0% found this document useful (0 votes)
302 views36 pages

Classical Mechanics 1

This document discusses classical mechanics and relativity. It begins by summarizing two textbooks on classical mechanics and recommending them for understanding basic concepts. It then discusses additional resources on Newtonian mechanics, special relativity, and general relativity. The document provides an overview of the topics that will be covered in the course, including Newton's laws of motion, forces like gravity and electromagnetism, and an introduction to special relativity.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
302 views36 pages

Classical Mechanics 1

This document discusses classical mechanics and relativity. It begins by summarizing two textbooks on classical mechanics and recommending them for understanding basic concepts. It then discusses additional resources on Newtonian mechanics, special relativity, and general relativity. The document provides an overview of the topics that will be covered in the course, including Newton's laws of motion, forces like gravity and electromagnetism, and an introduction to special relativity.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 36

Dynamics and Relativity

–1–
Abstract:

• Tom Kibble and Frank Berkshire, “Classical Mechanics”

• Douglas Gregory, “Classical Mechanics”

Both of these books are well written and do an excellent job of explaining the funda-
mentals of classical mechanics. If you’re struggling to understand some of the basic
concepts, these are both good places to turn.

• S. Chandrasekhar, “Newton’s Principia (for the common reader)”

Want to hear about Newtonian mechanics straight from the horse’s mouth? This is
an annotated version of the Principia with commentary by the Nobel prize winning
astrophysicist Chandrasekhar who walks you through Newton’s geometrical proofs.
Although, in fairness, Newton is sometimes easier to understand than Chandra.

• A.P. French, “Special Relativity”

A clear introduction, covering the theory in some detail.

• Wolfgang Pauli, “Theory of Relativity”

Pauli was one of the founders of quantum mechanics and one of the great physicists of
the last century. Much of this book was written when he was just 21. It remains one
of the most authoritative and scholarly accounts of special relativity. It’s not for the
faint of heart. (But it is cheap).

A number of excellent lecture notes are available on the web. Links can be found on
the course webpage: http://www.damtp.cam.ac.uk/user/tong/relativity.html
Contents

1. Newtonian Mechanics 1
1.1 Newton’s Laws of Motion 2
1.1.1 Newton’s Laws 3
1.2 Inertial Frames and Newton’s First Law 4
1.2.1 Galilean Relativity 5
1.3 Newton’s Second Law 7
1.4 Looking Forwards: The Validity of Newtonian Mechanics 9

2. Forces 10
2.1 Potentials in One Dimension 10
2.1.1 Moving in a Potential 12
2.1.2 Equilibrium: Why (Almost) Everything is a Harmonic Oscillator 15
2.2 Potentials in Three Dimensions 17
2.2.1 Central Forces 19
2.2.2 Angular Momentum 20
2.3 Gravity 21
2.3.1 The Gravitational Field 21
2.3.2 Escape Velocity 23
2.3.3 Inertial vs Gravitational Mass 24
2.4 Electromagnetism 24
2.4.1 The Electric Field of a Point Charge 26
2.4.2 Circles in a Constant Magnetic Field 26
2.4.3 An Aside: Maxwell’s Equations 29
2.5 Friction 29
2.5.1 Dry Friction 30
2.5.2 Fluid Drag 30
2.5.3 An Example: The Damped Harmonic Oscillator 31
2.5.4 Terminal Velocity with Quadratic Friction 33

3. Interlude: Dimensional Analysis 38

–1–
1. Newtonian Mechanics
Classical mechanics is an ambitious theory. Its purpose is to predict the future and
reconstruct the past, to determine the history of every particle in the Universe.

The theory of classical mechanics was formulated by Newton in 1687, building on


earlier insights of Galileo. Starting from a few simple axioms, Newton constructed a
mathematical framework which is powerful enough to explain a broad range of phe-
nomena, from the orbits of the planets, to the motion of the tides, to the scattering of
elementary particles. Before it can be applied to any specific problem, the framework
needs just a single input: a force. With this in place, it is merely a matter of turning
a mathematical handle to reveal what happens next.

We start this course by exploring the framework of Newtonian mechanics, under-


standing the axioms and what they have to tell us about the way the Universe works.
We then move on to look at a number of forces that are at play in the world. Nature is
kind and the list is surprisingly short. Moreover, many of forces that arise have special
properties, from which we will see new concepts emerging such as energy and conserva-
tion principles. Finally, for each of these forces, we turn the mathematical handle. We
turn this handle many many times. In doing so, we will see how classical mechanics is
able to explain large swathes of what we see around us.

Despite its wild success, Newtonian mechanics is not the last word in theoretical
physics. It struggles in extremes: the realm of the very small, the very heavy or the
very fast. We finish these lectures with an introduction to special relativity, the theory
which replaces Newtonian mechanics when the speed of particles is comparable to the
speed of light. We will see how our common sense ideas of space and time are replaced
by something more intricate and more beautiful, with surprising consequences. Time
goes slow for those on the move; lengths get smaller; mass is merely another form of
energy.

Ultimately, the framework of classical mechanics falls short of its ambitious goal to
tell the story of every particle in the Universe. Yet it provides the basis for all that
follows. Some of the Newtonian ideas do not survive to later, more sophisticated,
theories of physics. Even the seemingly primary idea of force will fall by the wayside.
Instead other concepts that we will meet along the way, most notably energy, step to
the fore. But all subsequent theories are built on the Newtonian foundation. Moreover,
developments in the past 300 years have confirmed what is perhaps the most important
legacy of Newton: the laws of Nature are written in the language of mathematics. In
this course, we take the first steps towards understanding these laws.

–1–
1.1 Newton’s Laws of Motion
Classical mechanics is all about the motion of particles. We start with a definition.

Definition: A particle is an object of insignificant size. This means that if you


want to say what a particle looks like at a given time, the only information you have
to specify is its position.

During this course, we will treat electrons, tennis balls, falling cats and planets as
particles. In all of these cases, this means that we only care about the position of the
object and our analysis will not, for example, be able to say anything about the look on
the cat’s face as it falls. However, it’s not immediately obvious that we can meaningfully
assign a single position to a complicated object such as a spinning, mewing cat. Should
we describe its position as the end of its tail or the tip of its nose? We will not provide
an immediate answer to this question, but we will return to it in Section 5 where we
will show that any object can be treated as a point-like particle if we look at the motion
of its centre of mass.

To describe the position of a particle we need a reference frame. This is a choice


of origin, together with a set of axes which, for now, we pick to be Cartesian. With
respect to this frame, the position of a particle is specified by a vector x, which we
denote using bold font. Since the particle moves, the position depends on time, resulting
in a trajectory of the particle described by

x = x(t)

In these notes we will also use both the notation x(t) and r(t) to describe the trajectory
of a particle.

The velocity of a particle is defined to be z

dx(t)
v≡
dt x

Throughout these notes, we will often denote differentiation with


y
respect to time by a “dot” above the variable. So we will also write

v = ẋ Figure 1:

The acceleration of the particle is defined to be


d2 x(t)
a ≡ ẍ =
dt2

–2–
A Comment on Vector Differentiation
The derivative of a vector is defined by differentiating each of the components. So, if
x = (x1 , x2 , x3 ) then
 
dx dx1 dx2 dx3
= , ,
dt dt dt dt
Geometrically, the derivative of a path x(t) lies tangent to the path (a fact which you
will see in the Vector Calculus course).

In this course, we will be working with vector differential equations. These should
be viewed as three, coupled differential equations – one for each component. We will
frequently come across situations where we need to differentiate vector dot-products
and cross-products. The meaning of these is easy to see if we use the chain rule on each
component. For example, given two vector functions of time, f(t) and g(t), we have
d df dg
(f · g) = ·g+f ·
dt dt dt
and
d df dg
(f × g) = ×g+f ×
dt dt dt
As usual, it doesn’t matter what order we write the terms in the dot product, but
we have to be more careful with the cross product because, for example, df/dt × g =
−g × df/dt.

1.1.1 Newton’s Laws


Newtonian mechanics is a framework which allows us to determine the trajectory x(t)
of a particle in any given situation. This framework is usually presented as three axioms
known as Newton’s laws of motion. They look something like:
• N1: Left alone, a particle moves with constant velocity.

• N2: The acceleration (or, more precisely, the rate of change of momentum) of a
particle is proportional to the force acting upon it.

• N3: Every action has an equal and opposite reaction.


While it is worthy to try to construct axioms on which the laws of physics rest, the
trite, minimalistic attempt above falls somewhat short. For example, on first glance,
it appears that the first law is nothing more than a special case of the second law. (If
the force vanishes, the acceleration vanishes which is the same thing as saying that the
velocity is constant). But the truth is somewhat more subtle. In what follows we will
take a closer look at what really underlies Newtonian mechanics.

–3–
1.2 Inertial Frames and Newton’s First Law

Placed in the historical context, it is understandable that Newton wished to stress the
first law. It is a rebuttal to the Aristotelian idea that, left alone, an object will naturally
come to rest. Instead, as Galileo had previously realised, the natural state of an object
is to travel with constant speed. This is the essence of the law of inertia.

However, these days we’re not bound to any Aristotelian dogma. Do we really need
the first law? The answer is yes, but it has a somewhat different meaning.

We’ve already introduced the idea of a frame of reference: a Cartesian coordinate


system in which you measure the position of the particle. But for most reference frames
you can think of, Newton’s first law is obviously incorrect. For example, suppose the
coordinate system that I’m measuring from is rotating. Then, everything will appear
to be spinning around me. If I measure a particle’s trajectory in my coordinates as
x(t), then I certainly won’t find that d2 x/dt2 = 0, even if I leave the particle alone. In
rotating frames, particles do not travel at constant velocity.

We see that if we want Newton’s first law to fly at all, we must be more careful about
the kind of reference frames we’re talking about. We define an inertial reference frame
to be one in which particles do indeed travel at constant velocity when the force acting
on it vanishes. In other words, in an inertial frame

ẍ = 0 when F = 0

The true content of Newton’s first law can then be better stated as

• N1 Revisited: Inertial frames exist.

These inertial frames provide the setting for all that follows. For example, the second
law — which we shall discuss shortly — should be formulated in inertial frames.

One way to ensure that you are in an inertial frame is to insist that you are left alone
yourself: fly out into deep space, far from the effects of gravity and other influences,
turn off your engines and sit there. This is an inertial frame. However, for most
purposes it will suffice to treat axes of the room you’re sitting in as an inertial frame.
Of course, these axes are stationary with respect to the Earth and the Earth is rotating,
both about its own axis and about the Sun. This means that the Earth does not quite
provide an inertial frame and we will study the consequences of this in Section 6.

–4–
1.2.1 Galilean Relativity
Inertial frames are not unique. Given one inertial frame, S, in which a particle has
coordinates x(t), we can always construct another inertial frame S ′ in which the particle
has coordinates x′ (t) by any combination of the following transformations,

• Translations: x ′ = x + a, for constant a.

• Rotations: x ′ = Rx, for a 3 × 3 matrix R obeying RT R = 1. (This also allows for


reflections if det R = −1, although our interest will primarily be on continuous
transformations).

• Boosts: x ′ = x + vt, for constant velocity v.

It is simple to prove that all of these transformations map one inertial frame to another.
Suppose that a particle moves with constant velocity with respect to frame S, so that
d2 x/dt2 = 0. Then, for each of the transformations above, we also have d2 x ′ /dt2 = 0
which tells us that the particle also moves at constant velocity in S ′ . Or, in other
words, if S is an inertial frame then so too is S ′ . The three transformations generate a
group known as the Galilean group.

The three transformations above are not quite the unique transformations that map
between inertial frames. But, for most purposes, they are the only interesting ones!
The others are transformations of the form x ′ = λx for some λ ∈ R. This is just a
trivial rescaling of the coordinates. For example, we may choose to measure distances
in S in units of meters and distances in S ′ in units of parsecs.

We have already mentioned that Newton’s second law is to be formulated in an


inertial frame. But, importantly, it doesn’t matter which inertial frame. In fact, this
is true for all laws of physics: they are the same in any inertial frame. This is known
as the principle of relativity. The three types of transformation laws that make up the
Galilean group map from one inertial frame to another. Combined with the principle
of relativity, each is telling us something important about the Universe

• Translations: There is no special point in the Universe.

• Rotations: There is no special direction in the Universe.

• Boosts: There is no special velocity in the Universe

The first two are fairly unsurprising: position is relative; direction is relative. The third
perhaps needs more explanation. Firstly, it is telling us that there is no such thing as

–5–
“absolutely stationary”. You can only be stationary with respect to something else.
Although this is true (and continues to hold in subsequent laws of physics) it is not
true that there is no special speed in the Universe. The speed of light is special. We
will see how this changes the principle of relativity in Section 7.

So position, direction and velocity are relative. But acceleration is not. You do
not have to accelerate relative to something else. It makes perfect sense to simply say
that you are accelerating or you are not accelerating. In fact, this brings us back to
Newton’s first law: if you are not accelerating, you are sitting in an inertial frame.

The principle of relativity is usually associated to Einstein, but in fact dates back
at least as far as Galileo. In his book, “Dialogue Concerning the Two Chief World
Systems”, Galileo has the character Salviati talk about the relativity of boosts,

Shut yourself up with some friend in the main cabin below decks on some
large ship, and have with you there some flies, butterflies, and other small
flying animals. Have a large bowl of water with some fish in it; hang up a
bottle that empties drop by drop into a wide vessel beneath it. With the
ship standing still, observe carefully how the little animals fly with equal
speed to all sides of the cabin. The fish swim indifferently in all directions;
the drops fall into the vessel beneath; and, in throwing something to your
friend, you need throw it no more strongly in one direction than another,
the distances being equal; jumping with your feet together, you pass equal
spaces in every direction. When you have observed all these things carefully
(though doubtless when the ship is standing still everything must happen
in this way), have the ship proceed with any speed you like, so long as the
motion is uniform and not fluctuating this way and that. You will discover
not the least change in all the effects named, nor could you tell from any of
them whether the ship was moving or standing still.

Galileo Galilei, 1632

Absolute Time
There is one last issue that we have left implicit in the discussion above: the choice of
time coordinate t. If observers in two inertial frames, S and S ′ , fix the units – seconds,
minutes, hours – in which to measure the duration time then the only remaining choice
they can make is when to start the clock. In other words, the time variable in S and
S ′ differ only by

t′ = t + t0

–6–
This is sometimes included among the transformations that make up the Galilean
group.

The existence of a uniform time, measured equally in all inertial reference frames,
is referred to as absolute time. It is something that we will have to revisit when we
discuss special relativity. As with the other Galilean transformations, the ability to
shift the origin of time is reflected in an important property of the laws of physics. The
fundamental laws don’t care when you start the clock. All evidence suggests that the
laws of physics are the same today as they were yesterday. They are time translationally
invariant.

Cosmology
Notably, the Universe itself breaks several of the Galilean transformations. There was
a very special time in the Universe, around 13.7 billion years ago. This is the time of
the Big Bang (which, loosely translated, means “we don’t know what happened here”).

Similarly, there is one inertial frame in which the background Universe is stationary.
The “background” here refers to the sea of photons at a temperature of 2.7 K which
fills the Universe, known as the Cosmic Microwave Background Radiation. This is the
afterglow of the fireball that filled all of space when the Universe was much younger.
Different inertial frames are moving relative to this background and measure the radi-
ation differently: the radiation looks more blue in the direction that you’re travelling,
redder in the direction that you’ve come from. There is an inertial frame in which this
background radiation is uniform, meaning that it is the same colour in all directions.

To the best of our knowledge however, the Universe defines neither a special point,
nor a special direction. It is, to very good approximation, homogeneous and isotropic.

However, it’s worth stressing that this discussion of cosmology in no way invalidates
the principle of relativity. All laws of physics are the same regardless of which inertial
frame you are in. Overwhelming evidence suggests that the laws of physics are the
same in far flung reaches of the Universe. They were the same in first few microseconds
after the Big Bang as they are now.

1.3 Newton’s Second Law


The second law is the meat of the Newtonian framework. It is the famous “F = ma”,
which tells us how a particle’s motion is affected when subjected to a force F. The
correct form of the second law is
d
(mẋ) = F(x, ẋ) (1.1)
dt

–7–
This is usually referred to as the equation of motion. The quantity in brackets is called
the momentum,

p ≡ mẋ

Here m is the mass of the particle or, more precisely, the inertial mass. It is a measure
of the reluctance of the particle to change its motion when subjected to a given force
F. In most situations, the mass of the particle does not change with time. In this case,
we can write the second law in the more familiar form,

mẍ = F(x, ẋ) (1.2)

For much of this course, we will use the form (1.2) of the equation of motion. However,
in Section 5.3, we will briefly look at a few cases where masses are time dependent and
we need the more general form (1.1).

Newton’s second law doesn’t actually tell us anything until someone else tells us what
the force F is in any given situation. We will describe several examples in the next
section. In general, the force can depend on the position x and the velocity ẋ of the
particle, but does not depend on any higher derivatives. We could also, in principle,
consider forces which include an explicit time dependence, F(x, ẋ, t), although we won’t
do so in these lectures. Finally, if more than one (independent) force is acting on the
particle, then we simply take their sum on the right-hand side of (1.2).

The single most important fact about Newton’s equation is that it is a second order
differential equation. This means that we will have a unique solution only if we specify
two initial conditions. These are usually taken to be the position x(t0 ) and the velocity
ẋ(t0 ) at some initial time t0 . However, exactly what boundary conditions you must
choose in order to figure out the trajectory depends on the problem you are trying to
solve. It is not unusual, for example, to have to specify the position at an initial time
t0 and final time tf to determine the trajectory.

The fact that the equation of motion is second order is a deep statement about
the Universe. It carries over, in essence, to all other laws of physics, from quantum
mechanics to general relativity to particle physics. Indeed, the fact that all initial
conditions must come in pairs — two for each “degree of freedom” in the problem
— has important ramifications for later formulations of both classical and quantum
mechanics.

–8–
For now, the fact that the equations of motion are second order means the following:
if you are given a snapshot of some situation and asked “what happens next?” then
there is no way of knowing the answer. It’s not enough just to know the positions of
the particles at some point of time; you need to know their velocities too. However,
once both of these are specified, the future evolution of the system is fully determined
for all time.

1.4 Looking Forwards: The Validity of Newtonian Mechanics


Although Newton’s laws of motion provide an excellent approximation to many phe-
nomena, when pushed to extreme situation they are found wanting. Broadly speaking,
there are three directions in which Newtonian physics needs replacing with a different
framework: they are
• When particles travel at speeds close to the speed light, c ≈ 3 × 108 ms−1 ,
the Newtonian concept of absolute time breaks down and Newton’s laws need
modification. The resulting theory is called special relativity and will be described
in Section 7. As we will see, although the relationship between space and time
is dramatically altered in special relativity, much of the framework of Newtonian
mechanics survives unscathed.

• On very small scales, much more radical change is needed. Here the whole frame-
work of classical mechanics breaks down so that even the most basic concepts,
such as the trajectory of a particle, become ill-defined. The new framework that
holds on these small scales is called quantum mechanics. Nonetheless, there are
quantities which carry over from the classical world to the quantum, in particular
energy and momentum.

• When we try to describe the forces at play between particles, we need to introduce
a new concept: the field. This is a function of both space and time. Familiar
examples are the electric and magnetic fields of electromagnetism. We won’t have
too much to say about fields in this course. For now, we mention only that the
equations which govern the dynamics of fields are always second order differential
equations, similar in spirit to Newton’s equations. Because of this similarity, field
theories are again referred to as “classical”.
Eventually, the ideas of special relativity, quantum mechanics and field theories are
combined into quantum field theory. Here even the concept of particle gets subsumed
into the concept of a field. This is currently the best framework we have to describe
the world around us. But we’re getting ahead of ourselves. Let’s firstly return to our
Newtonian world....

–9–
1. Newton’s Laws of Motion

“So few went to hear him, and fewer understood him, that oftimes he did,
for want of hearers, read to the walls. He usually stayed about half an hour;
when he had no auditors he commonly returned in a quarter of that time.”

Appraisal of a Cambridge lecturer in classical mechanics, circa 1690

1.1 Introduction

The fundamental principles of classical mechanics were laid down by Galileo and New-
ton in the 16th and 17th centuries. In 1686, Newton wrote the Principia where he
gave us three laws of motion, one law of gravity and pretended he didn’t know cal-
culus. Probably the single greatest scientific achievement in history, you might think
this pretty much wraps it up for classical mechanics. And, in a sense, it does. Given
a collection of particles, acted upon by a collection of forces, you have to draw a nice
diagram, with the particles as points and the forces as arrows. The forces are then
added up and Newton’s famous “F = ma” is employed to figure out where the par-
ticle’s velocities are heading next. All you need is enough patience and a big enough
computer and you’re done.

From a modern perspective this is a little unsatisfactory on several levels: it’s messy
and inelegant; it’s hard to deal with problems that involve extended objects rather than
point particles; it obscures certain features of dynamics so that concepts such as chaos
theory took over 200 years to discover; and it’s not at all clear what the relationship is
between Newton’s classical laws and quantum physics.

The purpose of this course is to resolve these issues by presenting new perspectives
on Newton’s ideas. We shall describe the advances that took place during the 150
years after Newton when the laws of motion were reformulated using more powerful
techniques and ideas developed by some of the giants of mathematical physics: people
such as Euler, Lagrange, Hamilton and Jacobi. This will give us an immediate practical
advantage, allowing us to solve certain complicated problems with relative ease (the
strange motion of spinning tops is a good example). But, perhaps more importantly,
it will provide an elegant viewpoint from which we’ll see the profound basic principles
which underlie Newton’s familiar laws of motion. We shall prise open “F = ma” to
reveal the structures and symmetries that lie beneath.

–1–
Moreover, the formalisms that we’ll develop here are the basis for all of fundamental
modern physics. Every theory of Nature, from electromagnetism and general relativity,
to the standard model of particle physics and more speculative pursuits such as string
theory, is best described in the language we shall develop in this course. The new
formalisms that we’ll see here also provide the bridge between the classical world and
the quantum world.

There are phenomena in Nature for which these formalisms are not particularly
useful. Systems which are dissipative, for example, are not so well suited to these
new techniques. But if you want to understand the dynamics of planets and stars and
galaxies as they orbit and spin, or you want to understand what’s happening at the
LHC where protons are collided at unprecedented energies, or you want to know how
electrons meld together in solids to form new states of matter, then the foundations
that we’ll lay in in this course are a must.

1.2 Newtonian Mechanics: A Single Particle


In the rest of this section, we’ll take a flying tour through the basic ideas of classical
mechanics handed down to us by Newton. We’ll start with a single particle.

A particle is defined to be an object of insignificant size. e.g. an electron, a tennis


ball or a planet. Obviously the validity of this statement depends on the context: to
first approximation, the earth can be treated as a particle when computing its orbit
around the sun. But if you want to understand its spin, it must be treated as an
extended object.

The motion of a particle of mass m at the position r is governed by Newton’s Second


Law F = ma or, more precisely,

F(r, ṙ) = ṗ (1.1)

where F is the force which, in general, can depend on both the position r as well as
the velocity ṙ (for example, friction forces depend on ṙ) and p = mṙ is the momentum.
Both F and p are 3-vectors which we denote by the bold font. Equation (1.1) reduces
to F = ma if ṁ = 0. But if m = m(t) (e.g. in rocket science) then the form with ṗ is
correct.

General theorems governing differential equations guarantee that if we are given r


and ṙ at an initial time t = t0 , we can integrate equation (1.1) to determine r(t) for all
t (as long as F remains finite). This is the goal of classical dynamics.

–2–
Equation (1.1) is not quite correct as stated: we must add the caveat that it holds
only in an inertial frame. This is defined to be a frame in which a free particle with
ṁ = 0 travels in a straight line,

r = r0 + vt (1.2)

Newtons’s first law is the statement that such frames exist.

An inertial frame is not unique. In fact, there are an infinite number of inertial frames.
Let S be an inertial frame. Then there are 10 linearly independent transformations
S → S ′ such that S ′ is also an inertial frame (i.e. if (1.2) holds in S, then it also holds
in S ′ ). These are

• 3 Rotations: r′ = Or where O is a 3 × 3 orthogonal matrix.

• 3 Translations: r′ = r + c for a constant vector c.

• 3 Boosts: r′ = r + ut for a constant velocity u.

• 1 Time Translation: t′ = t + c for a constant real number c

If motion is uniform in S, it will also be uniform in S ′ . These transformations make


up the Galilean Group under which Newton’s laws are invariant. They will be impor-
tant in section 2.4 where we will see that these symmetries of space and time are the
underlying reason for conservation laws. As a parenthetical remark, recall from special
relativity that Einstein’s laws of motion are invariant under Lorentz transformations
which, together with translations, make up the Poincaré group. We can recover the
Galilean group from the Poincaré group by taking the speed of light to infinity.

1.2.1 Angular Momentum


We define the angular momentum L of a particle and the torque τ acting upon it as

L= r×p , τ =r×F (1.3)

Note that, unlike linear momentum p, both L and τ depend on where we take the
origin: we measure angular momentum with respect to a particular point. Let us cross
both sides of equation (1.1) with r. Using the fact that ṙ is parallel to p, we can write
d
dt
(r × p) = r × ṗ. Then we get a version of Newton’s second law that holds for angular
momentum:

τ = L̇ (1.4)

–3–
1.2.2 Conservation Laws
From (1.1) and (1.4), two important conservation laws follow immediately.

• If F = 0 then p is constant throughout the motion

• If τ = 0 then L is constant throughout the motion


Notice that τ = 0 does not require F = 0, but only r × F = 0. This means that F
must be parallel to r. This is the definition of a central force. An example is given by
the gravitational force between the earth and the sun: the earth’s angular momentum
about the sun is constant. As written above in terms of forces and torques, these
conservation laws appear trivial. In section 2.4, we’ll see how they arise as a property
of the symmetry of space as encoded in the Galilean group.

1.2.3 Energy
Let’s now recall the definitions of energy. We firstly define the kinetic energy T as

T = 12 m ṙ · ṙ (1.5)

Suppose from now on that the mass is constant. We can compute the change of kinetic
energy with time: dT dt
= ṗ · ṙ = F · ṙ. If the particle travels from position r1 at time t1
to position r2 at time t2 then this change in kinetic energy is given by
Z t2 Z t2 Z r2
dT
T (t2 ) − T (t1 ) = = F · ṙ dt = F · dr (1.6)
t1 dt t1 r1

where the final expression involving the integral of the force over the path is called the
work done by the force. So we see that the work done is equal to the change in kinetic
energy. From now on we will mostly focus on a very special type of force known as a
conservative force. Such a force depends only on position r rather than velocity ṙ and
is such that the work done is independent of the path taken. In particular, for a closed
path, the work done vanishes.
I
F · dr = 0 ⇔ ∇×F=0 (1.7)

It is a deep property of flat space R3 that this property implies we may write the force
as

F = −∇V (r) (1.8)

for some potential V (r). Systems which admit a potential of this form include gravi-
tational, electrostatic and interatomic forces. When we have a conservative force, we

–4–
necessarily have a conservation law for energy. To see this, return to equation (1.6)
which now reads
Z r2
T (t2 ) − T (t1 ) = − ∇V · dr = −V (t2 ) + V (t1 ) (1.9)
r1

or, rearranging things,

T (t1 ) + V (t1 ) = T (t2 ) + V (t2 ) ≡ E (1.10)

So E = T + V is also a constant of motion. It is the energy. When the energy is


considered to be a function of position r and momentum p it is referred to as the
Hamiltonian H. In section 4 we will be seeing much more of the Hamiltonian.

1.2.4 Examples
• Example 1: The Simple Harmonic Oscillator

This is a one-dimensional system with a force proportional to the distance x to the


origin: F (x) = −kx. This force arises from a potential V = 21 kx2 . Since F 6= 0,
momentum is not conserved (the object oscillates backwards and forwards) and, since
the system lives in only one dimension, angular momentum is not defined. But energy
E = 12 mẋ2 + 21 kx2 is conserved.

• Example 2: The Damped Simple Harmonic Oscillator

We now include a friction term so that F (x, ẋ) = −kx−γ ẋ. Since F is not conservative,
energy is not conserved. This system loses energy until it comes to rest.

• Example 3: Particle Moving Under Gravity

Consider a particle of mass m moving in 3 dimensions under the gravitational pull of


a much larger particle of mass M. The force is F = −(GMm/r 2 )r̂ which arises from
the potential V = −GMm/r. Again, the linear momentum p of the smaller particle
is not conserved, but the force is both central and conservative, ensuring the particle’s
total energy E and the angular momentum L are conserved.

1.3 Newtonian Mechanics: Many Particles


It’s easy to generalise the above discussion to many particles: we simply add an index
to everything in sight! Let particle i have mass mi and position ri where i = 1, . . . , N
is the number of particles. Newton’s law now reads

Fi = ṗi (1.11)

–5–
where Fi is the force on the ith particle. The subtlety is that forces can now be working
between particles. In general, we can decompose the force in the following way:
X
Fi = Fij + Fext
i (1.12)
j6=i

where Fij is the force acting on the ith particle due to the j th particle, while Fext
i is the
th
external force on the i particle. We now sum over all N particles
X X X
Fi = Fij + Fext
i
i i,j with j6=i i
X X
= (Fij + Fji ) + Fext
i (1.13)
i<j i

where, in the second line, we’ve re-written the sum to be over all pairs i < j. At
this stage we make use of Newton’s third law of motion: every action has an equal an
opposite reaction. Or, in other words, Fij = −Fji . We see that the first term vanishes
and we are left simply with
X
Fi = Fext (1.14)
i

where we’ve defined the total external force to be Fext = i Fext


P
P i . We now define the
total mass of the system M = i mi as well as the centre of mass R
P
mi ri
R= i (1.15)
M
Then using (1.11), and summing over all particles, we arrive at the simple formula,

Fext = M R̈ (1.16)

which is identical to that of a single particle. This is an important formula. It tells that
the centre of mass of a system of particles acts just as if all the mass were concentrated
there. In other words, it doesn’t matter if you throw a tennis ball or a very lively cat:
the center of mass of each traces the same path.

1.3.1 Momentum Revisited


P
The total momentum is defined to be P = i pi and, from the formulae above, it is
simple to derive Ṗ = Fext . So we find the conservation law of total linear momentum
for a system of many particles: P is constant if Fext vanishes.

–6–
P
Similarly, we define total angular momentum to be L = i Li . Now let’s see what
happens when we compute the time derivative.
X
L̇ = ri × ṗi
i
!
X X
= ri × Fij + Fext
i (1.17)
i j6=i
X X
= ri × Fji + ri × Fext
i (1.18)
i,j with i6=j i

The last term in this expression is the definition of total external torque: τ ext = i ri ×
P

Fext
i . But what are we going to do with the first term on the right hand side? Ideally we
would like it to vanish! Let’s look at the circumstances under which this will happen.
We can again rewrite it as a sum over pairs i < j to get
X
(ri − rj ) × Fij (1.19)
i<j

which will vanish if and only if the force Fij is parallel to the line joining to two particles
(ri − rj ). This is the strong form of Newton’s third law. If this is true, then we have a
statement about the conservation of total angular momentum, namely L is constant if
τ ext = 0.

Most forces do indeed obey both forms of Newton’s third law: 1


Fij = −Fji and Fij is parallel to (ri −rj ). For example, gravitational
and electrostatic forces have this property. And the total momentum
and angular momentum are both conserved in these systems. But
some forces don’t have these properties! The most famous example
is the Lorentz force on two moving particles with electric charge Q.
This is given by,
2
Fij = Qvi × Bj (1.20)

where vi is the velocity of the ith particle and Bj is the magnetic


field generated by the j th particle. Consider two particles crossing Figure 1: The
magnetic field for
each other in a “T” as shown in the diagram. The force on particle
two particles.
1 from particle 2 vanishes. Meanwhile, the force on particle 2 from
particle 1 is non-zero, and in the direction

F21 ∼ ↑ ×⊗ ∼ ← (1.21)

–7–
Does this mean that conservation of total linear and angular momentum is violated?
Thankfully, no! We need to realise that the electromagnetic field itself carries angular
momentum which restores the conservation law. Once we realise this, it becomes a
rather cheap counterexample to Newton’s third law, little different from an underwater
swimmer who can appear to violate Newton’s third law if we don’t take into account
the momentum of the water.

1.3.2 Energy Revisited


1
mi ṙ2i . Let us
P
The total kinetic energy of a system of many particles is T = 2 i
decompose the position vector ri as
ri = R + r̃i (1.22)
where r̃i is the distance from the centre of mass to the particle i. Then we can write
the total kinetic energy as
2
mi r̃˙ i
X
T = 12 M Ṙ2 + 21 (1.23)
i

Which shows us that the kinetic energy splits up into the kinetic energy of the centre
of mass, together with an internal energy describing how the system is moving around
its centre of mass. As for a single particle, we may calculate the change in the total
kinetic energy,
XZ XZ
ext
T (t2 ) − T (t1 ) = Fi · dri + Fij · dri (1.24)
i i6=j

Like before, we need to consider conservative forces to get energy conservation. But
now we need both
• Conservative external forces: Fext
i = −∇i Vi (r1 , . . . , rN )

• Conservative internal forces: Fij = −∇i Vij (r1 , . . . , rN )


where ∇i ≡ ∂/∂ri . To get Newton’s third law Fij = −Fji together with the requirement
that this is parallel to (ri −rj ), we should take the internal potentials to satisfy Vij = Vji
with
Vij (r1 , . . . rN ) = Vij (|ri − rj |) (1.25)
so that Vij depends only on the distance between the ith and j th particles. We also
insist on a restriction for the external forces, Vi (r1 , . . . rN ) = Vi (ri ), so that the force
on particle i does not depend on the positions of the other particles. Then, following
the steps we took in the single particle case, we can define the total potential energy
P P
V = i Vi + i<j Vij and we can show that H = T + V is conserved.

–8–
1.3.3 An Example
Let us return to the case of gravitational attraction between two bodies but, unlike
in Section 1.2.4, now including both particles. We have T = 21 m1 ṙ21 + 21 m2 ṙ22 . The
potential is V = −Gm1 m2 /|r1 − r2 |. This system has total linear momentum and total
angular mometum conserved, as well as the total energy H = T + V .

–9–
Chapter 1

Introduction to Dynamics

1.1 Introduction and Review

Dynamics is the science of how things move. A complete solution to the motion of a system
means that we know the coordinates of all its constituent particles as functions of time.
For a single point particle moving in three-dimensional space, this means we want to know
its position vector r(t) as a function of time. If there are many particles, the motion is
described by a set of functions ri (t), where i labels which particle we are talking about. So
generally speaking, solving for the motion means being able to predict where a particle will
be at any given instant of time. Of course, knowing the function ri (t) means we can take
its derivative and obtain the velocity vi (t) = dri /dt at any time as well.
The complete motion for a system is not given to us outright, but rather is encoded in a
set of differential equations, called the equations of motion. An example of an equation of
motion is
d2x
m = −mg (1.1)
dt2

with the solution

x(t) = x0 + v0 t − 12 gt2 (1.2)

where x0 and v0 are constants corresponding to the initial boundary conditions on the
position and velocity: x(0) = x0 , v(0) = v0 . This particular solution describes the vertical
motion of a particle of mass m moving near the earth’s surface.
In this class, we shall discuss a general framework by which the equations of motion may
be obtained, and methods for solving them. That “general framework” is Lagrangian Dy-
namics, which itself is really nothing more than an elegant restatement of Isaac Newton’s
Laws of Motion.

3
4 CHAPTER 1. INTRODUCTION TO DYNAMICS

1.1.1 Newton’s laws of motion

Aristotle held that objects move because they are somehow impelled to seek out their
natural state. Thus, a rock falls because rocks belong on the earth, and flames rise because
fire belongs in the heavens. To paraphrase Wolfgang Pauli, such notions are so vague as to
be “not even wrong.” It was only with the publication of Newton’s Principia in 1687 that
a theory of motion which had detailed predictive power was developed.
Newton’s three Laws of Motion may be stated as follows:

I. A body remains in uniform motion unless acted on by a force.


II. Force equals rate of change of momentum: F = dp/dt.
III. Any two bodies exert equal and opposite forces on each other.

Newton’s First Law states that a particle will move in a straight line at constant (possibly
zero) velocity if it is subjected to no forces. Now this cannot be true in general, for suppose
we encounter such a “free” particle and that indeed it is in uniform motion, so that r(t) =
r0 + v0 t. Now r(t) is measured in some coordinate system, and if instead we choose
to measure r(t) in a different coordinate system whose origin R moves according to the
function R(t), then in this new “frame of reference” the position of our particle will be

r ′ (t) = r(t) − R(t)


= r0 + v0 t − R(t) . (1.3)

If the acceleration d2R/dt2 is nonzero, then merely by shifting our frame of reference we have
apparently falsified Newton’s First Law – a free particle does not move in uniform rectilinear
motion when viewed from an accelerating frame of reference. Thus, together with Newton’s
Laws comes an assumption about the existence of frames of reference – called inertial frames
– in which Newton’s Laws hold. A transformation from one frame K to another frame K′
which moves at constant velocity V relative to K is called a Galilean transformation. The
equations of motion of classical mechanics are invariant (do not change) under Galilean
transformations.
At first, the issue of inertial and noninertial frames is confusing. Rather than grapple with
this, we will try to build some intuition by solving mechanics problems assuming we are
in an inertial frame. The earth’s surface, where most physics experiments are done, is not
an inertial frame, due to the centripetal accelerations associated with the earth’s rotation
about its own axis and its orbit around the sun. In this case, not only is our coordinate
system’s origin – somewhere in a laboratory on the surface of the earth – accelerating, but
the coordinate axes themselves are rotating with respect to an inertial frame. The rotation
of the earth leads to fictitious “forces” such as the Coriolis force, which have large-scale
consequences. For example, hurricanes, when viewed from above, rotate counterclockwise
in the northern hemisphere and clockwise in the southern hemisphere. Later on in the course
we will devote ourselves to a detailed study of motion in accelerated coordinate systems.
Newton’s “quantity of motion” is the momentum p, defined as the product p = mv of a
particle’s mass m (how much stuff there is) and its velocity (how fast it is moving). In
1.1. INTRODUCTION AND REVIEW 5

order to convert the Second Law into a meaningful equation, we must know how the force
F depends on the coordinates (or possibly velocities) themselves. This is known as a force
law. Examples of force laws include:

Constant force: F = −mg

Hooke’s Law: F = −kx

Gravitation: F = −GM m r̂/r 2

v
Lorentz force: F = qE +q ×B
c

Fluid friction (v small): F = −b v .

Note that for an object whose mass does not change we can write the Second Law in the
familiar form F = ma, where a = dv/dt = d2r/dt2 is the acceleration. Most of our initial
efforts will lie in using Newton’s Second Law to solve for the motion of a variety of systems.
The Third Law is valid for the extremely important case of central forces which we will
discuss in great detail later on. Newtonian gravity – the force which makes the planets orbit
the sun – is a central force. One consequence of the Third Law is that in free space two
isolated particles will accelerate in such a way that F1 = −F2 and hence the accelerations
are parallel to each other, with
a1 m2
=− , (1.4)
a2 m1
where the minus sign is used here to emphasize that the accelerations are in opposite
directions. We can also conclude that the total momentum P = p1 + p2 is a constant, a
result known as the conservation of momentum.

1.1.2 Aside : inertial vs. gravitational mass

In addition to postulating the Laws of Motion, Newton also deduced the gravitational force
law, which says that the force Fij exerted by a particle i by another particle j is

ri − rj
Fij = −Gmi mj , (1.5)
|ri − rj |3

where G, the Cavendish constant (first measured by Henry Cavendish in 1798), takes the
value
G = (6.6726 ± 0.0008) × 10−11 N · m2 /kg2 . (1.6)
Notice Newton’s Third Law in action: Fij + Fji = 0. Now a very important and special
feature of this “inverse square law” force is that a spherically symmetric mass distribution
has the same force on an external body as it would if all its mass were concentrated at its
6 CHAPTER 1. INTRODUCTION TO DYNAMICS

center. Thus, for a particle of mass m near the surface of the earth, we can take mi = m
and mj = Me , with ri − rj ≃ Re r̂ and obtain

F = −mgr̂ ≡ −mg (1.7)

where r̂ is a radial unit vector pointing from the earth’s center and g = GMe /Re2 ≃ 9.8 m/s2
is the acceleration due to gravity at the earth’s surface. Newton’s Second Law now says
that a = −g, i.e. objects accelerate as they fall to earth. However, it is not a priori clear
why the inertial mass which enters into the definition of momentum should be the same
as the gravitational mass which enters into the force law. Suppose, for instance, that the
gravitational mass took a different value, m′ . In this case, Newton’s Second Law would
predict
m′
a=− g (1.8)
m

and unless the ratio m′ /m were the same number for all objects, then bodies would fall
with different accelerations. The experimental fact that bodies in a vacuum fall to earth at
the same rate demonstrates the equivalence of inertial and gravitational mass, i.e. m′ = m.

1.2 Examples of Motion in One Dimension

To gain some experience with solving equations of motion in a physical setting, we consider
some physically relevant examples of one-dimensional motion.

1.2.1 Uniform force

With F = −mg, appropriate for a particle falling under the influence of a uniform gravita-
tional field, we have m d2x/dt2 = −mg, or ẍ = −g. Notation:

dx d2x 7
¨¨˙ = d x ,
ẋ ≡ , ẍ ≡ , ẍ etc. (1.9)
dt dt2 dt7

With v = ẋ, we solve dv/dt = −g:

Zv(t) Zt
dv = ds (−g) (1.10)
v(0) 0

v(t) − v(0) = −gt . (1.11)

Note that there is a constant of integration, v(0), which enters our solution.
1.2. EXAMPLES OF MOTION IN ONE DIMENSION 7

We are now in position to solve dx/dt = v:

Zx(t) Zt
dx = ds v(s) (1.12)
x(0) 0

Zt
 
x(t) = x(0) + ds v(0) − gs (1.13)
0
= x(0) + v(0)t − 12 gt2 . (1.14)

Note that a second constant of integration, x(0), has appeared.

1.2.2 Uniform force with linear frictional damping

In this case,
dv
m = −mg − γv (1.15)
dt
which may be rewritten
dv γ
= − dt (1.16)
v + mg/γ m
d ln(v + mg/γ) = −(γ/m)dt . (1.17)

Integrating then gives


 
v(t) + mg/γ
ln = −γt/m (1.18)
v(0) + mg/γ
 
mg mg
v(t) = − + v(0) + e−γt/m . (1.19)
γ γ

Note that the solution to the first order ODE mv̇ = −mg − γv entails one constant of
integration, v(0).
One can further integrate to obtain the motion
 
m mg mg
x(t) = x(0) + v(0) + (1 − e−γt/m ) − t. (1.20)
γ γ γ

The solution to the second order ODE mẍ = −mg − γ ẋ thus entails two constants of
integration: v(0) and x(0). Notice that as t goes to infinity the velocity tends towards
the asymptotic value v = −v∞ , where v∞ = mg/γ. This is known as the terminal veloc-
ity. Indeed, solving the equation v̇ = 0 gives v = −v∞ . The initial velocity is effectively
“forgotten” on a time scale τ ≡ m/γ.
Electrons moving in solids under the influence of an electric field also achieve a terminal
velocity. In this case the force is not F = −mg but rather F = −eE, where −e is the
8 CHAPTER 1. INTRODUCTION TO DYNAMICS

electron charge (e > 0) and E is the electric field. The terminal velocity is then obtained
from
v∞ = eE/γ = eτ E/m . (1.21)
The current density is a product:

current density = (number density) × (charge) × (velocity)

j = n · (−e) · (−v∞ )

ne2 τ
= E. (1.22)
m
The ratio j/E is called the conductivity of the metal, σ. According to our theory, σ =
ne2 τ /m. This is one of the most famous equations of solid state physics! The dissipation
is caused by electrons scattering off impurities and lattice vibrations (“phonons”). In high
purity copper at low temperatures (T < ∼ 4 K), the scattering time τ is about a nanosecond
−9
(τ ≈ 10 s).

1.2.3 Uniform force with quadratic frictional damping

At higher velocities, the frictional damping is proportional to the square of the velocity.
The frictional force is then Ff = −cv 2 sgn (v), where sgn (v) is the sign of v: sgn (v) = +1
if v > 0 and sgn (v) = −1 if v < 0. (Note one can also write sgn (v) = v/|v| where |v| is
the absolute value.) Why all this trouble with sgn (v)? Because it is important that the
frictional force dissipate energy, and therefore that Ff be oppositely directed with respect to
the velocity v. We will assume that v < 0 always, hence Ff = +cv 2 .
2
p is a terminal velocity, since setting v̇ = −g + (c/m)v = 0 gives v = ±v∞ ,
Notice that there
where v∞ = mg/c. One can write the equation of motion as

dv g
= 2 (v 2 − v∞
2
) (1.23)
dt v∞

and using  
1 1 1 1
= − (1.24)
v 2 − v∞
2 2v∞ v − v∞ v + v∞
we obtain
dv 1 dv 1 dv
= −
v2 2
− v∞ 2v∞ v − v∞ 2v∞ v + v∞
 
1 v∞ − v
= d ln
2v∞ v∞ + v
g
= 2 dt . (1.25)
v∞
1.2. EXAMPLES OF MOTION IN ONE DIMENSION 9

Assuming v(0) = 0, we integrate to obtain


 
1 v∞ − v(t) gt
ln = 2 (1.26)
2v∞ v∞ + v(t) v∞
which may be massaged to give the final result

v(t) = −v∞ tanh(gt/v∞ ) . (1.27)

Recall that the hyperbolic tangent function tanh(x) is given by

sinh(x) ex − e−x
tanh(x) = = x . (1.28)
cosh(x) e + e−x
Again, as t → ∞ one has v(t) → −v∞ , i.e. v(∞) = −v∞ .
Advanced Digression: To gain an understanding of the constant c, consider a flat surface
of area S moving through a fluid at velocity v (v > 0). During a time ∆t, all the fluid
molecules inside the volume ∆V = S · v ∆t will have executed an elastic collision with the
moving surface. Since the surface is assumed to be much more massive than each fluid
molecule, the center of mass frame for the surface-molecule collision is essentially the frame
of the surface itself. If a molecule moves with velocity u is the laboratory frame, it moves
with velocity u − v in the center of mass (CM) frame, and since the collision is elastic, its
final CM frame velocity is reversed, to v − u. Thus, in the laboratory frame the molecule’s
velocity has become 2v − u and it has suffered a change in velocity of ∆u = 2(v − u). The
total momentum change is obtained by multiplying ∆u by the total mass M = ̺ ∆V , where
̺ is the mass density of the fluid. But then the total momentum imparted to the fluid is

∆P = 2(v − u) · ̺ S v ∆t (1.29)

and the force on the fluid is


∆P
F = = 2S ̺ v(v − u) . (1.30)
∆t
Now it is appropriate to average this expression over the microscopic distribution of molec-
ular velocities u, and since on average hui = 0, we obtain the result hF i = 2S̺v 2 , where
h· · · i denotes a microscopic average over the molecular velocities in the fluid. (There is a
subtlety here concerning the effect of fluid molecules striking the surface from either side –
you should satisfy yourself that this derivation is sensible!) Newton’s Third Law then states
that the frictional force imparted to the moving surface by the fluid is Ff = −hF i = −cv 2 ,
where c = 2S̺. In fact, our derivation is too crude to properly obtain the numerical prefac-
tors, and it is better to write c = µ̺S, where µ is a dimensionless constant which depends
on the shape of the moving object.

1.2.4 Crossed electric and magnetic fields

Consider now a three-dimensional example of a particle of charge q moving in mutually


perpendicular E and B fields. We’ll throw in gravity for good measure. We take E = E x̂,
10 CHAPTER 1. INTRODUCTION TO DYNAMICS

B = B ẑ, and g = −gẑ. The equation of motion is Newton’s 2nd Law again:
m r̈ = mg + qE + qc ṙ × B . (1.31)
The RHS (right hand side) of this equation is a vector sum of the forces due to gravity plus
the Lorentz force of a moving particle in an electromagnetic field. In component notation,
we have
qB
mẍ = qE + ẏ (1.32)
c
qB
mÿ = − ẋ (1.33)
c
mz̈ = −mg . (1.34)

The equations for coordinates x and y are coupled, while that for z is independent and may
be immediately solved to yield
z(t) = z(0) + ż(0) t − 12 gt2 . (1.35)

The remaining equations may be written in terms of the velocities vx = ẋ and vy = ẏ:
v̇x = ωc (vy + uD ) (1.36)
v̇y = −ωc vx , (1.37)

where ωc = qB/mc is the cyclotron frequency and uD = cE/B is the drift speed for the
particle. As we shall see, these are the equations for a harmonic oscillator. The solution is

vx (t) = vx (0) cos(ωc t) + vy (0) + uD sin(ωc t) (1.38)

vy (t) = −uD + vy (0) + uD cos(ωc t) − vx (0) sin(ωc t) . (1.39)
Integrating again, the full motion is given by:
x(t) = x(0) + A sin δ + A sin(ωc t − δ) (1.40)
y(r) = y(0) − uD t − A cos δ + A cos(ωc t − δ) , (1.41)
where q  
1 2 −1 ẏ(0) + uD
A= 2
ẋ (0) + ẏ(0) + uD , δ = tan . (1.42)
ωc ẋ(0)
Thus, in the full solution of the motion there are six constants of integration:
x(0) , y(0) , z(0) , A , δ , ż(0) . (1.43)
Of course instead of A and δ one may choose as constants of integration ẋ(0) and ẏ(0).

1.3 Pause for Reflection

In mechanical systems, for each coordinate, or “degree of freedom,” there exists a cor-
responding second order ODE. The full solution of the motion of the system entails two
constants of integration for each degree of freedom.
Chapter 2

Systems of Particles

2.1 Work-Energy Theorem

Consider a system of many particles, with positions ri and velocities ṙi . The kinetic energy
of this system is X X
1 2
T = Ti = 2 mi ṙi . (2.1)
i i
Now let’s consider how the kinetic energy of the system changes in time. Assuming each
mi is time-independent, we have
dTi
= mi ṙi · r̈i . (2.2)
dt
Here, we’ve used the relation
d  dA
A2 = 2 A · . (2.3)
dt dt
We now invoke Newton’s 2nd Law, mi r̈i = Fi , to write eqn. 2.2 as Ṫi = Fi · ṙi . We integrate
this equation from time tA to tB :
ZtB
(B) (A) dTi
Ti − Ti = dt
dt
tA
ZtB X (A→B)
= dt Fi · ṙi ≡ Wi , (2.4)
tA i

where Wi(A→B) is the total work done on particle


P i during its motion from state A to state
B, Clearly P
the total kinetic energy is T = i Ti and the total work done on all particles is
W (A→B) = i Wi(A→B) . Eqn. 2.4 is known as the work-energy theorem. It says that
In the evolution of a mechanical system, the change in total kinetic energy is equal to the
total work done: T (B) − T (A) = W (A→B) .

11
12 CHAPTER 2. SYSTEMS OF PARTICLES

Figure 2.1: Two paths joining points A and B.

2.2 Conservative and Nonconservative Forces

For the sake of simplicity, consider a single particle with kinetic energy T = 12 mṙ 2 . The
work done on the particle during its mechanical evolution is
ZtB
(A→B)
W = dt F · v , (2.5)
tA

where v = ṙ. This is the most general expression for the work done. If the force F depends
only on the particle’s position r, we may write dr = v dt, and then
ZrB
(A→B)
W = dr · F (r) . (2.6)
rA

Consider now the force


F (r) = K1 y x̂ + K2 x ŷ , (2.7)
where K1,2 are constants. Let’s evaluate the work done along each of the two paths in fig.
2.1:
ZxB ZyB
W (I) = K1 dx yA + K2 dy xB = K1 yA (xB − xA ) + K2 xB (yB − yA ) (2.8)
xA yA
ZxB ZyB
W (II) = K1 dx yB + K2 dy xA = K1 yB (xB − xA ) + K2 xA (yB − yA ) . (2.9)
xA yA
2.2. CONSERVATIVE AND NONCONSERVATIVE FORCES 13

Note that in general W (I) 6= W (II) . Thus, if we start at point A, the kinetic energy at point
B will depend on the path taken, since the work done is path-dependent.
The difference between the work done along the two paths is
W (I) − W (II) = (K2 − K1 ) (xB − xA ) (yB − yA ) . (2.10)
Thus, we see that if K1 = K2 , the work is the same for the two paths. In fact, if K1 = K2 ,
the work would be path-independent, and would depend only on the endpoints. This is
true for any path, and not just piecewise linear paths of the type depicted in fig. 2.1. The
reason for this is Stokes’ theorem:
I Z
dℓ · F = dS n̂ · ∇ × F . (2.11)
∂C C
Here, C is a connected region in three-dimensional space, ∂C is mathematical notation for
the boundary of C, which is a closed path1 , dS is the scalar differential area element, n̂ is
the unit normal to that differential area element, and ∇ × F is the curl of F :
 
x̂ ŷ ẑ
∂ ∂ ∂ 
∇ × F = det  ∂x ∂y ∂z
Fx Fy Fz
     
∂Fz ∂Fy ∂Fx ∂Fz ∂Fy ∂Fx
= − x̂ + − ŷ + − ẑ . (2.12)
∂y ∂z ∂z ∂x ∂x ∂y
For the force under consideration, F (r) = K1 y x̂ + K2 x ŷ, the curl is
∇ × F = (K2 − K1 ) ẑ , (2.13)
which is a constant. The RHS of eqn. 2.11 is then simply proportional to the area enclosed
H
by C. When we compute the work difference in eqn. 2.10, we evaluate the integral dℓ · F
C
along the path γII−1 ◦ γI , which is to say path I followed by the inverse of path II. In this
case, n̂ = ẑ and the integral of n̂ · ∇ × F over the rectangle C is given by the RHS of eqn.
2.10.
When ∇ × F = 0 everywhere in space, we can always write F = −∇U , where U (r) is the
potential energy. Such forces are called conservative forces because the total energy of the
system, E = T + U , is then conserved during its motion. We can see this by evaluating the
work done,
ZrB
(A→B)
W = dr · F (r)
rA
ZrB
= − dr · ∇U
rA

= U (rA ) − U (rB ) . (2.14)


1
If C is multiply connected, then ∂C is a set of closed paths. For example, if C is an annulus, ∂C is two
circles, corresponding to the inner and outer boundaries of the annulus.
14 CHAPTER 2. SYSTEMS OF PARTICLES

The work-energy theorem then gives

T (B) − T (A) = U (rA ) − U (rB ) , (2.15)

which says
E (B) = T (B) + U (rB ) = T (A) + U (rA ) = E (A) . (2.16)
Thus, the total energy E = T + U is conserved.

2.2.1 Example : integrating F = −∇U

If ∇ × F = 0, we can compute U (r) by integrating, viz.


Zr
U (r) = U (0) − dr ′ · F (r ′ ) . (2.17)
0

The integral does not depend on the path chosen connecting 0 and r. For example, we can
take
(x,0,0)
Z (x,y,0)
Z (x,y,z)
Z
′ ′ ′ ′
U (x, y, z) = U (0, 0, 0) − dx Fx (x , 0, 0) − dy Fy (x, y , 0) − dz ′ Fz (x, y, z ′ ) . (2.18)
(0,0,0) (x,0,0) (z,y,0)

The constant U (0, 0, 0) is arbitrary and impossible to determine from F alone.


As an example, consider the force

F (r) = −ky x̂ − kx ŷ − 4bz 3 ẑ , (2.19)

where k and b are constants. We have





∂Fz ∂Fy
∇×F x = − =0 (2.20)
∂y ∂z
 
 ∂Fx ∂Fz
∇×F y = − =0 (2.21)
∂z ∂x
 
 ∂Fy ∂Fx
∇×F z = − =0, (2.22)
∂x ∂y
so ∇ × F = 0 and F must be expressible as F = −∇U . Integrating using eqn. 2.18, we
have
(x,0,0)
Z (x,y,0)
Z (x,y,z)
Z
3
U (x, y, z) = U (0, 0, 0) + dx′ k · 0 + dy ′ kxy ′ + dz ′ 4bz ′ (2.23)
(0,0,0) (x,0,0) (z,y,0)

= U (0, 0, 0) + kxy + bz 4 . (2.24)


2.3. CONSERVATIVE FORCES IN MANY PARTICLE SYSTEMS 15

Another approach is to integrate the partial differential equation ∇U = −F . This is in fact


three equations, and we shall need all of them to obtain the correct answer. We start with
the x̂-component,
∂U
= ky . (2.25)
∂x
Integrating, we obtain
U (x, y, z) = kxy + f (y, z) , (2.26)
where f (y, z) is at this point an arbitrary function of y and z. The important thing is that
it has no x-dependence, so ∂f /∂x = 0. Next, we have
∂U
= kx =⇒ U (x, y, z) = kxy + g(x, z) . (2.27)
∂y
Finally, the z-component integrates to yield
∂U
= 4bz 3 =⇒ U (x, y, z) = bz 4 + h(x, y) . (2.28)
∂z
We now equate the first two expressions:

kxy + f (y, z) = kxy + g(x, z) . (2.29)

Subtracting kxy from each side, we obtain the equation f (y, z) = g(x, z). Since the LHS is
independent of x and the RHS is independent of y, we must have

f (y, z) = g(x, z) = q(z) , (2.30)

where q(z) is some unknown function of z. But now we invoke the final equation, to obtain

bz 4 + h(x, y) = kxy + q(z) . (2.31)

The only possible solution is h(x, y) = C + kxy and q(z) = C + bz 4 , where C is a constant.
Therefore,
U (x, y, z) = C + kxy + bz 4 . (2.32)

Note that it would be very wrong to integrate ∂U/∂x = ky and obtain U (x, y, z) = kxy +
C ′ , where C ′ is a constant. As we’ve seen, the ‘constant of integration’ we obtain upon
integrating this first order PDE is in fact a function of y and z. The fact that f (y, z) carries
no explicit x dependence means that ∂f /∂x = 0, so by construction U = kxy + f (y, z) is a
solution to the PDE ∂U/∂x = ky, for any arbitrary function f (y, z).

2.3 Conservative Forces in Many Particle Systems

X
1 2
T = 2 mi ṙi (2.33)
i
X X 
U= V (ri ) + v |ri − rj | . (2.34)
i i<j
16 CHAPTER 2. SYSTEMS OF PARTICLES

Here, V (r) is the external (or one-body) potential, and v(r−r ′ ) is the interparticle potential,
which we assume to be central, depending only on the distance between any pair of particles.
The equations of motion are
mi r̈i = Fi(ext) + Fi(int) , (2.35)
with
∂V (ri )
Fi(ext) = − (2.36)
∂ri

X ∂v |ri − rj | X (int)
Fi(int) =− ≡ Fij . (2.37)
ri
j j

Here, Fij(int) is the force exerted on particle i by particle j:



(int) ∂v |ri − rj | ri − rj ′ 
Fij =− =− v |ri − rj | . (2.38)
∂ri |ri − rj |

Note that Fij(int) = −Fji(int) , otherwise known as Newton’s Third Law. It is convenient to
abbreviate rij ≡ ri − rj , in which case we may write the interparticle force as

Fij(int) = −r̂ij v ′ rij . (2.39)

2.4 Linear and Angular Momentum


P
Consider now the total momentum of the system, P = i pi . Its rate of change is

(int) (int)
Fij +Fji =0
X X (ext) z }| {
X
dP
= ṗi = Fi + Fij(int) = Ftot
(ext)
, (2.40)
dt
i i i6=j

since the sum over all internal forces cancels as a result of Newton’s Third Law. We write
X
P = mi ṙi = M Ṙ (2.41)
i
X
M= mi (total mass) (2.42)
i
P
i mi ri
R= P (center-of-mass) . (2.43)
i mi

Next, consider the total angular momentum,


X X
L= ri × pi = mi ri × ṙi . (2.44)
i i
2.4. LINEAR AND ANGULAR MOMENTUM 17

The rate of change of L is then

dL X 
= mi ṙi × ṙi + mi ri × r̈i
dt
i
X X
= ri × Fi(ext) + ri × Fij(int)
i i6=j
(int)
rij ×Fij =0
X zX }| {
(ext) 1 (int)
= ri × Fi + 2 (ri − rj ) × Fij
i i6=j
(ext)
= Ntot . (2.45)

Finally, it is useful to establish the result


X X 2
T = 1
2 mi ṙi2 = 12 M Ṙ2 + 1
2 mi ṙi − Ṙ , (2.46)
i i

which says that the kinetic energy may be written as a sum of two terms, those being the
kinetic energy of the center-of-mass motion, and the kinetic energy of the particles relative
to the center-of-mass.
Recall the “work-energy theorem” for conservative systems,

final
Z final
Z final
Z
0= dE = dT + dU
initial initial
initial (2.47)
XZ
= T (B) − T (A) − dri · Fi ,
i

which is to say
XZ
(B) (A)
∆T = T −T = dri · Fi = −∆U . (2.48)
i

In other words, the total energy E = T + U is conserved:


X X X 
1 2
E= 2 mi ṙi + V (ri ) + v |ri − rj | . (2.49)
i i i<j

Note that for continuous systems, we replace sums by integrals over a mass distribution,
viz. Z
X 
mi φ ri −→ d3r ρ(r) φ(r) , (2.50)
i

where ρ(r) is the mass density, and φ(r) is any function.

You might also like