial frames. In the rest frame of the particle,
U
= (
c,
0).
So U · U = c
2
.
In any other frame, Y = γ
u
(c, u). So
Y · Y = γ
2
u
(c
2
|u|
2
) = c
2
as expected.
Transformation of velocities revisited
We have seen that velocities cannot be simply added in relativity. However, the
4-velocity does transform linearly, according to the Lorentz transform:
U
0
= ΛU.
In frame
S
, consider a particle moving with speed
u
at an angle
θ
to the
x
axis
in the
xy
plane. This is the most general case for motion not parallel to the
Lorentz boost.
Its 4-velocity is
U =
γ
u
c
γ
u
u cos θ
γ
u
u sin θ
0
, γ
u
=
1
p
1 u
2
/c
2
.
With frames
S
and
S
0
in standard configuration (i.e. origin coincide at
t
= 0,
S
0
moving in x direction with velocity v relative to S),
U
0
=
γ
u
0
c
γ
u
0
u
0
cos θ
0
γ
u
0
u
0
sin θ
0
0
=
γ
v
γ
v
v/c 0 0
γ
v
v/c γ
v
0 0
0 0 1 0
0 0 0 1
γ
u
c
γ
u
u cos θ
γ
u
u sin θ
0
Instead of evaluating the whole matrix, we can divide different rows to get useful
results.
The ratio of the first two lines gives
u
0
cos θ
0
=
u cos θ v
1
uv
c
2
cos θ
,
just like the composition of parallel velocities.
The ratio of the third to second line gives
tan θ
0
=
u sin θ
γ
v
(u cos θ v)
,
which describes aberration, a change in the direction of motion of a particle due
to the motion of the observer. Note that this isn’t just a relativistic effect! If
you walk in the rain, you have to hold your umbrella obliquely since the rain
seems to you that they are coming from an angle. The relativistic part is the
γ
v
factor in the denominator.
This is also seen in the aberration of starlight (
u
=
c
) due to the Earth’s
orbital motion. This causes small annual changes in the apparent positions of
stars.
4-momentum
Definition (4-momentum). The 4-momentum of a particle of mass m is
P = mU =
u
c
u
The 4-momentum of a system of particles is the sum of the 4-momentum of the
particles, and is conserved in the absence of external forces.
The spatial components of P are the relativistic 3-momentum,
p =
u
u,
which differs from the Newtonian expression by a factor of
γ
u
. Note that
|p|
as |u| c.
What is the interpretation of the time component
P
0
(i.e. the first time
component of the P vector)? We expand for |u| c:
P
0
= c =
mc
p
1 |u|
2
/c
2
=
1
c
mc
2
+
1
2
m|u|
2
+ ···
.
We have a constant term
mc
2
plus a kinetic energy term
1
2
m|u|
2
, plus more
tiny terms, all divided by
c
. So this suggests that
P
0
is indeed the energy for
a particle, and the remaining
···
terms are relativistic corrections for our old
formula
1
2
m|u|
2
(the mc
2
term will be explained later). So we interpret P as
P =
E/c
p
Definition
(Relativistic energy)
.
The relativistic energy of a particle is
E
=
P
0
c
.
So
E = c
2
= mc
2
+
1
2
m|u|
2
+ ···
Note that E as |u| c.
For a stationary particle, we obtain
E = mc
2
.
This implies that mass is a form of energy.
m
is sometimes called the rest mass.
The energy of a moving particle,
u
c
2
, is the sum of the rest energy
mc
2
and kinetic energy m(γ
u
1)c
2
.
Since
P ·P
=
E
2
c
2
|p|
2
is a Lorentz invariant (lengths of 4-vectors are always
Lorentz invariant) and equals
m
2
c
2
in the particle’s rest frame, we have the
general relation between energy and momentum
E
2
= |p|
2
c
2
+ m
2
c
4
In Newtonian physics, mass and energy are separately conserved. In relativity,
mass is not conserved. Instead, it is just another form of energy, and the total
energy, including mass energy, is conserved.
Mass can be converged into kinetic energy and vice versa (e.g. atomic bombs!)
Massless particles
Particles with zero mass (
m
= 0), e.g. photons, can have non-zero momentum
and energy because they travel at the speed of light (γ = ).
In this case,
P · P
= 0. So massless particles have light-like (or null)
trajectories, and no proper time can be defined for such particles.
Other massless particles in the Standard Model of particle physics include
the gluon.
For these particles, energy and momentum are related by
E
2
= |p|
2
c
2
.
So
E = |p|c.
Thus
P =
E
c
1
n
,
where n is a unit (3-)vector in the direction of propagation.
According to quantum mechanics, fundamental “particles” aren’t really
particles but have both particle-like and wave-like properties (if that sounds
confusing, yes it is!). Hence we can assign it a de Broglie wavelength, according
to the de Broglie relation:
|p| =
h
λ
where h 6.63 × 10
34
m
2
kg s
1
is Planck’s constant.
For massless particles, this is consistent with Planck’s relation:
E =
hc
λ
= hν,
where ν =
c
λ
is the wave frequency.
Newton’s second law in special relativity
Definition (4-force). The 4-force is
F =
dP
dτ
This equation is the relativistic counterpart to Newton’s second law.
It is related to the 3-force F by
F = γ
u
F · u/c
F
Expanding the definition of the 4-force componentwise, we obtain
dE
dτ
= γ
u
F · u
dE
dt
= F · u
and
dp
dτ
= γ
u
F
dp
dt
= F
Equivalently, for a particle of mass m,
F = mA,
where
A =
dU
dτ
is the 4-acceleration.
We have
U = γ
u
c
u
So
A = γ
u
dU
dt
= γ
u
˙γ
u
c
γ
u
a + ˙γ
u
u.
where a =
du
dt
and ˙γ
u
= γ
3
u
a·u
c
2
.
In the instantaneous rest frame of a particle, u = 0 and γ
u
= 1. So
U =
c
0
, A =
0
a
Then
U · A
= 0. Since this is a Lorentz invariant, we have
U · A
= 0 in all
frames.
8.6 Particle physics
Many problems can be solved using the conservation of 4-momentum,
P =
E/c
p
,
for a system of particles.
Definition
(Center of momentum frame)
.
The center of momentum (CM) frame,
or zero momentum frame, is an inertial frame in which the total 3-momentum is
P
p = 0.
This exists unless the system consists of one or more massless particle moving
in a single direction.
Particle decay
A particle of mass m
1
decays into two particles of masses m
2
and m
2
.
We have
P
1
= P
2
+ P
3
.
i.e.
E
1
= E
2
+ E
3
p
1
= p
2
+ p
3
.
In the CM frame (i.e. the rest frame of the original particle),
E
1
= m
1
c
2
=
q
|p
2
|
2
c
2
+ m
2
2
c
4
+
q
|p
3
|
2
c
2
+ m
2
2
c
4
m
2
c
2
+ m
3
c
2
.
So decay is possible only if
m
1
m
2
+ m
3
.
(Recall that mass is not conserved in relativity!)
Example. A possible decay path of the Higgs’ particle can be written as
h γγ
Higgs’ particle 2 photons
This is possible by the above criterion, because m
h
0, while m
γ
= 0.
The full conservation equation is
P
h
=
m
h
c
0
= P
γ
1
+ P
γ
2
So
p
γ
1
= p
γ
2
E
γ
1
= E
γ
2
=
1
2
m
h
c
2
.
Particle scattering
When two particles collide and retain heir identities, the total 4-momentum is
conserved:
P
1
+ P
2
= P
3
+ P
4
In the laboratory frame
S
, suppose that particle 1 travels with speed
u
and
collides with particle 2 (at rest).
1
2
21
φ
θ
In the CM frame S
0
,
p
0
1
+ p
0
2
= 0 = p
0
3
+ p
0
4
.
Both before and after the collision, the two particles have equal and opposite
3-momentum.
1 2
p
1
p
2
1
2
p
3
p
4
The scattering angle
θ
0
is undetermined and can be thought of as being random.
However, we can derive some conclusions about the angles
θ
and
φ
in the
laboratory frame.
(staying in
S
0
for the moment) Suppose the particles have equal mass
m
.
They then have the same speed v in S
0
.
Choose axes such that
P
0
1
=
v
c
v
v
0
0
, P
0
2
=
v
c
v
v
0
0
and after the collision,
P
0
3
=
v
c
v
v cos θ
0
v
v sin θ
0
0
, P
0
4
=
v
c
v
v cos θ
0
v
v sin θ
0
0
.
We then use the Lorentz transformation to return to the laboratory frame
S
.
The relative velocity of the frames is v. So the Lorentz transform is
Λ =
γ
v
γ
v
v/c 0 0
γ
v
v/c γ
v
0 0
0 0 1 0
0 0 0 1
and we find
P
1
=
u
c
u
u
0
0
, P
2
=
mc
0
0
0
where
u =
2v
1 + v
2
/c
2
,
(cf. velocity composition formula)
Considering the transformations of P
0
3
and P
0
4
, we obtain
tan θ =
sin θ
0
γ
v
(1 + cos θ
0
)
=
1
γ
v
tan
θ
0
2
,
and
tan φ =
sin θ
0
γ
v
(1 cos θ
0
)
=
1
γ
v
cot
θ
0
2
.
Multiplying these expressions together, we obtain
tan θ tan φ =
1
γ
2
v
.
So even though we do not know what
θ
and
φ
might be, they must be related
by this equation.
In the Newtonian limit, where |v| c, we have γ
v
1. So
tan θ tan φ = 1,
i.e. the outgoing trajectories are perpendicular in S.
Particle creation
Collide two particles of mass
m
fast enough, and you create an extra particle of
mass M .
P
1
+ P
2
= P
3
+ P
4
+ P
5
,
where P
5
is the momentum of the new particle.
In the CM frame,
1 2
v v
P
1
+ P
2
=
2
v
c
0
We have
P
3
+ P
4
+ P
5
=
(E
3
+ E
4
+ E
5
)/c
0
So
2
v
c
2
= E
3
+ E
4
+ E
5
2mc
2
+ Mc
2
.
So in order to create this new particle, we must have
γ
v
1 +
M
2m
.
Alternatively, it occurs only if the initial kinetic energy in the CM frame satisfies
2(γ
v
1)mc
2
Mc
2
.
If we transform to a frame in which the initial speeds are
u
and 0 (i.e. stationary
target), then
u =
2v
1 + v
2
/c
2
Then
γ
u
= 2γ
2
v
1.
So we require
γ
u
2
1 +
M
2m
2
1 = 1 +
2M
m
+
M
2
2m
.
This means that the initial kinetic energy in this frame must be
m(γ
u
1)c
2
2 +
M
2m
Mc
2
,
which could be much larger than
Mc
2
, especially if
M m
, which usually the
case. For example, the mass of the Higgs’ boson is 130 times the mass of the
proton. So it would be much advantageous to collide two beams of protons head
on, as opposed to hitting a fixed target.