Sei sulla pagina 1di 158

Classical Mechanics Notes (IIC)

2019
i

Table of Contents

Preface iv
Useful Constants vi

1 Newtonian Mechanics: Revision 1


Summary 1
1.1 Point Particle 1
1.2 Space and Time 1
1.2.1 The Relativity Principle 1
1.2.2 Inertial Reference Frames 1
1.3 Physical Quantities and Units 2
1.4 Newton’s Laws of Motion 2
1.4.1 Newton’s First Law 2
1.4.2 Newton’s Second and Third Laws 2
1.5 Motion of a Single Particle 3
1.6 General Motion 8
1.7 Conservation Laws 11
1.7.1 Single Particle Systems 11
1.7.2 Conservation of Mechanical Energy 13
1.7.3 Conditions for Existence of a Conservative Potential Function 16
1.7.4 Gradient 17
1.7.5 Separable Forces 17
1.8 Many Particle Systems 18

2 Oscillations 29
Summary 29
2.1 Introduction 29
2.2 Simple Pendulum 29
2.2.1 Motion Near Equilibrium; the Harmonic Oscillator 32
2.2.2 Solution of the Linear Harmonic Oscillator Equation 34
2.2.3 Phase Difference 45
2.3 Phase space 52
2.3.1 Simple Harmonic Motion: No Damping 53
2.3.2 The Under Damped Harmonic Oscillator 54
2.3.3 The Critically Damped Harmonic Oscillator 55
2.3.4 The Over damped Oscillator 56
2.4 Nonlinear Oscillator 56
2.4.1 Method of Successive Approximations 56
2.5 Chaotic Motion: An Introduction 58
2.5.1 Self­Limiting Oscillator 59
2.5.2 Driven Oscillator 60
Exercises 61

3 Central Forces and Celestial Mechanics 63


Summary 63
3.1 Introduction 63
3.2 Potential Energy in a General Central Field 63
3.3 Law of gravity 64
3.3.1 Potential Energy in a Gravitational Field. 64
3.3.2 Gravitational Force between a Uniform Spherical Shell and a Particle 66
3.3.3 Potential of a Uniform Spherical Shell 68
ii Table of Contents

3.3.4 Potential and Field of a Thin Ring 69


3.3.5 Angular Momentum 70
3.3.6 Law of Areas 71
3.3.7 Orbit of a Particle in a Central­Force Field 71
Example 73
3.3.8 Energy Equation of the Orbit 74
3.3.9 Kepler’s Laws of Planetary Motion 75
3.3.10 Orbits in an inverse­Square Field 76
3.3.11 Orbital Parameters from Initial Conditions 80
Example 81
3.3.12 Orbital Energies in the Inverse­Square Field 81
Example 82
3.3.13 Limits of the Radial Motion 83
3.3.14 Periodic Time of Orbital Motion 83
3.4 Motion in an Inverse­Square Repulsive Field 84
3.4.1 Scattering of Atomic Particles 84
Example 86

4 Noninertial Reference Frames 90


Summary 90
4.1 Introduction 90
4.2 Translation of the Coordinate System 90
4.3 Inertial Forces 91
4.4 Angular Velocity; Rate of Change of a Vector 92
4.5 General Motion of the Coordinate System 93
Examples 95
4.6 Dynamics of a Particle in a Rotating Coordinate System 96
4.7 Effects of the Earth’s Rotation 97
4.7.1 Static Effects. The Plumb Line 97
4.7.2 Dynamic Effects 99
4.7.3 The Foucault Pendulum 102

5 Lagrangian and Hamiltonian Mechanics 105


5.1 Generalised Coordinates and Degrees of Freedom 105
5.2 Generalised Forces 106
5.2.1 Generalised Forces for Conservative Systems 107
5.3 Lagrangian Equations of Motion 107
5.4 Some Applications of Lagrange’s Equations 109
5.4.1 Harmonic Oscillator 109
5.4.2 Single Particle in a Central Field 110
5.4.3 Particle Sliding on a Movable Inclined Plane 110
5.5 Generalised Momenta 112
5.6 Hamilton’s Variational Principle and Euler’s Equations 112
5.7 Homogeneous Functions and the Kinetic Energy 114
Euler’s Theorem 114
Homogeneous Functions and the Kinetic Energy 115
5.8 Conservation Laws 115
5.8.1 Energy Conservation 115
5.8.2 Momentum Conservation 116
Conservation of angular momentum 117
5.8.3 Cyclic or Ignorable Coordinates 117
Example 118
5.9 The Hamiltonian Function 118
Example 1 118
Example 2 119
iii

5.10 Hamilton’s Equations 120


Example 1 121
Example 2 121
5.11 Poisson Brackets 121

A Coordinate Systems and Kinematic Variables 129


A.1 Coordinate Systems and Vectors 129
1.1.1 Cartesian Coordinate System 129
1.1.2 Change of Coordinate System 130
1.1.3 Spherical Coordinate System 131
1.1.4 Cylindrical Coordinates 131
A.2 Position Vector of a Particle 132
A.3 Velocity Vector 132
A.4 Acceleration Vector 133
A.5 Vector Integration 133
A.6 Relative Displacement, Velocity and Acceleration 134
A.7 Derivatives of Products of Vectors 134
A.8 Tangential and Normal Components of Acceleration 134
A.9 Velocity and Acceleration in Plane Polar Coordinates 136
A.10 Velocity and Acceleration in Spherical and Cylindrical Coordinates 138
A.11 Angular Velocity 139

B Conics 143
B.1 Cartesian Form 143
B.2 Polar Form 144

C 146
C.1 Elliptic Integrals 146
C.2 Curvilinear coordinates 146
C.3 Tensors 147
3.3.1 Eigenvalues and Diagonalisation of a Symmetric Tensor 148
3.3.2 Evaluation of Determinants 149

Index 150
iv Preface

Preface
Classical mechanics is a very old subject. It is one of the major sub­fields of physics, and is concerned with
the set of physical laws describing the motion of bodies under the action of a system of forces. The study
of the motion of bodies is an ancient one, making classical mechanics one of the oldest and largest subjects
in science, engineering and technology. Its formal structure was proposed by Newton in the seventeenth
century, while the mathematical structure was further developed by Lagrange in the late eighteenth century
and by Hamilton in the nineteenth.
Classical mechanics describes the motion of macroscopic objects, from projectiles to parts of machinery,
as well as astronomical objects, such as spacecraft, planets, stars, and galaxies. Besides this, many
specializations within the subject deal with gases, liquids, solids, and other specific sub­topics. Classical
mechanics provides extremely accurate results as long as the domain of study is restricted to large objects
and the speeds involved do not approach the speed of light. When the objects being dealt with become
sufficiently small, it becomes necessary to introduce the other major sub­field of mechanics, quantum
mechanics, which reconciles the macroscopic laws of physics with the atomic nature of matter and handles
the wave­particle duality of atoms and molecules. In the case of high velocity objects approaching the
speed of light, classical mechanics is enhanced by special relativity. General relativity unifies special
relativity with Newton’s law of universal gravitation, allowing physicists to handle gravitation at a deeper
level.
The term classical mechanics was coined in the early 20th century to describe the system of physics
begun by Isaac Newton and many contemporary 17th century natural philosophers, building upon the
earlier astronomical theories of Johannes Kepler, which in turn were based on the precise observations of
Tycho Brahe and the studies of terrestrial projectile motion of Galileo.
The initial stage in the development of classical mechanics is often referred to as Newtonian mechanics,
and is associated with the physical concepts employed by and the mathematical methods invented
by Newton himself, in parallel with Leibniz, and others. Later, more abstract and general methods
were developed, leading to reformulations of classical mechanics known as Lagrangian mechanics and
Hamiltonian mechanics. These advances were largely made in the 18th and 19th centuries, and they extend
substantially beyond Newton’s work, particularly through their use of analytical mechanics. Ultimately,
the mathematics developed for these were central to the creation of quantum mechanics.
Recent research in classical mechanics has been in the application of modern mathematical tools
to understand, in particular, the transition between regular and chaotic behaviour. Many fascinating
discoveries about the earth and space in the past decades, since the launching of artificial satellites, have
been due to the direct application of classical mechanics.
Every scientific theory is based on some hypotheses which are suggested by observations, but represent
an idealisation of them. The theory is tested against experiment and if persistent discrepancies are found
attempts have to be made to modify the theory. Gradually the hypotheses acquire the status of ’laws of
nature’. Since its formulation by Newton, the range of validity of the three Laws of Motion on which
classical mechanics is based, has been extended, but for the small scale phenomena of atomic, nuclear
and particle physics it has been superseded by quantum mechanics and for phenomena involving speeds
approaching the speed of light, by relativity.
Quantum physics and relativistic mechanics are extensions of classical mechanics in the sense that they
reproduce the results of classical mechanics in the appropriate limiting cases. These theories do not
imply that classical mechanics has lost its value, but rather reinforces our confidence in the correctness of
classical mechanics in its own vast range of applicability.

Textbooks:
The notes for the course are sufficient for in introduction to the material covered. We’ll work with the
notes as reference, but there are many good textbooks available. Most textbooks on Classical Mechanics
cover more material than what we’ll discuss in the course. If you intend to continue with physics, it is a
good idea to have a a book on Classical Mechanics to refer to. When you browse in the library you may
find a textbook that appeals to you, but following are recommended:
Recommended textbook: Analytical Mechanics, Fowles and Cassidy, Saunders College Publishing
Recommended textbook: Classical Mechanics, T W Kibble
v

A very useful textbook is ’Classical Mechanics’ by H Goldstein. The treatment in this book is very
detailed and it is probably best to get some background elsewhere first before you tackle this standard
reference book on classical mechanics.

DPJ November 2019


vi Preface

Useful Constants

Physical Constants

universal gravitational constant N m kg

Defined Standard Values

standard gravitational acceleration ms


normal atmospheric pressure Pa

Properties of the Earth


mass kg
m s
radius (polar) km
(equatorial) km
(mean) km
semi­major axis of orbit km
eccentricity of orbit
orbital period (sidereal year) s
mean orbital velocity km s
surface escape velocity km s
rotational angular velocity s

Properties of the Sun and Moon

mass of Sun kg

m s
mass of Moon kg
semi­major axis of lunar orbit km
lunar orbital period (sidereal month) s
1

Chapter 1
Newtonian Mechanics: Revision
Summary
In this chapter we shall revise some basic concepts in classical mechanics. We shall re­examine Newton’s
Laws of Motion on which all of classical mechanics is based. We shall show, using simple vector
calculus, how these laws, formulated for point particles, can be applied to systems composed of very large
numbers of particles. We shall formulate expressions for kinetic energy, linear and angular momentum for
extended bodies. We shall examine what is meant by a conservative force and derive the conditions for the
conservation of mechanical energy and linear and angular momentum.

1.1 Point Particle


A very useful concept in mechanics is the idealised particle or point particle, an entity with mass but no
spatial extent. We shall make extensive use of this concept in discussing the principles of mechanics. A
rigid body can then be considered as a collection of particles (atoms) that do not change their relative
positions. So for a rigid body we can determine where every particle of the system is if we know where at
least three particles are at any one time.

1.2 Space and Time


We assume that space and time are continuous. It is meaningful to say that an event took place at a specific
point in space at a specific time and that there are universal standards of time and length which allow
meaningful comparison of measurement by observers in different places.
In classical mechanics we assume that there is a universal time scale (two synchronized clocks will
always agree), that the geometry of space is Euclidean, and that there is in principle no limit to the accuracy
with which measurements can be made.

1.2.1 The Relativity Principle


In Newtonian mechanics we normally consider bodies as falling towards the earth due to gravitational
attraction, rather than falling towards some other fixed point in space. So position has meaning relative to a
chosen point. Velocity also only has relative significance. Given two bodies, it is in principle not possible
to decide which is at rest and which is moving. This fundamental statement is the principle of relativity.
Acceleration, on the other hand, retains an absolute meaning since it is experimentally possible to
distinguish between uniform and accelerated motion. Two un­accelerated observers must arrive at the
same result when performing identical experiments while an accelerated observer may come to a different
conclusion. The relativity principle states that two un­accelerated observers are equivalent as discussed in
Chapter 4.

1.2.2 Inertial Reference Frames


To specify positions and times, each observer may choose a zero on a time scale and an origin for a set of
three coordinate axes. Collectively these are known as a frame of reference. Any event can be specified by
three position coordinates and a time.
In view of the relativity principle, the frames of reference used by un­accelerated observers are equivalent
and are called frames. The laws of physics expressed in inertial frames (in terms of, for example,
2 Chapter 1 Newtonian Mechanics: Revision

x, y, z and t) are identical but will differ from the laws expressed in an accelerated frame of reference (see
Chapter 4).
For all practical purposes we can take an inertial frame to be one that moves with uniform velocity
relative to the ’fixed’ stars in which the centre of the solar system moves with uniform velocity. Frequently
it is even enough to take the earth as an inertial reference frame.

1.3 Physical Quantities and Units


A physical quantity is a quantity that can be measured relative to some chosen physical unit. In classical
mechanics we need only three physical quantities to describe all phenomena. The choice of base
quantities, in terms of which we measure all physical quantities, is arbitrary. We use mass, length and
time. (Another choice is length, force and time.) All other quantities can be expressed in terms of the
three chosen base quantities. We use the SI (Systeme International) or MKS system of units: kilogram
(mass), metre (length) and second (time).

1.4 Newton’s Laws of Motion

I. Every body continues in its state of rest or of uniform motion in a straight line, unless it is
compelled by a force to change that state.

II. The change of the state of motion is proportional to the resultant applied force and takes place in
the direction of the resultant force.

III. To every action there is always an equal and opposite reaction or, the mutual actions of two
bodies are always equal and oppositely directed.

1.4.1 Newton’s First Law


In a sense Newton’s first law (NI) amounts to a definition of an inertial reference system. Such a system
is one in which Newton’s first law holds.

1.4.2 Newton’s Second and Third Laws


Each body can be characterized by a scalar constant, its inertial mass m , which is a measure of how
difficult it is to change the state of rest or uniform motion of the object. We refer to the resistance of an
object to change its state of motion as the inertia of the object. Its linear momentum is defined to be its
(inertial) mass, times velocity,

The equation of motion that specifies how the body will move, Newton’s second law (NII), is then

and if the mass of the body remains constant,

where is the total force acting on the body and the acceleration .
If we denote the force acting the body due to the by
Section 1.5 Motion of a Single Particle 3

where
Newton’s third law (NIII) states that
(1.1)

The quantities mass and force can have meaning only if they can be measured. Inertial masses of two
bodies can, in principle, be compared by subjecting them to equal forces, but this is only possible to do if
we know that the two forces are equal in the first place. One case where this will be true follows from
Newton’s third law. If two bodies are isolated from all other matter, then according to Newton II and III

(1.2)
so that the accelerations are oppositely directed and inversely proportional to the masses. If the first body
is a standard reference body, the mass of the second body can be defined in terms of the first body. Since
acceleration can be measured, force can be measured once mass is defined.

Note that Newton’s laws of motion are formulated for point particles. One of the appealing properties of
this formulation is that it is easily extended to include many particle systems.

1.5 Motion of a Single Particle


In general the forces acting on a particle are functions of position, time, and velocity. Given the position
and velocity of a system of particles at a given time, we have to determine all the forces acting at the time
and then determine the positions and velocities using equations of motion, or other approaches, a short
interval later. In general this cannot be done in terms of known analytic functions.

In this section we do some revision and introduce a few new concepts and mathematical tricks.
1.5.0.1 Kinetic Energy, Potential Energy and Turning Points in one Dimension

Newton’s second law of motion can be written as

This vector equation can be reduced to three scalar equations. In Cartesian coordinates:

(1.3)

It is convenient to introduce a shorthand notation for time derivitaves. We do this simply because it saves
time. For each time derivatie we add a dot over the variable we take the time derivative of. So we can write

and

(1.4)

If there will be no acceleration in the and directions. Let the initial velocity be in the
­direction, .
Explain, using equations, in which direction the particle will move if
4 Chapter 1 Newtonian Mechanics: Revision

In this case we have

where the quantity is the kinetic energy of the particle. We have

and

(1.5)

The integral is the work done on the particle by the force This is the work­energy theorem in
one dimension, which states that the work done on a system is equal to the change in kinetic energy. Up to
now we have not said anything about the nature of the force, so the work energy theorem is a completely
general statement applicable to any type of force. If a force depends on position only, then we can
a function as

constant (1.6)

and hence

(1.7)

The function is called the potential energy function; it is defined to within an arbitrary additive
constant. Consequently, if we choose the reference point as we may write

(1.8)

constant (1.9)

where is called the total (mechanical) energy. In other words, for one dimensional motion, if the
applied force is a function of position only, the sum of the kinetic and potential energies remains constant
throughout the motion of the particle. The force is said to be conservative. Non­conservative forces, for
which a potential function cannot be defined, include friction.
Why is the constant in Equation (1.6) zero if we choose the reference point at
The motion of the particle moving in one dimension under the action of a conservative force can be
obtained by solving

(1.10)
Section 1.5 Motion of a Single Particle 5

which can be written as

(1.11)

thus giving as a function of


Prove that Equation (1.11) is correct.
From equation (1.10) we see that the velocity is real only if the potential is less or equal to the total
energy. In other words the particle is confined to regions where

as illustrated in figure 1.

Figure 1.1. Potential energy showing the allowed region and the turning points for a given value of
the total energy

1.5.0.2 Motion in a Resisting Medium. Terminal Velocity

An object moving through a  fluid is subject to viscous resistance. If the resistance is proportional to the
velocity, we can write it as where is a positive constant.
Explain why must be positive.
Take the ­axis to be positive upward and consider a particle moving vertically through air.
6 Chapter 1 Newtonian Mechanics: Revision

( x , y0 , z 0 )


r

Figure 1.2.Particle moves along dashed line. and are constant.

If we assume that the resistance is linear in , then, with

(1.12)

NII becomes

(1.13)

(1.14)

When we take the scalar product with respect to the equation of motion becomes a scalar equation,

(1.15)

If we ignore the dependence of the gravitational acceleration on height, we have a velocity dependent
force When we try to determine the vertical velocity component
directly from the equation of motion, we have to evaluate the integral

(1.16)

where is an integration constant. It is immediately obvious that there is a problem. In order to determine
in this way, we already have to know what is as a function of . There are two variables in the equation
motion (1.15), and For this example we can separate the two variables by rearranging the equation of
motion:
(1.17)

and then calculate an expression for as a function of time:

(1.18)

(1.19)
Section 1.5 Motion of a Single Particle 7

This equation can be inverted to solve for

(1.20)

(1.21)

where is the terminal velocity and the characteristic time. After sufficient time
the velocity approaches the limiting value, The terminal velocity is that velocity for which the
magnitude of the resisting force is equal to the weight of the object and the total force is then equal to zero.
The displacement as a function of time follows from

Note that the analysis is valid for a particle moving vertically up or down. If the particle is initially
moving upwards, will be positive and if it is initially moving downwards, will be negative.
Discuss the last statement about
Let us now consider the posibility that the viscous resistance is proportional to the speed squared,
resistance where is a positive constant. The vertical component of the equation of motion
now becomes

The minus sign in ( ) refers to motion in the positive direction (remember we took up as positive) and
the plus sign refers to motion in the negative direction.
Explain why the assignment of the signs, +­, is correct.
Now we have

(rising)

(falling) (1.22)

where Solving for

(rising)

(falling). (1.23)

Show that and ( Hint: and What is the sign of and

for positive and negative ? )


If the body is released from rest at time then and

Once again the terminal speed is attained after a few characteristic times have elapsed.
8 Chapter 1 Newtonian Mechanics: Revision

x 
ri ' O'


 ro '
ri

Figure 1.3. is the position vector of prticle relative to some fixed point

1.6 General Motion


We have been a bit casual in the interpretation and application of Newtons laws in the previous section.
What do we really mean with the ’motion of a body’ and the application of Newton’s second law to
describe the motion of a body?
Let us consider a system made up of particles, with particle of mass and at position . We can
think of the particles as the atoms in the system, or the mass in an infinitely small volume of the body.
Applying Newton’s laws to each particle, the total force on particle can be expressed as

,ext (1.24)

where ,ext is the resultant force exerted on particle by everything outside our system of particles and
the force that particle (part of our system) exerts on particle For convenience of notation we set
since the particle does not exert a force on itself. The equation of motion for particle is

,ext (1.25)

Summing over all particles we get

,ext (1.26)

The double summation in (1.26) can be expressed as

from NIII

(1.27)

where the first step follows from interchanging the dummy variables and This tells us that the internal
Section 1.6 General Motion 9

forces do not contribute to the equation of motion and equation (1.26) simplifies, so that we have.

,ext (1.28)

In general the system of particles can change shape as it moves in space. So what do we mean by the
’motion of the body’. Let us try to find some point associated with the body that will allow us to talk about
the motion of the body in an unambiguous way.
Let be the origin of a point fixed in space and a point ’fixed’ to the body. Then the position of
particle can be written as
(1.29)
(see figure 1.3) where is the position of the point relative to the origin of our fixed coordinate system
and the position of particle relative to The equation of motion for our system can now be written as

,ext tot,ext

tot,ext (1.30)


ri '

ri CM


rcm
z

Figure 1.4.Centre of mass of a system of particles: cm

We now the centre of mass position vector of the system:

cm (1.31)

If we set cm we can write

cm tot,ext (1.32)

where is the total mass of the system. The vector is now measured relative to the centre
of mass, so the vector is the centre of mass position measured relative to the centre of mass and
hence is zero. The equation of motion for the system of particles can therefore be written as

cm tot,ext (1.33)
10 Chapter 1 Newtonian Mechanics: Revision

If we talk about the motion of the centre of mass, we have a unique way of describing the translational
motion of the body. Equation (1.33) tells us that the motion of the centre of mass can be described as if we
had a point particle with mass at cm and a force equal to the total external force acting on the system
acting on the particle at cm This is not the whole story, the body can change shape and orientation relative
to the centre of mass, so we need to have a way of describing this as well.
Note that if we interpret Newtons laws of motion as definitions for the motion of point particles we can
interpret the laws of motion uniquely, but not the motion of extended objects, as we see from the discussion
above for the motion of the centre of mass.
We can get a description of the motion relative to the centre of mass by once again starting with the
equation of motion for particle

,ext (1.34)

Taking the cross product with and summing over all particles, we get

,ext (1.35)

With

cm

as shown in figure 4, we get

cm cm

cm cm cm

cm

cm cm (1.36)

where the last step follows from the definition of the centre of mass. The first term on the right of equation
(1.35) can be written as

,ext cm tot,ext ,ext (1.37)

and the second term can be written as

cm

from NIII

from NIII. (1.38)


Section 1.7 Conservation Laws 11

Since is the position of particle relative to particle it follows that equation (1.38) will be zero
for central internal forces, in other words if the force that particle exerts on particle is parallel to the line
between particles and From equations (1.33), (1.35), 1.36), (1.37) and (1.38) it follows that for central
internal forces,

cm cm cm tot,ext (1.39)

,ext (1.40)

Note that equation (1.39) follows directly form equation (1.33).


By definition
(1.41)
is the linear momentum of a particle with mass and velocity We define the angular momentum of a
particle taken about a point as
(1.42)
where is the position of the particle relative to the point about which the angular momentum is taken.
The torque, or moment, of a force is given by

(1.43)

where is the position of the point of application of the force relative to the point about which the torque
is taken. From (1.42) it follows that

(1.44)

and hence equations (1.39) and (1.40) can be expressed as

cm,O cm,O tot,ext tot,O (1.45)

cm cm tot,cm (1.46)

This gives us a second equation for the motion of the centre of mass and an equation to describe the motion
of asystem of particles relative to the centre of mass.
Note that in equations (1.45) and (1.46), we indicate the point (O or cm) relative to which the vector
quantities are measured.
In order to describe the general motion of a system, we have to use the equation of motion for the centre
of mass equation (1.33) and then we still need to describe the motion relative to the centre of mass for
which we can use equation (1.46). These equations are general, they are applicable to rigid bodies as well
as deformable bodies.

1.7 Conservation Laws


Many of the important conclusions of mechanics can be derived from conservation laws, which indicate
under what conditions various mechanical quantities are constant in time.

1.7.1 Single Particle Systems


12 Chapter 1 Newtonian Mechanics: Revision

1.7.1.1 Conservation of Linear Momentum

For an isolated particle of mass and linear momentum

tot . (1.47)

If the particle is isolated, tot and

(1.48)
constant (1.49)

which states that the momentum of an isolated particle is conserved. This is a specific case of a more
general rule that the total linear momentum of an isolated system of particles is conserved. Note that
Equation (1.47) is a vector equation and it follows that the component of the momentum perpendicular to a
force is always conserved.
Explain the last statement. Use equations to assist you in explaining it.

1.7.1.2 Conservation of Angular Momentum

The angular momentum of a particle about a point is defined as

where is the position vector of the particle measured from The moment of a force or torque, about
is

We can write the torque as

(1.50)

where the last identity follows from the fact that the cross product of two parallel vectors is zero. As a
consequence we can write

(1.51)

Note that both and depend on the position of the reference point It follows that

The Angular Momentum of a particle about a point is conserved if the total external torque about the
same point is zero.

From equation 1.45 it follows that the angular momentum of a extended body taken about its centre of
mass is constant if the total torque about the centre of mass is zero. This is special case of the more general
situation: The angular momentum of a system of particles about a point is conserved if the total
external torque about the same point is zero. See exercises for proof.

Note that equations (1.47) and (1.51) are vector equations. The conservation laws apply to each
component of the vector expressions. Therefore, if the force or torque is zero in a given direction, the
linear or angular momentum in that direction is conserved, even when the total force or torque is not zero.
Use equations to confirm the last statement.
Section 1.7 Conservation Laws 13

1.7.2 Conservation of Mechanical Energy


Consider the work done by an external force on a particle going from point to point By definition
the work done on the particle is given be the line integral

(1.52)

If is the total force acting on the particle and we assume a constant mass, then, since

it follows that

(1.53)

and we have

(1.54)

We recall that the scalar quantity is called the kinetic energy of the particle The line integral
represents the work done on the particle by the force as the particle moves along a given path of motion.
The work principle (or work­energy theorem), equation (1.54), states that the work done on the
particle is equal to the change in kinetic energy.
Note that is equal to the distance travelled multiplied by the component of the force in
the direction of displacement. In general the value of the integral will depend on the path of
integration.
When the force is a function of position only, it is said to define a static force.
Explain what a line integral is. (See for example the discussion at
http://en.wikipedia.org/wiki/Line_integral of a line integral.)
A force is said to be conservative if all closed line integrals of the force are zero:

(1.55)

(The O over the integral sign indicates a closed path, a path that starts and ends at the same point.)
Physically it is clear that a system cannot be conservative if there are dissipative forces such as friction
present. This is easy to see for friction, since friction is always negative and no closed line integral

friction (1.56)

can vanish.
14 Chapter 1 Newtonian Mechanics: Revision

For a conservative force, the work done by the force is independent of the path and we can define the
change in potential energy as the negative of the work done in moving a particle from to

(1.57)

is called the potential or potential energy. It is a unique function of position to within an additive
constant.

Prove that as defined in equation(1.57) is a well defined function of position.


Since the work done on a conservative system is equal to the negative of the change in potential energy,
and the change in kinetic energy is equal to the work done on the system, the work energy theorem for a
conservative system implies that

or
constant, (1.58)
or combining (1.57) and (1.54) we have for a conservative system the result

(1.59)

The sum of the kinetic and potential energies, is called the mechanical energy of the system.
The last two equations are statements of the

Energy Conservation Theorem: If the forces acting on a particle are conservative, the total mechanical
energy of the system, T+V, is conserved.

1.7.2.1 Relation between the potential and the force for a conservative system

For a conservative force we can show that

(1.60a)

Here we used the notation for the gradient of a scalar function ,

grad (1.61)

and where is shorthand for the gradient operator, a vector operator,

It is not surprising that equation (1.60a) can be derived in several ways from our discussions above. For
example, the conservation of mechanical energy for a conservative system can be written as

where the time rate of change of kinetic energy can be expressed as

(1.62a)

The change in kinetic energy in a small time interval is


Section 1.7 Conservation Laws 15

since We now have

(1.63)

(Note that we derived this relationship before ­ here we derived the work­energy equation for a conservative
force force from the conservation of mechanical energy.) Using the chain rule for differentiation, we can
write the time rate of change of the potential energy of a conservative system as

We can write this in terms of the gradient of

grad (1.64)

We therefore have

(1.65)

Differentiating (1.58) with respect to time we therefore get with the help of equations (1.62a) and (1.65)
that

Since this must hold for any velocity of the particle, we require

In terms of components

Note that we can add any constant to in equation (1.60a) without affecting the results. This tells us
that the zero of is arbitrary.
We can derive equation (1.60a) in an alternative way. If equation (1.55) holds, the work must be
independent of the path of integration between points to and it is equal to the negative of the change
in the potential energy,

(1.66)

( is an example of a scalar field line integral). We can re­write this as

(1.67)

where is the component of in the direction of Now consider the work done along the differential
path length where and are close, then

or

(1.68)
16 Chapter 1 Newtonian Mechanics: Revision

This is equivalent to (1.60a), .


For example, if describes a three dimensional harmonic oscillator

1.7.3 Conditions for Existence of a Conservative Potential Function


If a potential energy function exists, we know that, in Cartesian coordinates,

(1.69)

If we take partial derivatives of the force components we find that, for example,

Since the order of differentiation can be reversed, and it follows that

These are necessary conditions for a potential function to exist. So if is conservative we find that a
convenient way of expressing the criteria for a force field to be conservative is in terms of the curl of is
zero:

(1.70)

(If you are not familiar with the curl notation, you can easily see where it comes from. The curl operator,
replaces the first vector in the cross product of two vectors and the expression on
the right­hand­side of (1.70) follows naturally.)
This is also a sufficient condition as we can see from the following argument. By definition the line
integral about any closed loop is zero for a conservative force:

From Stokes’ theorem it follows that we can write

(1.71)

where is an open two sided surface bounded by a closed non­intersecting curve The directed element
of area is with a small element of area and a unit vector normal to directed towards
the same side of at every point. (The direction of is related to the direction of the integration along the
path using the right­hand rule.) If the right­hand of (1.71) is zero for any closed path, then all surface
integrals must be zero. This is only possible if vanishes everywhere. (You will
come across Stokes theorem again in the second semester in the Electromagnetism course where you will
study it in more detail. If you are not familiar with Stokes Theorem, accept the result for the moment.)
In summary, a sufficient and necessary condition for a force to be conservative is that the curl of the
force vanishes:
(1.72)
Section 1.7 Conservation Laws 17

1.7.4 Gradient
Let be a function of , the gradient of is sometimes written as grad
Let and be two neighbouring points, then

Now if the distance is fixed, then the scalar product is a maximum when is in the direction of
Hence points in the direction in which changes most rapidly. Its magnitude is the rate of change in
this direction.
Physically the relation

means that a particle is urged to move in the direction of maximum decrease in potential energy.

1.7.5 Separable Forces


If it is possible to choose a coordinate system such that the components of the force can be expressed in
terms of the respective coordinates only,

the force is separable. The curl of such a force is zero and the field is conservative. The equations of
motion are very simple because each component can be treated separately.
Show that for a separable force in a Cartesian coordinate system and in the curvilinear
coordinate systems discussed in Appendix A.

1.7.5.1 Motion of a Projectile in a Uniform Gravitational Field

Negligible Air Resistance

Choose the ­axis to be vertical with the positive direction upwards. The force acting on the particle is
which is separable and hence conservative. The potential function can be defined as

where the horizontal plane defined by is used as the reference plane. For the simple
force only depends on the vertical displacement from the reference plane. The total
mechanical energy is conserved, so

If then

The equation of motion reads

which can be integrated directly to give


18 Chapter 1 Newtonian Mechanics: Revision

Further integration yields

where and are the initial velocity and position respectively. For the components of this
equation are

(1.73)

We can eliminate time from the and equations to get

This means that the trajectory lies in the plane defined by and the axis. Eliminating the time
from the and equations we have

where and Hence the path is a parabola in the plane defined by and the ­axis.
For convenience we can rotate the coordinate system about the ­axis to new ­ and ­axes with the
­axis in the direction so that the motion takes place in the ­plane.

Figure 1.5. Path of a projectile moving in three dimensions.

1.8 Many Particle Systems


When considering a many particle system we have to distinguish between internal forces and external
forces. For a system of particles the equation of motion (Newton’s Second Law) can be expressed as

ext (1.74)

where is the force that particle exerts on particle and ext the resultant force on particle from all
Section 1.8 Many Particle Systems 19

the forces originating from outside the system. Summing (1.74) over all particles yields

ext (1.75)

Figure 1.6. Centre of mass of a system of particles.

For a system of particle of masses . . ., whose position vectors are . . ., the


centre of mass of the system as defined before is the point cm given by

cm (1.76)

where is the total mass of the system. Assuming that all the internal forces satisfy
Newton’s third law and lie along the line that joins the two particles, we find that in the double summation
in equation (1.75) for each there is also a which is equal and opposite:
from NIII. Consequently this double sum vanishes. The equation of motion for the system of particles
therefore reduces to

ext

cm (1.77)

where

cm

cm

The acceleration of the centre of mass of a system of particle is the same as that of a single particle with
mass equal to the total mass of the system and acted upon by the vector sum of the external forces.

Note that purely internal forces have no effect on the motion of the centre of mass.
20 Chapter 1 Newtonian Mechanics: Revision

The linear momentum of the system is the vector sum of the linear momenta of the individual particles:

tot

From (1.76) it follows by differentiating that

tot cm

It follows that

tot ext

tot
(1.78)

1.8.0.2 Linear Momentum

A consequence of (1.77) is that the total linear momentum of a system of particles,

constant, (1.79)

provided the total external force on the system is zero. It also follows that if the total force in some
direction is zero, the linear momentum in that direction is conserved.
1.8.0.3 Angular Momentum

The total angular momentum of a system of particles is obtained by summing over the angular momenta of
all the particles.

Therefore

ext (1.80)

The last term can be considered as a sum of pairs of the form

using the equality of action and reaction. But we made the assumption that the internal forces lie along the
line that joins the particles, hence is parallel to and hence

It follows that

ext

ext (1.81)

The Total Angular Momentum is constant in time if the total external applied torque is zero.
Section 1.8 Many Particle Systems 21

It is worth to note that equation (1.81) is a vector equation and hence the angular momentum is con­
served component by component: can be conserved even if and or is not conserved.

Vectors involved in shift of reference point for angular momentum.

The angular momentum is defined with respect to a specific point. Let cm be the centre of
mass coordinates then

cm cm

cm cm cm cm

cm cm cm cm (1.82)

The last two terms both contain the term which is equal to the total mass time the centre of
mass measured with respect to the position of the centre of mass and is hence zero. The angular momentum
can thus be written as

cm cm (1.83)

where cm is the total linear momentum of the system.


In other words equation (1.83) states that the total angular momentum of the system about a point
is the angular momentum of a point particle of mass located at the centre of mass plus the angular
momentum of motion about the centre of mass. It also means that

cm cm

ext,cm ext,O (1.84)

1.8.0.4 Energy

The total work done in moving a system of particles from an initial configuration 1, to a final configuration
22 Chapter 1 Newtonian Mechanics: Revision

2 is

ext (1.85)

Here the second term is the internal work done on the system. The equations of motion can again the used
to reduce these integrals to

(1.86)

and hence the work can still be written as

Transforming to centre of mass coordinates, we can write

cm cm

cm cm cm

We have already argued that the last term is zero, hence

cm cm (1.87)

cm cm (1.88)

The kinetic energy, like the angular momentum, has two parts: the kinetic energy of the centre of mass and
the kinetic energy of motion relative to the centre of mass.
If the external forces can be derived from the gradient of a potential field, the first term in (1.85) can be
written as

ext

(1.89)

If the internal forces are also conservative, and can be obtained from a potential function. In order
to satisfy NIII and the condition that the internal forces point along the line joining two particles, the
Section 1.8 Many Particle Systems 23

potential function can only depend on the distance between the two particles:

(1.90)

The two forces are then automatically equal in magnitude and opposite in direction,and point along the
line joining the particles:

(1.91)

(1.92)

Show that the provious two equations are correct.


When all the forces are conservative, the second term in (1.85) can be written as the sum over pairs in
the form

(1.93)

If the difference vector is and stands for the gradient with respect to then

and

so that each term in the pair has the form

The total work arising from internal forces then reduces to

(1.94)

Here we set The factor appears since in summing over both and each term in a pair appears
twice. So if both internal and external forces can be derived from potentials, it is possible to define a
potential energy of the system

(1.95)

such that the total energy


(1.96)
is conserved. For a rigid body the internal potential remains constant and can be set equal to
zero.
24 Chapter 1 Newtonian Mechanics: Revision

Note that equation (1.85) can also be expressed as a term for the work done on the centre of mass plus a
term for the work done relative to the centre of mass: We write

cm

cm

where is measured relative to the centre of mass.

cm

cm

tot ext cm (1.97)

The first term gives the work done on the centre of mass by the total external force acting on the system,
while the second term is the work done by all forces moving particles relative to the centre of mass. The
same force can contribute to both terms.
Consider the term

For a rigid body, each point remains at a fixed distance from the centre of mass and can only move on
a spherical surface relative to the centre of mass. If we consider the work done over a very short period
of time, we only have to think of an infinitesimal displacement about the instantaneous axis of rotation.
Choose an instantaneous cylindrical coordinate system with origin at the centre of mass and the ­axis
along the instantaneous axis of rotation. Then

For a rigid body is the same for each particle. We have

Since

and

it follows that

total
Section 1.8 Many Particle Systems 25

This tells us that the work done by all the forces acting on the rigid body in rotating it about an axis through
the centre of mass is equal to the component of the total torque parallel to the axis of rotation time the
angle through which the body is rotated.
26 Chapter 1 Newtonian Mechanics: Revision

Exercises
1. Explain why it is necessary to know the positions of at least three particles that make up a rigid body in
order to be able to specify where all the other particles in the body are.
2. Explain what is meant by ’force’.
3. Consider a particle
(a) Show that the same acceleration is measured for in two reference frames moving at constant
velocity with respect to each other.
(b) Show that different accelerations are measured for in two reference frames accelerating with
respect to each other.
4. Imagine that you measure the force (weight) that the gravitational field of the earth exerts on an object
on the surface of the earth using a spring balance. Now imagine that you are in the Space Shuttle
accelerating away from the earth. Do not worry about why the Shuttle is accelerating, and assume that
it is moving away perpendicular to the surface of the earth and only consider what you read on the
spring balance.
(a) What force will you now read on the spring balance? Is it the weight that you read?
(b) Now imagine that you are not aware that you are accelerating with the respect to the earth when
you do the measurement, and you pull on the object to accelerate it relative to the Shuttle.
i. What acceleration will you measure relative to the Shuttle?
ii. What is the acceleration with respect to the earth?
(c) What are the physical forces exerted on the object?
(d) Since we know from NII that is it valid for an earth bound observer and an astronaut in
the Shuttle to equate the total force exerted on the object and the mass times acceleration as
measured in their reference frames?
(e) How can an observer in the Shuttle tell that there is a force from outside the Shuttle acting on the
object?
(f) Do your conclusions tell you that there is a problem with Newton’s second law of motion?
(g) Explain what is meant by an inertial reference frame.
5. Explain whether it is possible to isolate particles.
6. A particle of mass is initially at rest. A constant force is applied at time After a time
the force suddenly doubles and remains constant after that. Find the speed of the particle and the total
displacement at time
[ ]
7. A particle of mass is initially at rest. A constant force acts on the particle for a time The
force then increases linearly with time such that after an additional interval the force is equal to
Show that the total distance the particle travels in the time is
8. Find the velocity and the position as functions of for a particle of mass which starts from rest at
at time and subject to the following forces:
(a)
(b)
(c)
(d)
[ ]
9. A block slides on a horizontal surface which has been lubricated with a heavy oil such that the block
experiences a viscous resistance in the opposite direction to the velocity and that varies with speed
according to

If the initial velocity is at time


(a) find as a function of time Discuss the allowed values of
(b) For find the displacement as a function of time
Exercises 27

(c) Show that for the block will not travel further than
[ ]
10. For a particle moving in a straight line, find the velocity as a function of the displacement for a
particle of mass which starts from rest at and subject to the following forces:
(a) ,
(b)
(c)
[ ]
11. A force acting on a particle of mass varies with distance according to the power law

(a) Find the potential energy function.


(b) If and at time fine as a function of
(c) Determine the turning points of the motion.
[ (n odd),

(n even) ]
12. A particle of mass is released from rest a distance from a fixed origin of force that attracts the
particle according to

Show that the time required for the particle to reach the origin of the force is
[ ]
13. The force acting on a particle of mass moving along the ­axis is given by

where is a constant. The particle passes through the origin with velocity at
time Find as a function of
[ ]
14. A block is projected up an inclined plane with initial speed If the inclination of the plane is (angle
the plane makes with horizontal) and the coefficient of sliding friction between the block and plane is
find the total time required for the block to return to the point of projection. Is it possible to find a
value of for which the block will just come to rest as it returns to the initial point?
[ No ­ explain why ]
15. Consider a block pushed backwards and forwards in a straight line on a table with friction. Draw the
applied force acceleration diagram.
16. Prove equation (1.50).
17. Discuss the derivation of Eq. (1.68) and explain why Eq. (1.68) is equivalent to Eq. (1.60a).
18. Show that the position of the centre of mass of a system of particles is determined by the distribution of
the particles and does not depend on the coordinate system used to calculate it.
19. Determine which of the following forces are conservative:
(a)
(b)
(c)
(d)
(e)
28 Chapter 1 Newtonian Mechanics: Revision

[ conservative; conservative; not; not; conservative ]


20. Find the value of the constant for which the following forces are conservative.
(a)
(b)
[ ]
21. Find the force for each of the following potential energy functions.
(a)
(b)
(c)
(d)
[
]
22. Find the potential functions for the forces in exercise 19 that are conservative. In each case use two
methods.
[ ]
23. Show that the inverse­square law force in three dimensions is conservative.
[Hint: Calculate the curl in spherical coordinates ­ see appendix C]
24. A particle of mass moves in a force field given by the potential If the particle passes
through the origin with a speed what will the speed be when the particle passes through the point

[ ]
25. Show that the motion of a projectile of mass subjected to a retarding force linear in the velocity,
in a uniform gravitational field is described by the equation

if the projectile is launched from the origin at time with velocity The vector is given w.r.t. a
Cartesian coordinate system with the ­axis pointing vertically upwards
[ ]
26. Consider a system of particles. Prove that the total momentum of the system is conserved if the total
external force on the system is zero. Remember to take all internal forces into account.
27. Prove that the angular momentum of a system of particles about a point is conserved if the total
external torque about the same point is zero. Remember to take all internal forces and torques into
account.
28. Explain why we can write Eq. (1.90) in the form given.
29. A uniform cylinder of mass and radius rolls in a horizontal surface.
(a) Show that the total torque about the centre of mass due to friction is parallel to the axis of
symmetry. [Hint. Use cylindrical coordinates and take into account that there will be friction at
every contact point.]
(b) i. Show that the work done by friction can be expressed as a sum of two terms: the work done
on the centre of mass plus the work done in rotating the cylinder. Give an explicit expression
for the work done relative to the centre of mass when the cylinder rotates through an angle
about its axis of symmetry.
ii. Show that if the cylinder rolls without slipping the total work done by friction is zero and if it
slips finite work is done by friction.
29

Chapter 2
Oscillations
Summary
In this chapter we shall examine ways of describing simple periodic motion, a type of motion found in
many situations. We shall start with a Simple Pendulum, and then simplify the system to the simple
harmonic oscillator. We shall solve the second order linear differential equation that we derive from
Newton’s second law of motion and then successively introduce a damping term and a driving force and
find solutions to the new differential equations that we derive. We shall interpret the solutions to the
differential equations and describe the physical motion of the damped and driven harmonic oscillators.
More complicated motion can be described by introducing Fourier transforms and Green’s functions,
mathematical tools that you will recognize from a course on solving differential equations, but introduced
here to find solutions to differential equations that describe physical systems. In the final sections we shall
briey look at chaotic motion by examining simple models of the nonlinear oscillator, the self­limiting
oscillator and the nonlinear driven oscillator.

2.1 Introduction
All around us we see things that move with periodic motion. The moon rotates about the earth and the
earth rotates about the sun periodically. The string of a guitar oscillates periodically when plucked. Even
things that we cannot see, like the air molecules in a wind instrument vibrate in a regular manner. The
essential feature here is periodicity.
The simplest form of periodic motion is simple harmonic motion.

2.2 Simple Pendulum

Figure 2.1. Simple pendulum.


30 Chapter 2 Oscillations

Consider a bob of mass supported by a thin rigid rod of length of negligible mass ( )
suspended at the other end on a frictionless pivot so that the bob is constrained to swing in a plane.
Assuming uniform gravitational acceleration, in plane polar coordinates, the gravitational force on the bob
is
(2.1)
The tension in the rod is radial, so the equation of motion is

(2.2)

but is constant so

(2.3)
We can solve the equation of motion (or try to). Equation (2.3) is a second order, non­linear differential
equation and the solution is not straight forward. Let us approach this problem in a slightly different way,
we shall return to the differential equation later: The velocity of the bob can be written as
(2.4)
(2.5)
The last line follows since the rod is rigid so that The kinetic energy becomes

where the distance moved by the bob is The tangential motion of the bob is described by

(2.6)

and the tangential component of the force can be expressed as

(2.7)

The tangential equation can therefore be treated as one dimensional. The force is a function of
position only, therefore a potential energy function exists :

(2.8)

If we chose then

(2.9)
1 The radial force ( cos ¡ ( )) ^ does not contribute to a change in energy since the only contribution to the line integral
along the curve comes from the tangential part.

¢ = ¡ =¡ ( cos ¡ ) ^¡ sin ^ ^

= ¡ ¡ sin
31

V
E1

½ mvtop2
2mgl
E1

 

Figure 2.2.Potential energy of pendulum bob.

Recall that the work energy theorem states that the work done on a system is equal to the change in
kinetic energy. So

or

(2.10)

where is the total mechanical energy for the system. We could have written down equation (2.10)
directly since we know that the system is conservative.
Suppose that initially and the bob is given a push that starts it moving with a speed . Then the
total energy is Now if ( in Figure 2.2), the motion is confined between two
angles where

(2.11)

These are the points where the kinetic energy vanishes, so the pendulum bob is instantaneously at rest.
The motion is thus an oscillation of amplitude
Note that

(2.12)

If the bob is pushed so hard that ( in Figure 2.2) then the kinetic energy will never vanish
and bob will still have positive kinetic energy when it reaches the upward vertical. With the speed at the
top top

top (2.13)

In this case the motion is a continuous revolution.


32 Chapter 2 Oscillations

s
­A V=0 A

Figure 2.3. Potential energy near the minumum for the harmonic oscillator, .

2.2.1 Motion Near Equilibrium; the Harmonic Oscillator


A particle can be in equilibrium only if the total force on it is zero. For a conservative force, from (1.7),
this only happens if the gradient of the potential energy curve is zero. Let us consider motion in one
dimension only. If we are interested in small displacements, we can expand the potential energy curve
around where is assumed to have a local minimum, in a Maclaurin­Taylor series:

(2.14)

We can choose since the potential energy is defined only to within a constant. At equilibrium
so the first non­zero term is the quadratic term. Thus for we can
approximate

(2.15)

where We call this the harmonic approximation. The motion near almost any point of
equilibrium can be described by a potential function of this form. The potential energy curve is a parabola
at this level of approximation.
For the harmonic approximation gives a parabola as shown in Fig. 3. The energy must be greater
or equal to zero. There are two points where From equation (2.10) it follows
that and the motion is an oscillation between these points.
If Fig. 4, the curve for the harmonic approximation is an inverted parabola and two kinds of
motion is possible.
33

E1

V=0 s

E2

Figure 2.4. Potential energy near maximum, .

If (measured from top of potential at the particle may approach the minimum up to a
point where it comes to rest momentarily and then reverses direction ( in Figure 4). If ( in
Figure 4) the particle has sufficient energy to overcome the barrier and will never come to rest.
The force corresponding to the potential is

(2.16)

in the harmonic approximation. It is attractive towards the equilibrium position if the potential is close to a
minimum and repulsive if the potential is close to a maximum, and respectively.
For convenience of notation we can set
We can use conservation of energy directly to solve for and integrate to obtain:

(2.17)

(2.18)

We can try to invert this equation to get as a function of An alternative, which we shall explore, is to
solve the equation of motion for the linear harmonic oscillator.

Exercise 1 Try to invert equation (2.18) to get an expression for as a function of


34 Chapter 2 Oscillations

For the simple pendulum that we used as introduction to this chapter, we have

There is no analytic solution for this equation.

2.2.2 Solution of the Linear Harmonic Oscillator Equation


In the harmonic approximation the equation of motion near an extremum in the potential energy becomes

The equation

(2.19)

is a homogeneous second order linear differential equation.

differential equation expresses a relationship amongst a function and its derivatives

function and its derivatives appear in equation to first order only; coefficients
multiplying the function and its derivatives are constants or functions of
linear
the independent variable

order highest derivative is second order

homogeneous sum of function and its derivatives is zero:

A second order linear differential equation has two independent solutions. Such equations have the
property that their solutions satisfy the superposition principle: if and are solutions, then so is
any linear combination

where and are constants. So if and are solutions of Eq. (2.19), then we have

If and are linearly independent ( this is the general solution. The general solution
of a second order differential equation therefore has two arbitrary constants that are fixed by the initial
conditions.
Another way to see that we need two constants to solve Eq. (2.19) is to notice that it is of second order
so its solution may be obtained from integrating twice, hence we get two arbitrary constants of integration,
so we have to get two independent solutions and

Case 1: so has a maximum at (Fig. 4)

(2.20)

The functions and satisfy this differential equation. The general solution is
(2.21)
35

A small increase in time leads to an exponential increase in the magnitude of The equilibrium is
unstable. In addition, as the distance from the extremum increases the quadratic approximation to the
potential energy ceases to be valid.

Exercise 2 Discuss equation (2.21) in relation to Figure (4). Explain why the equation is consistent with
the conclusions about the motion of a particle near a local maximum in the potential energy that we arrived
at before.

Exercise 3 Find values for and in equation (2.21) that correspond to the two situations illustrated in
figure (4).

Case 2: so has a minimum at (Fig. 3) Simple harmonic oscillator.


The equation of motion can be written as

(2.22)

where is the angular frequency


The functions and are solutions of this equation. In general

(2.23)

where and are determined by the initial conditions. If at the particle is at with velocity ,
then
(2.24)

Exercise 4 Prove equation (2.24).

An alternative way of writing (2.23) is

(2.25)

where
(2.26)

Exercise 5 Prove equation (2.26). Recall that

The constant is the . It defines the limits between which the particle oscillates,
maximum The motion of the particle is a periodic oscillation of period

The is the number of oscillations per unit time


36 Chapter 2 Oscillations

This discussion applies to the motion of a particle near a stable equilibrium point of potential energy
function. For sufficiently small displacements in the vicinity of a local minimum in the potential energy
any system behaves like a simple harmonic oscillator. The second derivative of the potential energy
determines the frequency or period, which consequently depends on the details of the system, but the
general motion does not depend on the specific details of the system.

A T

x t

-A

Figure 2.5. Simple harmonic motion.

2.2.2.1 Energy in Harmonic Motion

The total mechanical energy is given by

(2.27)

The velocity as a function of displacement is then

For to be real, must lie between This implies that the maximum value of the amplitude ,
is

(2.28)

Exercise 6 Discuss in your own words the derivation of equation (2.28) for the magnitude of the ampli­
tude.

The maximum value of occurs for so we have

(2.29)

and

2.2.2.2 Complex Representation

It is frequently convenient to describe periodic motion in terms of complex numbers. This allows us to
37

treat the solution for all in the same manner. Let us take as a trial solution of (2.19), then

(2.30)
Consequently

where and is the angular frequency as before. The general solution is then

(2.31)
Since must be real, and must be complex conjugates :

Exercise 7 Prove that

Now since an alternate way of writing the solution is


(2.32)
(2.33)
or
(2.34)
(2.35)
as we had before (see equation (2.25)), but now valid for all

Exercise 8 Prove equations (2.33) and (2.35).

The period is given by

and the frequency by

The period and frequency become imaginary for


2.2.2.3 Damped Harmonic Motion

Let us consider the situation where and our pendulum experiences a linear drag force such as would
be exerted by a viscous medium and only consider the effect on small oscillations. If is the displacement
from equilibrium then the restoring force is and the retarding force is where is a constant. The
differential equation of motion is

and by re­arranging terms


(2.36)
38 Chapter 2 Oscillations

Exercise 9 Show that the scalar component of the restoring force for an harmonic oscillator is equal to

Equation (2.36) is another second order linear differential equation, so we need two independent
solutions. Let us take as a trial solution once more, then

(2.37)

for all This will be the case if satisfies the equation


(2.38)

The two roots are given by

(2.39)

where and Recall that There are three physically distinct cases:

I
II
III

I. Over damping. The two roots are real and negative, let them be and The general solution
is
(2.40)
The motion is non­oscillatory and the displacement decays in a exponential manner.

Exercise 10 What are the expressions for and Show that can overshoot the equilibrium point
once only, and discuss when this can happen in .terms of the initial position and velocity.

II. In the case of critical damping, the two roots are equal and so (2.39) is not the general solution since
there is really only one solution, . To get the full solution we return to (2.36). For equal roots we can
factor this equation as

Make the substitution which gives


39

So

(2.41)

This is the general solution (two integration constants) and is also non­oscillating. The displacement
decays to zero asymptotically with time.

Exercise 11 What is the value of Show that can overshoot the equilibrium point once only, and
discuss when this can happen in .terms of the initial position and velocity.

III. If the resistance constant is small enough so that is complex.

(2.42)

The general solution is

(2.43)

Since must be real, we must have

(2.44)

The solution represents an oscillation with exponentially decreasing amplitude and angular
frequency The natural frequency is less than the frequency of the undamped oscillator
40 Chapter 2 Oscillations

x = A e­ t
A

x = ­A e­ t

Figure 2.6.Displacement versus time for the underdamped harmonic oscillator.

For weak damping, that is for we can expand as a function of and get the approximate
relation

(2.45)

2.2.2.4 Energy Considerations

The total energy of the oscillator at any time is the sum of the kinetic and potential energies;

(2.46)

Differentiate this expression with respect to time:

(2.47)

From the equation of motion (2.36) the expression in brackets is

consequently

(2.48)

This is always negative and represents the rate at which the energy is dissipated by friction.
2.2.2.5 Quality Factor

The rate of energy loss of a weakly damped harmonic oscillator can be characterised by the quality factor
It is defined to be times the energy stored in the oscillator divided by the energy lost in a single
period of oscillation . If the oscillator is weakly damped, the energy lost will be small and will
41

be large. The rate of energy dissipation is given by Equation (2.48), From (2.44)

The energy lost during a single cycle is

If we change the variable of integration to then and the integral over the period
transforms to an integral from to The value of the integral does not depend on
the initial phase .

(2.49)

For the exponential factor does not change much over a cycle and we can take it from
inside the integral. Therefore

(2.50)

where we used the relations and


We can define a time constant and write the magnitude of the energy loss per cycle as

(2.51)

where the energy stored at any time is

(2.52)

(See exercise 13.) The energy remaining in the oscillator dies away exponentially with time constant We
see that the quality factor is

(2.53)

For weak damping the period of oscillation is much smaller than the time constant which
characterises the energy loss rate of the oscillator. is large under these circumstances.
Piano string 3000
Crystal in a digital watch 10
Atomic clock 10
Neutron star 10
Excited Fe nucleus 3
2.2.2.6 Forced Harmonic Motion: Resonance
42 Chapter 2 Oscillations

Consider a linearly damped harmonic oscillator that is driven by an external harmonic force ext , that is a
force that varies sinusoidally with time.

ext

It is convenient to write the force as

ext

and demand that the real and imaginary parts are satisfied by the resulting differential equation.
The total force is the sum of three forces and the equation of motion becomes

ext

ext (2.54)

The solution of this non­homogeneous linear differential equation is given by the sum of two parts: The
first is the solution of the homogeneous equation which we solved before. The second
is any particular solution that satisfies equation 2.54. The homogeneous equation represents an oscillation
which decays exponentially and is called the term (see equation (2.44)). It is reasonable to
expect that the solution that depends on the applied force will also have a sinusoidal time dependence. So
we try a solution of the form

which must satisfy

at all times. We must have

(2.55)

where and as before. We can also write this expression as

(2.56)

or

Equating real and imaginary parts

(2.57)
(2.58)

where the or Upon dividing the second equation by the first,


we get

(2.59)

By squaring both sides of (2.57) and (2.58) and adding we get

(2.60)
43

hence

(2.61)

We have found a particular solution of (2.54). The real part is simply

The general solution is obtained by adding to this particular solution the general solution of the
homogeneous equation, namely (2.40), (2.41) or (2.44). In the case where the damping is less than critical
( we obtain

(2.62)

Here and are arbitrary constants fixed by the initial conditions while and are given by (2.61) and
(2.59). The second term, the transient, dies away exponentially with time.
The amplitude (2.61) assumes a maximum value at a frequency , the resonance frequency. To find
the resonance frequency we determine and solve for :

(2.63)

Exercise 12 Show that the resonance frequency

For weak damping, or, equivalently, the resonance frequency is very close to
the frequency of the free oscillator with no damping.

|A|


c = 1/8 m
c = 1/16 m



   

Figure 2.7. Amplitude versus driving frequency.


44 Chapter 2 Oscillations

If we expand the right­hand side of (2.63) as a series in and retain the first two terms we get

(2.64)

Let then we can write the approximate natural frequency of the free running oscillator with
damping (2.45) as

and the approximate resonance frequency as

The natural frequency of the damped oscillator lies between the frequency of the free oscillator and the
resonance frequency of the forced damped oscillator.
The steady­state (when the transient term has died away) amplitude at the resonant frequency,
from (2.61) and (2.63) is

(2.65)

In the case of weak damping, we can neglect and we have

(2.66)

So the amplitude of the induced oscillation becomes very large if the damping constant is small, and
conversely. In a mechanical system it may or may not be desirable to have large amplitudes. For the
mountings of electrical motors, for example, the stiffness of rubber or spring mounts have to chosen to
ensure that the resulting resonant frequency is far from the running frequency of the motor.
The sharpness of the resonance peak is of interest. For weak damping, we can make the
following approximation for

With these approximations the amplitude for small and can be written as

(2.67)

Since

and, using equation (2.65),


45

When

(2.68)

This means that is a measure of the width of the resonance curve.


The total energy of the free oscillator is

(2.69)

so at the energy is half that at resonance. Thus is called the half­width of the resonance
peak. Note the inverse relationship, the smaller the half­width the higher the peak (see equation (2.66)).
Another way of designating the sharpness of the resonance peak is the quality factor of the resonant
system.

(2.70)

As for the damped oscillator we can show that this is times the energy stored in the oscillator divided
by the energy lost in a single period of oscillation (in the steady state). For weak damping

The width at the half­energy points is approximately

or, since

giving the fractional width of the resonance peak.


Electrically driven quartz crystal oscillators are used to control the frequency of radio broadcasting
stations. The of the quartz crystals in such applications is of the order of

2.2.3 Phase Difference


The difference in phase between the applied driving force and the steady­state response is given by
equation. (2.59):

and is plotted in Figure 8. The phase difference is small for small so that the response is in phase with
the driving force. At the phase, and thus the response is o out of phase with the driving
o
force near resonance. For large values of the value of approaches hence the system is out of
phase with the driving force.
In the low frequency limit, the amplitude becomes

(2.71)
46 Chapter 2 Oscillations


c = 1/16 m
c = 1/8 m


 


Figure 2.8. Phase angle versus driving frequency.

In other words for a spring, the spring constant and not the mass attached to the spring and not the
friction determines the response. In the low frequency limit the mass is pushed back and forth by a force
acting against the retarding force of the spring.
The ratio of the maximum possible amplitude and the amplitude for the same strength of driving force
when in the case of weak damping is

max
(2.72)

The result is the of the oscillator. For no friction the resultant amplitude at resonance can be enormous.
In order to understand why the response at resonance can be very large, think about pushing a child on a
swing. For a large amplitude, we want to transfer the maximum amount of energy to the swing. It makes
sense to push the swing while it is moving and to push in the direction of motion, or in the direction of
the velocity. A small force applied judiciously can result in a large amplitude. The maximum transfer of
power occurs when is a maximum. So we want the driving force and the velocity to be in phase for
the maximum transfer of power. Since the velocity is o out of phase with the displacement, we want the
driving force and the displacement to be o out of phase.
Let us verify this conclusion. The steady­state solution is, from equation. (2.62),

(2.73)

for a driving force


(2.74)
At resonance, for small damping so that

(2.75)

The velocity at resonance is

so that indeed the driving force and velocity is in phase at resonance.


2.2.3.1 Beats

Consider the case when Then there is no damping of the transient solution and the general solution
47

Figure 2.9. Beats with

is a sum of two periodic solutions

where and Near resonance

Since changes little over the period so the first term can be thought of as a time
dependent amplitude. Then

(2.76)

where

(2.77)

It follows that

(2.78)

The time dependent amplitude varies between and This phenomenon is called
beats.
2.2.3.2 Function Periodic with period T: Fourier sum

If a function is periodic with period

(2.79)
48 Chapter 2 Oscillations

then can be written as an infinite sum, a Fourier sum,

(2.80)

where

(2.81)

If is real,
It follows from (2.80) that

(2.82)

Also

(2.83)

but

(2.84)

where is the Kronecker delta: if and otherwise. Hence

(2.85)

2.2.3.3 General Periodic Force

A general periodic force can be written as a Fourier series as discussed in the previous paragraph:

ext (2.86)

Since ext is real, a given term, and its complex conjugate is always present in the
sum. Because of the linearity of the equations

ext (2.87)

the solution can be written as


transient (2.88)

where is related to and precisely as in equation (2.55):

(2.89)

where This may be verified by direct substitution in the differential equation

ext
49

2.2.3.4 Square Well Force

Consider a force of the form

F(t)

Figure 2.10. Square well force.

Then

It follows that

Note that for all even values of The amplitudes are (see equation (2.89))

odd

and the position of the oscillator is given by

transient.

If is close to one of the values one pair of amplitudes and will be much larger than the rest,
and the oscillation will be almost simple harmonic motion at this frequency. In general the amplitudes
decrease rapidly in magnitude with increasing values of so we can get a good approximation to by
retaining only a few terms of the series.
50 Chapter 2 Oscillations

2.2.3.5 Impulsive Forces; the Green’s Function Method

In many physical situations, such as during a collision, a very large force acts for a short time. Consider a
force that acts for a time The resulting change in momentum is

(2.90)

The integral is called the impulse, In the extreme case when as while
the position remains constant, the momentum of the particle changes discontinuously. This is a good
description of a sudden blow, for example. In reality, of course, the force always remains finite.
Consider an under damped simple harmonic oscillator. The equation of motion is

The displacement can be expressed as At a time the oscillator


experiences a blow of impulse If the duration of the blow is very short, it is still at
immediately after the blow (the particle cannot move a finite distance in an infinitely short interval) while
its velocity is The last expression follows from

At times equations of the form will be solutions of the equation


of motion and so will

The velocity after the impulse is

Taking the limit

(2.91)

(recall that )

Take ( the other possibility will change the sign of then

(2.92)

Hence

(2.93)
51

If the oscillator is subjected to a series of blows of impulse at time its position may be found by
adding together the corresponding set of solutions of the form (2.93) to obtain

transient (2.94)

where

(2.95)

The function is called the Green’s function for the oscillator. It represents the response of the
oscillator to an instantaneous blow delivered at a time (Recall that the transient is a solution of the
homogeneous equation.) To prove that transient is a solution to of the equation of
motion, we have to show that it satisfies the equation of motion at all times, apart from the times when the
impulse is delivered when the derivatives may be a bit dodgy, and that the initial conditions (position and
velocity) at the times are correctly given. (See exercises.)
The Green’s function for the undamped oscillator is

undamped

undamped (2.96)

This function can be derived as before, or by taking the limit when


We are now in a position to find a solution for an arbitrary force driving a linearly damped oscillator

The method is in a sense complementary to Fourier’s approach. We express as a sum of impulsive


forces. If we divide the time into small intervals then the impulse delivered between and will
be approximately when Then

transient

transient (2.97)

The transient term is the solution before the driving force is a applied, before that time we set .
From the definition is zero for or so the integrand is zero for This
solution is particularly convenient if the applied force is known numerically.
Let us check the validity of equation (2.97). The solution must satisfy

(2.98)

For and

(2.99)

The only problem occurs at It is sufficient to integrate over an infinitesimal interval around so

transient. (2.100)

The integral
52 Chapter 2 Oscillations

We know that for

(2.101)

and for

Recall that we integrate over For a small interval we can approximate

(2.102)

So in the interval as a function of is a step function that changes by at


We have

(2.103)

In the interval we can approximate the derivative of as a step function in We now


approximate the step function by linear interpolation in the interval so that we can take the
second derivative of

(2.104)

so that

(2.105)

We now perform the integration before we take the limit Therefore

(2.106)

This proves that equation (2.97) is a solution.


Note: The derivative of a step function is a Dirac delta function.

2.3 Phase space


A physical system that does not dissipate energy remains in motion or stationary for all time. If it undergoes
periodic motion it repeats its configuration each cycle. A system that does dissipate energy eventually
comes to rest. The evolution of a system can be graphically illustrated in a special space called phase
space. The coordinates of the phase space of a system of particles include the instantaneous positions and
velocities of all the particles. Visualising phase space for a system of particles is impossible, but the phase
space of a single particle whose motion is restricted to one dimension consists of all the points in a plane
whose horizontal component is the position of the particle and the vertical component is the velocity
The ’position’ of the particle in its phase space is given by its ’coordinates’ The future state of
motion of such a particle is completely specified if the position and velocity are known simultaneously,
since we can apply Newton’s laws of motion to the situation and determine where the particle goes to at
any subsequent time. We can picture the evolution of the motion by plotting the coordinates in phase space.
Section 2.3 Phase space 53

2.3.1 Simple Harmonic Motion: No Damping


The solutions for the position and velocity are

(2.107)

We can eliminate from these parametric equations:

(2.108)

or

(2.109)

This is the equation for an ellipse whose semimajor axis is and its semiminor axis is
.
x
1.5

0.5

-3 -2 -1 1 2 3
x
-0.5

-1

-1.5

Figure 2.11. Phase space plot for simple harmonic oscillator s

Note that the phase­path trajectories in the phase space plot never intersect. The existence of a point
common to two different trajectories would imply that two different trajectories could evolve from a single
set of conditions ( at some time This cannot happen since Newton’s Laws of motion
completely determine a unique future state of motion. Also note that the trajectories form closed paths. In
other words the motion repeats itself. This is a consequence of the conservation of mechanical energy of
the simple harmonic oscillator. Substituting and into (2.109) leads to

(2.110)

which is equivalent to

(2.111)

the energy equation (2.27) for the harmonic oscillator. Each closed phase space trajectory corresponds to
some definite, conserved energy.
54 Chapter 2 Oscillations

2.3.2 The Under Damped Harmonic Oscillator


The solutions for and for an under damped harmonic oscillator are

(2.112)

It is more difficult to eliminate from these equations. Substitute and into


(2.112) to obtain

(2.113)
(2.114)

0.4

0.2

x
-0.6 -0.4 -0.2 0.2 0.4 0.6 0.8

-0.2

Figure 2.12. Modified phase space plot for simple underdamped harmonic oscillator. Solid line
dashed line s

Now apply to (2.114) to get

(2.115)

We can square this expression to obtain

(2.116)

or

(2.117)

This expression is identical in form to (2.109). Since is a linear combination of and the ensemble
of points represent a modified phase space. The trajectory of the oscillator in this space is an ellipse
whose major and minor axes decrease exponentially with time. The trajectory starts off with a maximum
value of and then spirals inward towards the origin.
In the case of weak damping then

(2.118)

Hence (2.117) becomes

(2.119)

This is identical in form to (2.117) and hence the trajectory seen in the ­ plane is almost identical to the
modified phase space trajectory for weak damping.
Section 2.3 Phase space 55

Upon substituting and into (2.119) we get

(2.120)

Comparing this expression to (2.52) we see that (2.120) represents the total energy remaining in the
oscillator at any time The energy of the oscillator dies away exponentially and the spiral nature of its
phase space trajectory reects this fact.

2.3.3 The Critically Damped Harmonic Oscillator

y
0.5

0.4

0.3

0.2

0.1
x
0.2 0.4 0.6 0.8 1 1.2
-0.1

-0.2

Figure 2.13. Modified phase space plot for a critially damped simple harmonic oscillator. s

For the critically damped harmonic oscillator

(2.121)

or

(2.122)

The last equation shows that the phase space trajectory should approach a straight line whose slope is equal
to
56 Chapter 2 Oscillations

2.3.4 The Over damped Oscillator

y
x
0.2 0.4 0.6 0.8 1
-0.02

-0.04

-0.06

-0.08

-0.1

Figure 2.14. Phase space plot for overdamped simple harmonic oscillator. s s

Over damping occurs when the damping parameter is larger than the angular frequency The solution
for the motion is, Equation (2.40):

(2.123)

where

As in the case of critical damping the phase space trajectory approaches zero along a straight line.
However, two lines are possible. Let the motion start from rest at displacement Then

(2.124)

A little algebra shows that there are two possible linear combinations of and

(2.125)
(2.126)

The right­hand side of each of these equations die out with time and thus the phase space asymptotes are
given by the pairs of straight lines:

(2.127)
(2.128)

The phase space path of motion usually approach zero along the asymptote with slope

2.4 Nonlinear Oscillator

2.4.1 Method of Successive Approximations


Up to now we have only considered restoring forces that are linear in the displacement (harmonic
approximation). When a system is disturbed from its equilibrium position, the restoring force may be
nonlinear in the displacement. In the nonlinear case the restoring force can be expressed as

(2.129)
Section 2.4 Nonlinear Oscillator 57

in which the function represents the departure from linearity. The equation of motion can now be
written as
(2.130)
Note that is necessarily quadratic or of higher order in Solving this kind of equation normally
requires some method of approximation. We shall find an approximate solution for the cubic case using
the method of successive approximations.
Consider
(2.131)
Upon division by and using and we have

(2.132)

If then
(2.133)
Let us try a trial solution
(2.134)
where can take any value. We may expect this solution to be reasonable for small Inserting our trial
solution into the equation of motion gives

(2.135)

Upon rearranging terms we get

(2.136)

Excluding the trivial case where we see that our trial solution does not exactly satisfy the
differential equation. However, an approximation that is valid for small is obtained by setting the
expression in brackets equal to zero:

(2.137)

This is an approximation for the frequency of the freely running nonlinear oscillator. It depends on the
amplitude of the oscillation and in the limit of small amplitudes we get back the frequency of the linear
oscillator.
To obtain a better solution we must take into account the dangling term in (2.136) involving the third
harmonic We can try a second trial function of the form .

(2.138)

This leads to

terms in and higher multiples of (2.139)

Setting the first term in brackets equal to zero gives the same values for found before. Setting the second
term in brackets to zero yields a value for

(2.140)
58 Chapter 2 Oscillations

where we have assumed that the term in the denominator involving is small enough to neglect. Our
second solution can be expressed as

(2.141)

The process can be repeated to get higher order approximations.


This method is crude, but it illustrates two essential features of free oscillations. The frequency
of oscillation is amplitude dependent and the oscillation is not sinusoidal, but can be considered as a
superposition of harmonics.

2.5 Chaotic Motion: An Introduction


When the equations of motion are nonlinear an arbitrary sum of two solutions will in general not be a
solution. Consider the equation discussed in the previous section,

(2.142)

and assume that and are solutions of this equation. If we substitute the linear combination

into this equation, we get for the left­hand side of equation (2.142)

(2.143)

If we substitute the linear combination into the right­hand side of (2.142) and equate it to (2.143) we have

(2.144)

With a little algebra this leads to

(2.145)

Since and are solutions of the equation that varies with time, the only way that this equation can
be satisfied for all times is when This shows that an arbitrary linear combination of the two
solutions is not a solution in general.
The essence of chaotic motion of a nonlinear system is erratic and unpredictable behaviour. It is a
commonly occurring phenomenon and occurs, for example, in a driven oscillator ­ driven beyond its
linear regime. It occurs in the weather and in the convective fl ow of motion of a heated 
fluid, in electrical
circuits and in objects bound to our solar system. Chaotic behaviour in such systems manifests itself in
non repetitive behaviour. The oscillations are bounded, but each oscillation looks like nothing in the past.
The motion is still deterministic in nature. Given the forces that the system are subject to, it will evolve
according to the laws describing the motion. We just may not be able to calculate the evolution with any
degree of certainty.
Section 2.5 Chaotic Motion: An Introduction 59

2.5.1 Self-Limiting Oscillator

0.5

x
-1.5 -1 -0.5 0.5 1

-0.5

-1

Figure 2.15. Phase space plot for the self­limiting oscillator. Damping factor and starting values
­ solid curve ­ dashed curve.

As a special case let us look at an oscillator subject to a nonlinear damping force


whose equation of motion is

(2.146)

Let then we can write

(2.147)

Note that the nonlinear damping term is negative for all points inside the ellipse given by

(2.148)

It is zero for all points on the ellipse and positive for all points outside the ellipse. So no matter what the
current state of the oscillator (described by its current position in phase space), it will be driven toward
states whose phase space points lie on the ellipse. In other words, no matter how the motion is started, the
oscillator will ultimately vibrate with simple harmonic motion. Its behaviour is said to be self­limiting and
this ellipse in phase space is called its limit cycle.
A complete solution must normally be carried out numerically. For illustrative purposes we can simplify
things by setting and equal to one. This transforms the elliptical limit cycle to a circular one of
unit radius and the angular frequencies are then measured in units of The equation of motion then
becomes
(2.149)
This can be solved as a coupled set of linear differential equations:

(2.150)

These equations have an analytic solution:

(2.151)

as can be verified by direct substitution. This solution represents the final limiting motion on the unit circle
60 Chapter 2 Oscillations

2.5.2 Driven Oscillator


Chaotic behaviour can be induced by driving the oscillator harmonically. The equation of motion is then
(2.152)
where is the amplitude of the external driving force and is the driving frequency. This can be solved
as a system of three linear differential equations:

(2.153)
In Figure 16 the driven oscillator was started at . In other words it was started
from rest from position outside the limit cycle, and given a maximum ”shove” in the positive
direction. Only the first few cycles are shown in Figure 16. The oscillations fail to lock into the limit cycle
though they are confined to a region close to the limit cycle after a few cycles. The motion is deterministic,
by which we mean that given the same set of initial conditions, the same results will be obtained. It is
chaotic, in the sense that even very slight changes in the initial conditions or parameters will lead to
completely different phase space trajectories. Hence the motion of any chaotic system is unpredictable
since any values used to make predictions have a limited numerical precision, which therefore limits the
knowledge of the outcome of the motion.

y
1

0.5

-1.5 -1 -0.5 0.5 1


x
-0.5

-1

Figure 2.16. Phase space plot for the driven self­limiting oscillator. s
Starting values
Exercises 61

Exercises
1. Explain why the rod in Section (2.2) only exerts a radial force on the bob.
2. Prove Eqs. (2.24) and (2.25).
3. Prove the condition that the coefficients and must meet so that Eq. (2.31) is real.
4. A particle executing simple harmonic motion of amplitude passes through the equilibrium position
with speed What is the period of the oscillation?
[ ]
5. Two particles of mass and respectively, undergo simple harmonic oscillation of amplitude
and If the total energy of particle is twice the total energy of particle what is the ratio of their
periods:
[ ]
6. A particle undergoing simple harmonic motion has a speed when the displacement is and a speed
when the displacement is Find the period and the amplitude of the motion in terms of the
quantities given.
[ ]
7. Two springs of negligible mass that obey Hooke’s law have stiffness and respectively. The
springs are used in a vertical arrangement to support a single object of mass Write down the
equation of motion of the mass and show that the angular frequency of oscillation for the mass is
if the springs are tied in parallel and if the springs are tied in series.
8. A box of mass in which is placed a block of mass is suspended by a spring of negligible mass
that obeys Hooke’s law with stiffness The equilibrium length of the spring is The system is
constrained to move vertically only. If the spring is pulled down a distance from the equilibrium
position and then released, find the force of reaction between the block and the box as a function of
time. For what value of will the block just lift off the bottom of the box at the top of the vertical
oscillations. Neglect air resistance. Assume constant gravitational acceleration with magnitude
[ ]
9. Show that the ratio of two successive maxima in the displacement of a (under) damped harmonic
oscillator is constant. [Note: The maxima do not occur at the points of contact between the
displacement curve and the curve ]
10. Given that the amplitude of an under damped oscillator drops to of its original value after
complete cycles, show that the ratio of the period of oscillation to the period of the same oscillator with
no damping is given by

11. The terminal speed of a freely falling ball is When the ball is supported by a light elastic spring the
spring stretches by an amount Assume that air resistance is proportional to the velocity and show
that the natural frequency of oscillation of the system is given by

12. Find the driving frequency for which the maximum speed of a forced linearly damped harmonic
oscillator is greatest (only consider steady state solution).
[ ]
13. Discuss the claim that the energy of a damped harmonic oscillator at time is

14. Show that the expression for the quality factor given for an under­damped harmonic oscillator near
62 Chapter 2 Oscillations

resonance, is consistent with the definition of the quality factor given in section 5: It is defined
to be times the energy stored in the oscillator divided by the energy lost in a single period of
oscillation.
15. A linearly damped harmonic oscillator is subjected to a periodic force (only consider
steady­state solution).
(a) Show that the average rate at which the force does work is (
(b) Show that this is equivalent to the average rate at which energy is dissipated against the resistive
force.
(c) Show that is a maximum for the frequency of the free running oscillator.
(d) Find the values of for which is half its maximum value. [
]
16. Find the average value (steady­state solution) of the total energy of a linearly damped harmonic
oscillator subjected to a periodic force If is the work done against friction in one
period, show that when the ratio .( )
17. Prove equation (2.85).
18. Show that (2.88) is the solution of (2.87).
19. Explain carefully why Eq. (2.94) is a solution for a damped oscillator subjected to a series of blows of
impulse at time
20. Solve the problem of an oscillator under a simple periodic force
(turned on at by the Green’s function method, and verify that this reproduces the solution 2.62.
Assume that the damping is less than critical. Write as to perform the
integral.
21. Give an explanation why you think that the phase­space path of motion usually approaches zero along
the asymptote with slope in section (2.3.4).
22. Prove the last line in Equation (2.135). Hint: Write
23. The equation of motion for a simple pendulum has as tangential term

For small oscillation the period is well approximated by

Use the method of successive approximations to show that the period of a simple pendulum is better
approximated by the expression

where is the amplitude of oscillation if the amplitude is not very small.


Hint. . Solve for approximation then consider next non­zero term
in expansion.
63

Chapter 3
Central Forces and Celestial Mechanics
Summary
In this chapter we examine the motion of a particle subjected to a central force. First we shall give a
plausibility argument that the gravitational force of an extended body is central at large distances from
the gravitational system of the body. Then we shall look at motion in a general central field and later we
shall specialize to inverse central fields that include gravity. This will allow us to find descriptions of the
motion of planets around the sun. A last example that we shall look at the scattering of atomic particles as
an example of a repulsive inverse square force, the force between two like charges.

3.1 Introduction
A force whose line of action (that is the line that is parallel to the force and passes through its point of
application) passes through a single point or centre and whose magnitude depends on the distance from
that centre is called a central force. Central forces include gravity and electrostatic forces.

3.2 Potential Energy in a General Central Field


A general isotropic central field can be expressed in the following way:

(3.1)

in which is a unit radial vector. To test whether the field is conservative, we calculate the curl of It is
convenient to use spherical coordinates for which the curl is given by

For our central force The curl reduces to

(3.2)

Thus the curl vanishes for any central field and hence all central fields are conservative. The potential
function exists and if we choose a point infinitely far from the centre of the force as reference point, it is
given by

(3.3)

This allows us to calculate the potential energy function given the force function. Conversely, if we know
the potential energy function, we have

giving the force function for a central field.


64 Chapter 3 Central Forces and Celestial Mechanics

3.3 Law of gravity


Newton stated his law of universal gravitation in 1666:

Every particle in the universe attracts every other particle in the universe by a force that varies directly
as the product of the masses of the particles and inversely with the square of their distance apart. The
direction of the force is along the straight line that joins the two particles.

(3.4)

where is the force on particle of mass exerted by particle of mass The vector
is directed along the line segment that runs from particle to particle

Figure 3.1. Action and reaction in Newton’s law of gravity.

The law of action and reaction requires that The constant of proportionality is known
as the universal constant of gravitation. It can be determined in the laboratory by measuring the forces
between two spheres of known mass. The currently accepted values is

Nm kg

3.3.1 Potential Energy in a Gravitational Field.


Let be the work required to move a test particle of mass at constant speed, by an external force
along some prescribed path in the gravitational field of another particle of mass
at the origin of our coordinate system in Figure 2(a).
65

Figure 3.2. The work required to move a test particle from one point to another in gravitational field.

Since the force on the test practice is an external force must be applied to
move the particle (we think of moving the particle at constant speed). The work done in moving the
particle a distance is

(3.5)

We resolve into two components: the radial component parallel to and a component
perpendicular to as shown in Figure 2(b). Clearly

so is given by

(3.6)

where and are the radial distances of the particle at the beginning and end, respectively, of the path.
This verifies the fact that an inverse­square law force is conservative since the work is independent of the
path taken.

Note: In (3.5) and (3.6) we determined the work done against the gravitational field. Recalling that
for a conservative force we can write (3.6) is equivalent to the change in potential energy:
final initial .

If the resultant force on the test particle is zero, the velocity will be constant. Discuss whether it is
possible to move the test particle along an arbitrary path without changing the conclusions reached
above (Eq. (3.6). Keep in mind that in the discussion above the particle was assumed to move at
constant speed.

We can define the potential energy of a particle of mass at a given point in the gravitational field of
another particle as the work done in moving the particle from some (arbitrary) reference point to the point
66 Chapter 3 Central Forces and Celestial Mechanics

in question. It is convenient to take this reference point at infinity .

(3.7)

The gravitational potential is defined as the gravitational potential energy per unit mass:

It follows that if we have a number of particles, located at points then


the gravitational potential at point is given by

(3.8)

Prove Eq. (3.8) by determining the work done to move a test particle in the corresponding gravitational
force.
The gravitational field intensity is the ratio of the force field on a given particle and the mass of the
particle

(3.9)

It follows from definition that

(3.10)
(3.11)

Prove that for Eq.. (3.8).

Note that and give the gravitational potential and gravitational field intensity for a distribution
of point particles. It does not, for example, give the interaction potential between two extended objects.

3.3.2 Gravitational Force between a Uniform Spherical Shell and a


Particle
Consider a volume element centred on a distance from the centre of the spherical shell,

where is the angle makes with the axis and is the angle that the projection of on the ­
plane makes with the ­axis and is the density. Let be the distance from the origin to a test particle
of mass located on the ­axis. We assume that Then the gravitational force exerted on by a
uniform spherical shell of inner radius and outer radius is

gravity (3.12)

where is the distance from the test particle to the volume element and is a unit
vector from the test particle to the volume element .

2 It is important to note that it is not legitimate to define potential energy as the integral of unless we know that is
conservative, in other words, unless we know that the potential function exists.
67

Figure 3.3. Coordinates for calculating the gravitational force between a uniform sphere and a particle.

We can simplify the calculation by using the symmetry of the sphere. The force exerted on the
particle by can be decomposed into two components, one along the ­axis, and the
other in the plane perpendicular to the ­axis, .

For constant and and are constant. From symmetry it follows that the contribution of the
perpendicular components, for constant and vanishes when integrated over .

The force exerted by a ring of thickness and radius about the ­axis will therefore be
parallel to the ­axis, directed towards the origin. The magnitude is

The total force exerted by a shell of radius and thickness is obtained by integrating over

From the triangle we have from the cosine rule that

Differentiating on both sides with respect to yields


68 Chapter 3 Central Forces and Celestial Mechanics

We also have, using the cosine rule, that

Using these relations we get

(3.13)

is the volume of a spherical shell of radius and thickness . For a spherical shell of inner
radius i and outer radius o we have
o

shell
i

o i (3.14)

The gravitational force, shell on a particle outside a uniform spherical shell is therefore equal to

shell
shell (3.15)

where shell is the mass of the shell, is the distance from the centre of the shell and is the unit radial
vector from the origin. This means that a uniform spherical shell attracts an external particle as if the
whole mass of the shell were concentrated at the centre of the shell. This will be true for each concentric
spherical portion of a sphere. A uniform spherical body therefore attracts an external particle as if the
entire mass were located at its centre. This is also true for a nonuniform shell as long as the distribution of
mass is radially symmetric.
If the gravitational force between a test particle and a uniform spherical shell can be described as if the
shell were a point particle located at the centre of the shell, it follows that the gravitational field between
two uniform spherical shells can be described as if there were two point particles with masses equal to the
masses of the spherical shell located at the centre of the spherical shells.
Note that the gravitational force on a particle inside a uniform spherical shell is zero (See exercise at the
end of this chapter).

3.3.3 Potential of a Uniform Spherical Shell


Using the same notation as in Figure 3 the potential for a spherical shell of thickness for a point outside
the shell, can be expressed as

o i

(3.16)
69

where is the mass of the uniform shell. This is the same as the potential of a particle of mass located
at the centre of the shell.
The potential inside the shell is constant, and hence the potential force field is zero (exercise).

3.3.4 Potential and Field of a Thin Ring


The potential in the plane of a thin circular ring of radius and mass can be calculated as follows:

in which is the linear density of the ring ( For a point outside the circumference in the plane
of the ring it is convenient to express the integral in terms of the angle as defined in Figure 4

Figure 3.4. Coordinates for calculating the gravitational field of a ring.

We have, from the sine rule,

Differentiating

since Now so

hence the integral becomes

(3.17)

where

(3.18)
70 Chapter 3 Central Forces and Celestial Mechanics

is the complete elliptical integral of the first kind. There is no analytic solution. Approximations can be
found by expanding the integrand and integrating term by term:

(3.19)

The field intensity at a distance from the centre of the ring is then in the radial direction, since is not a
function of

(3.20)

The field is not given by an inverse­square law. If the first term dominates and the field is
approximately of the inverse­square type. This is true for a body of any shape: for distances large
compared to the linear dimensions of the body, the field tends to become predominantly inverse­square.
See http://en.wikipedia.org/wiki/Gravitational_potential for a general discussion.

3.3.5 Angular Momentum


The angular momentum theorem states that

(3.21)

For a particle in a central field the force acts in the direction of the radius vector . This means that
the cross product vanishes, and therefore there is zero moment. Consequently for a central field

and therefore

constant (3.22)

The angular momentum of a particle moving in a central field remains constant.

As a corollary, it follows that the path of a particle in a central field remains in a single plane because
the constant angular momentum is perpendicular to both and and is therefore normal to the plane in
which the particle moves. Thus it is possible, without loss of generality, to use plane polar coordinates
when investigating central motion.
3.3.5.1 Magnitude of Angular Momentum

Resolve the velocity in its radial and transverse components (plane polar coordinates):
71

the magnitude of the angular momentum is

constant (3.23)

for a particle moving in a central field.


We usually simply write

where we use the convention that This is always possible. In a central field the angular momentum
is conserved and therefore the angular direction of motion does not change. We simply choose the
coordinate system with which implies that points in the direction.

3.3.6 Law of Areas


The angular momentum of a particle is related to the rate at which the position vector sweeps out area.
The area defined by position vectors and representing the motion of a particle in a time interval
can be expressed as

Figure 3.5. Area swept out by the radius vector.

Upon division by and taking the limit when we have

(3.24)

From the definition of we can write

(3.25)

for the rate at which the radius sweeps out area. Since the angular momentum is constant in any central
field, it follows that the areal velocity is also constant in a central field.

3.3.7 Orbit of a Particle in a Central-Force Field


In a central field the equation of motion reads

In plane polar coordinates the radial component of is and the transverse component is
72 Chapter 3 Central Forces and Celestial Mechanics

and we can write the equation of motion as

The radial and transverse components of the equation of motion are

(3.26)

(3.27)

Explain what implies.

Equations (3.26) and (3.27) forms a coupled set of differential equations that requires the solution of
and There is no simple solution for arbitrary , but we can learn something about the general
behaviour of a particle in a central field be examining the equations a little more carefully.
Let be the angular momentum per unit mass,

(3.28)

Recall that the angular momentum is given by

constant. (3.29)
from which it follows that
constant. (3.30)
Recall that we choose the orientation of coordinate system to ensure that and
We can use equation (3.30) to eliminate from equation (3.26):

or

(3.31)

This looks like the equation of motion of a particle moving in one dimension with a force acting
on it. It is not a simple differential equation to solve, but we can learn something from it. The effective
force comes from the angular momentum and is non­zero unless the particle is moving directly towards
or directly away from the centre of the field, when the angular momentum is zero.
Explain why the particle moves in a straight line that passes through the centre of the field if the
angular momentum is zero.

For non­zero angular momentum, and will become infinitely large as tends to zero. Since
, behaves as a repulsive force that keeps the particle away from the centre of the field. So unless
faster than In other words, unless there is a contribution to
of the form , with the particle will experience an infinite repulsive force near the origin
if the angular momentum is non­zero and hence will never be found at the origin.
In your own words, explain the statement above about the form of as
73

To find the equation of the orbit (or path that the particle follows in space), we have to find the
relationship between and or To do this, we eliminate the time from equation (3.26). Let
then

(3.32)

Differentiating a second time, we have

(3.33)

Equation (3.26) can now be written as

(3.34)

This equation is the differential equation for the radial coordinate for a particle moving in a central
field since the solution will give us .

Example
A particle in a central field moves in a spiral orbit

Determine the form of the force function, the potential function and the how varies as a function of time.

We have

and

From (3.34)

Hence

and

The potential function is

For central fields is constant. Thus


74 Chapter 3 Central Forces and Celestial Mechanics

or

and so

where we take the constant of integration to be zero. So

3.3.8 Energy Equation of the Orbit


A central force is conservative, hence the total energy is constant. We can use this to give us
another equation that describes the motion of a particle in a central field.
The square of the speed in polar coordinates is given by

which leads to

constant. (3.35)

Eliminate to obtain the energy equation of the orbit :

(3.36)

is the effective potential of the orbit. The contribution is called the


centrifugal potential. It is not a real potential, but is a consequence of conservation of angular momentum.
Conservation of angular momentum thus results in a repulsive contribution to the potential energy.
From Eq. (3.36) we have

(3.37)
75

E
U(r)
2 2
L /(2mr )
V(r)

r0 r1 r

E<0

E1

Figure 3.6. Turning points for an attractive inverse square central force.

Let (an inverse square central force). Note that for an inverse square central force there is a
minimum energy allowed for the system,

How does the particle move if ?

It follows from equation (3.37) that vanishes when For there are two points
where this happens, and on Figure 6. The radial velocity will change sign at these radial distances
and we call these points the turning points (or apses) of the motion. For there is only one turning
point. This tells us that the motion will be confined to a region around the centre for energies less than
zero, and for energies greater than or equal to zero the object will approach up to a minimum distance and
then move away to infinity.
We can eliminate the time from the energy equation and express it in terms of Using (3.32),
(3.35) becomes

(3.38)

In this equation the only variables are and

Equations (3.34) and (3.38) give different equations If you solve these equations, will you get
the same result?

3.3.9 Kepler’s Laws of Planetary Motion


The fact that the planets move about the sun in such a way that the areal velocities are constant was
discovered by Johannes Kepler in 1609. Kepler deduced this rule and two others from the painstaking
study of planetary motion recorded by Tycho Brahe. Kepler’s three laws are:

(1) Each planet moves in an ellipse with the sun as focus.


(2) The radius vector sweeps out equal areas in equal times.
76 Chapter 3 Central Forces and Celestial Mechanics

(3) The square of the period of revolution about the sun is proportional to the cube of the major axis
of the orbit.

Newton showed that these laws are a consequences of the law of gravity. The second law, as we have
already shown, is a consequence of the fact the gravitational field of the sun is central.

3.3.10 Orbits in an inverse-Square Field


Let With the inclusion of the minus sign, the constant of proportionality is positive for an
attractive force and negative for a repulsive force. ( for a gravitational field.)
The orbit equation (3.34) becomes

(3.39)

Let so that

with solution

(3.40)
This gives the general solution

(3.41)

or

(3.42)

and finally

(3.43)

where the constants of integration and are determined from the initial conditions. This is the equation
of a conic section (ellipse, parabola or hyperbola) with the origin at a focus (see appendix B). The value of
determines the orientation of the orbit, so we can, without loss of generality in discussing the form of
the orbit, choose With

(3.44)

the equation can be written in standard form

(3.45)

where

(3.46)

We can also express the obit as

(3.47)

The constant is called the eccentricity and determines the shape of the orbit, while the semilatus
rectum, determines the scale. The different cases for an attractive potential are
77

ellipse
circle (special case of an ellipse)
parabola
hyperbola

Figure 3.7. The family of central conics.

From (3.45), is the value of for . The value for for is

(3.48)

In reference to the elliptic orbits of the planets around the sun, the distance is called the perihelion
and the distance the The corresponding distances for the moon around the earth, and orbits
of the earth’s artificial satellites, are the and distances, respectively.
The orbital eccentricities of the planets are quite small. ( See Table 3.1 below.) For example. in the case
of the earth km km The comets, on the other hand, tend to
have large orbital eccentricities (highly elongated orbits). Haley’s comet has with perihelion
distance km, while at aphelion it is beyond the orbit of Neptune. Many comets (nonrecurring)
have parabolic or hyperbolic orbits.

For a repulsive potential, the parameter is negative. That means that in Eq. (3.47) the eccentricity and
the right hand side of the equation is negative. The conic section that is then described by Eq. (3.47) for
the repulsive case can be written as
(3.49)
This corresponds to the branch of the hyperbola to the right in graph 7.
3.3.10.1 Energy Equation

The orbits can also be derived directly from the energy equation (3.38)

(3.50)

We consider an inverse potential of the form . With multiply the energy


78 Chapter 3 Central Forces and Celestial Mechanics

equation by

(3.51)

Multiply by and add to each side:

(3.52)

Define the variable

(3.53)

so

(3.54)

The left side is a sum of squares, so we find solutions only when the right side is also positive. This
means that the minimum allowed energy is

The general solution is

where is an arbitrary constant of integration.

(3.55)

These are the polar equations of conic sections with the origin at a focus . From comparison of (3.55) and
(3.42) it follows that
An additional bit of information we get from this derivation is that from (3.54) and the definition of it
follows that

(3.56)

For a repulsive potential, the energy is always positive and it follows that is greater than one for a
repulsive potential.

3.3.10.2 Elliptic Orbits ( )

Figure 3.8. Ellipse.


79

If we choose the axis so that then we can write the equation of the ellipse, (3.47), as

(3.57)

in Cartesian coordinates. Here

(3.58)

and

This is the equation of the ellipse with centre at and semimajor axis and semiminor axis
The semimajor axis is determined by the energy, the simeminor axis is determined by the angular
momentum and energy and the semilatus rectum by the angular momentum.

3.3.10.3 Hyperbolic Orbits ( )

For both attractive and repulsive cases, the Cartesian equation of the orbit is

(3.59)

where

(3.60)

This is the equation of a hyperbola with centre at and semi­axis and One orbit (the one on
the left in Fig. (9) corresponds to the attractive case and the other to the repulsive case. As before is
determined by the energy and by the angular momentum and energy.

Figure 3.9. Hyperbola.


80 Chapter 3 Central Forces and Celestial Mechanics

The directions in which becomes infinite are, in the repulsive case, and, in the
attractive case In both cases the angle through which the particle is deected from its
original asymptotic line of motion is

is the scattering angle . The semiminor axis is known as the impact parameter. In terms of the
limiting velocity (particle infinitely far from the centre) the energy so that from (3.58)
Also

(3.61)

Hence

(3.62)

3.3.11 Orbital Parameters from Initial Conditions


From (3.46) we find that the eccentricity can be expressed as

(3.63)

Let be the speed of the particle when Then, from the definition of the constant we have

the eccentricity is then given by

(3.64)

Is this equation for valid for all and


For a circular orbit, , we have

(3.65)

Now let us denote the quantity by

so that if the orbit is a circle. The expression for the eccentricity can then be written as

(3.66)

and the equation of the orbit can be written as

(3.67)
81

With

we can also write

(3.68)

(3.69)

We get the value of by setting thus

(3.70)

Example
A space shuttle is going around the earth in a circular orbit of radius A sudden blast of the rocket
motor increases the speed by percent. Find the furthest distance from the earth in the new orbit while
the perigee remains .

Let be the speed in the circular orbit, and let be the new initial speed, then

For to remain the perigee, must remain a turning point. At a turning point the velocity and position
vectors are perpendicular to each other So for to remain the perigee, the thrust must increase the
velocity in the direction of motion.
From (3.70) the expression of the new orbit is

Note: The new perigee distance is only if the change in velocity is in the direction of motion. If the
change in velocity is not in the tangential direction, the new perigee distance will be different from . If
the speed is decreased, from (3.64), the perigee distance has to change.

3.3.12 Orbital Energies in the Inverse-Square Field


Since the potential­energy function for an attractive inverse­square force field is given by

the energy equation of the orbit (3.38), reads


82 Chapter 3 Central Forces and Celestial Mechanics

or

Upon integration

where is the constant of integration. If we let and solve for we obtain

(3.71)

or

(3.72)

This is the polar equation of the orbit. If we compare it with (3.45), we find that

(3.73)

which is equivalent to equation (3.56) as it should be. This allows us to classify the orbits according to the
total energy as follows
closed orbits (ellipse or circle)
parabolic path
hyperbolic path

Since and is constant, the closed orbits are those for which and the open orbits are
those for which

Example
A comet is observed to have a speed when it is a distance from the sun, and its direction of motion
makes an angle with the radius vector from the sun. Find the eccentricity of the comet’s orbit.

In the sun’s gravitational field where is the mass of the sun, and is the mass of the
comet. The total energy is

constant (3.74)

and the orbit will be elliptic, parabolic or hyperbolic, according to whether is negative, zero or positive.
Accordingly, if is less than, equal to or greater than the orbit will be an ellipse, a parabola
or hyperbola, respectively. Now

The eccentricity from (3.73), therefore has the value

(3.75)

The product may be expressed in terms of the earth’s speed and orbital radius (assuming a
circular orbit), namely,
83

The equation giving the eccentricity can be written as

(3.76)

3.3.13 Limits of the Radial Motion

ea

Figure 3.10. Limits of radial motion: Ellipse, .

From the radial equation of the orbit, (3.72), we see that

(3.77)

(3.78)

In the case of an elliptical orbit, the major axis of the ellipse is given by

We then find

(3.79)

This means that the values of is determined entirely by the total energy as we saw in equation
(3.58)

3.3.14 Periodic Time of Orbital Motion


In Section 3.3.9 we showed that the areal velocity of a particle in a central field is constant. From (3.25)
and (3.28) the time, it takes a particle to move from point to point during which it sweeps out an
area is given by

(3.80)

The area of an ellipse is where and are the semimajor and semiminor axes, respectively. Thus
for a particle in an inverse­square field in an elliptic orbit the time required for the particle to complete
one orbital path is

(3.81)

Now for an ellipse

(3.82)
84 Chapter 3 Central Forces and Celestial Mechanics

1. Periods, semimajor axes and orbital eccentricities for planets in the solar system.

So we have

(3.83)

Further, from (3.44) and (3.48), the major axis is given by

(3.84)

We can therefore express the period as

(3.85)

Thus for an inverse­square force field the period depends only on the size of the major axis of an elliptic
orbit. Since for a planet of mass moving in the sun’s gravitational field, we can write the
period of the planet as

where Clearly, is the same for all planets. (Note: We have ignored the inuence of
the other solar and stellar objects on the period. Interactions with other ’large’ objects will perturb the
period somewhat.) Equation (3.85) is a mathematical statement of Kepler’s third law. If is expressed in
astronomical units ( astronomical unit earth km and is in years, then is unity.
In Table 1 data for the planets in the solar system are listed.

3.4 Motion in an Inverse-Square Repulsive Field

3.4.1 Scattering of Atomic Particles


Consider an incident where a high speed particle of charge and mass passes near a heavy particle of
charge of the same sign. The incident particle is repelled by a force given by Coulomb’s law :

where the position of is taken to be the origin. The differential equation of the orbit takes the form
Section 3.4 Motion in an Inverse­Square Repulsive Field 85

and so the equation for the orbit is

or from (3.72)

(3.86)

since The orbit is a hyperbola since for a repulsive potential the total energy
is always greater than zero. The eccentricity must therefore by greater than one
and the orbit is hyperbolic.
The incident particle approaches along one asymptote and recedes along another.

Figure 3.11. Hyperbolic path of a charged particle moving in the inverse­square repulsive field of another
charged particle.

The direction of the polar axis is chosen such that the initial position of the particle is at
The minimum distance of approach is when that is when Since when
then is also infinite when and the angle through which the particle is deected is

Furthermore, in (3.86) the denominator on the right vanishes at and Thus

(3.87)

from which it follows that

(3.88)

In scattering problems it is convenient to express the constant in terms of another quantity the impact
parameter. The impact parameter is the perpendicular distance from the origin (scattering centre) to the
initial line of motion of the particle. We have

(3.89)

where is the initial speed of the particle. Note that corresponds to the semiminor axis of the hyperbolic
86 Chapter 3 Central Forces and Celestial Mechanics

orbit in section 3. We know that the energy is constant and is equal to the initial kinetic energy
because the initial potential energy is zero ( Accordingly

(3.90)

This is, of course, a special case of (3.62).

Example
o
1 An alpha particle is emitted by radium with energy eV. It undergoes a deection of
upon passing near a gold nucleus. What is the value of the impact parameter?
For alpha particles and for gold where ­ is the charge of an electron. From

From equation (3.90)

m.
2 Calculate the distance of closest approach in the above example.
The distance of closest approach is given by the equation of the orbit, Eq. (3.86) for , so

(3.91)

Using (3.90) leads to

(3.92)

Note: The equation for becomes indeterminate when This is the case when
in other words when the scattering source is approached in a straight line. The angle
o
of deection is then and the value of can be found from the fact that is constant. At the
turning point the kinetic energy is zero and so

(3.93)
Exercises 87

Exercises
1. Show that the gravitational force on a particle inside a uniform spherical shell is zero.
2. Show that the gravitational potential inside a uniform spherical shell is constant, and hence the force
field is zero.
3. Assume that the earth is spherical and that the density is uniform. Show that a particle dropped into a
straight hole drilled through the earth, passing through the centre of the earth, will execute an harmonic
motion. Find the period of oscillation and show that it depends on the density, not the size of the earth.
[Hint: Use the results of the previous two exercises.]
4. A particle moving in a central field describes a spiral orbit of the form Show that the force
is inverse­cube and that varies logarithmically with time
5. The orbit of a particle moving in a central field is a circle of radius passing through the origin. Show
that and that the force law is inverse fifth power.
6. A satellite is going around the earth in a circular orbit of radius moon
It is desired to place the
satellite in a new orbit with its apogee distance equal to the radius of the moon’s orbit around the earth.
(a) In which direction should the thrusters be fired to get the maximum change in energy for the
minimum expenditure of fuel?
(b) Show that the perigee distance of the new orbit is equal to the radius of the circular orbit if the
velocity is changed tangent to the path.
(c) Determine the ratio of the final and initial speeds.
(d) Calculate the apogee distance if the speed ratio is percent too large and comment on the accuracy
required to achieve the new orbit.
7. Haley’s comet has an orbit with eccentricity and a perihelion distance of km.
Calculate the period and the speeds at perihelion and aphelion.
8. Assuming that the orbit of the earth around the sun is circular, let be the radius of the orbit and
the speed of the earth relative to the sun. If a comet is seen at a distance of astronomical units
(1 astronomical unit ) moving at a speed of times the earth’s speed, show that the orbit of the
comet about the sun will be hyperbolic, parabolic or elliptic depending on whether is greater than,
equal to or less than (Only consider the gravitational field of the sun).
9. A particle moves in an inverse square force field. Show that the product of the maximum and minimum
speeds is equal to where is the semimajor axis and the period.
10. Show that the radial part of the equation of motion for a particle in a central field is the
same as that of a particle undergoing rectilinear motion in an effective potential given by

where
11. Consider an attractive inverse square law force.
(a) Show that the orbit is a circle when

(b) Show that for a circular orbit the velocity is

(c) Show that for a circular orbit the total energy


88 Chapter 3 Central Forces and Celestial Mechanics

and the kinetic energy

12. The orbits of synchronous communications satellites are chosen so that viewed from the earth they
appear to be stationary. Find the radius of the orbits. (These satellites are also known as geostationary
satellites.)
13. Consider a projectile launched from the earth, (assumed to be a sphere of radius ) with speed and
angle with the vertical.
(a) Show that the escape velocity from the earth is

km.s
(b) If show that the projectile will reach a maximum height at a distance from the centre of
the earth given by the roots of the equation

(c) If is equal to the circular orbit velocity in an orbit just above the surface of the earth, show that

14. Show that the expressions for the ellipse and hyperbola as given in Equations (3.57) and (3.59) can be
derived from the orbit equation in polar coordinates, equation (3.47).
15. A body moves in a circular path in an attractive inverse square force field.
(a) Show that for a circular path the kinetic energy increases while the total enery decreases with
decrease in radius.
(b) To go from one circular path to another, how must the energy and momentum change?
16. (a) Show that the angular momentum for a particle moving in a central potential is conserved and
hence that it is justified, without loss of generality, to use plane polar coordinates to describe the
motion of a particle in a central field.
(b) Two objects of mass, and respectively, are fired vertically and horizontally from the same
point on the surface of the earth of radius and mass with the same speed
(i) Ignore friction in the atmosphere and show that if

then can escape from the gravitational field of the earth.


(ii) If is projected with the same speed, will it escape or not?
(iii) The general orbit equation for a particle of mass in a central force field, is

where is the angular momentum of the particle. The eccentricity, can be expressed as

Find the orbit equation, and sketch the orbit, that describes the orbit of particle if launched
horizontally from the earth with speed

[Examination November 1998]


17. A particle moves in a central potential.
(a) Show that the particle moves in a plane.
Exercises 89

(b) Given that in polar coordinates show that the radial


coordinate satisfies the differential equation

where is the angular momentum, the potential energy and is the total mechanical energy.
(c) Show that for a particle in an attractive inverse square central force there is a minimum allowed
energy for a given angular momentum and that the particle will move in a closed orbit if the energy
and will move in an open orbit if the energy
[Examination November 1999]
18. Derive Eq. (3.88).
90 Chapter 4 Noninertial Reference Frames

Chapter 4
Noninertial Reference Frames
Summary
Reference frames in general are non­inertial. In this chapter we look at how we have to formulated
Newton’s laws of motion in non­inertial reference frames. As examples we shall look at motion near the
surface of the Earth when the centre of the Earth, not the surface, is taken as a reference point in an inertial
reference frame. We shall look at a plumb line, projectile motion and the Faucault pendulum.

4.1 Introduction
Up to now we have used reference frames in which the laws of motion take on the simple form expressed
in Newton’s laws, in other words, we used inertial reference frames. A coordinate system fixed to the
surface of the earth, for example, is the most convenient one to use in expressing the motion of a projectile,
although the earth is rotating and accelerating and therefore the coordinate system is not an inertial
reference frame. In this chapter we shall look at how to describe motion in a moving reference frame.

4.2 Translation of the Coordinate System


The simplest motion of the coordinate system is that of pure translation. In Fig. (1) the primary coordinate
system is assumed to be an inertial reference frame while the secondary coordinate system is
moving. First we consider the case where the orientation of the axes of the moving coordinate system does
not change.

Figure 4.1. Relationship between the position vectors for two coordinate systems.

The position vector of particle is given by in the fixed system and by in the moving system. The
displacement of the origin, with respect to is given by thus
Section 4.3 Inertial Forces 91

Taking first and second derivatives with respect to we find the velocities and accelerations:

(4.1)
(4.2)

where and are the velocity and acceleration of in the fixed system and and the velocity and
acceleration in the moving system, while and are the velocity and acceleration of the moving origin.
In particular, if the moving system is not accelerating, then

that is the acceleration measured by observers in the two systems is identical.

4.3 Inertial Forces


If the primary system is inertial so that Newton’s second law

is valid in this reference frame, then from (4.2) the equation of motion in the moving system is

(4.3)

Thus an acceleration of the secondary reference system can be accounted for by an addition of to
the force We call this an inertial term. Inertial terms have no physical origin, but stem from the choice
of reference system. By definition, an inertial reference system is one in which there are no inertial terms.
If the primary system is inertial, then the secondary system is also inertial if it moves at constant velocity
relative to the primary system. If the observer were not aware that the system is accelerating, he or she will
suspect that there is an additional force acting on the system.
The inertial terms are sometimes called ’fictitious forces’ or ’inertial forces’. Whether one calls this term
a ’force’ or not is simply a matter of terminology. These terms have no physical origin.

Example
A block of wood rests on a rough table. If the table is accelerated in a horizontal direction, under what
conditions will the block slip?

Let be the coefficient of friction between the block and table. Then the force of friction has a
maximum value where is the mass of the block. In the moving reference frame, the total force is
measured as

where is the acceleration of the table. Hence in the moving reference frame the condition that the block
slips is

or
92 Chapter 4 Noninertial Reference Frames

4.4 Angular Velocity; Rate of Change of a Vector


Consider a rigid body that is rotating with angular speed about a fixed axis. Let be a unit vector along
the axis of rotation, whose direction is determined by a right­hand rule: it is the direction in which the
thumb of the right hand points when the fingers are curled in the direction of rotation. The angular velocity
is defined as

If we take the origin of the coordinate system to lie on the axis of rotation, then the velocity of a point in
the body at a position is given by

(4.4)

Note that and and are measured in a fixed coordinate system.

Figure 4.2. Angular and linear velocities.

To prove (4.4), we note that the point moves about the axis with a radius thus its
instantaneous speed is

Here is the period of revolution for a constant angular velocity. The direction of the velocity is in the
direction of since is perpendicular to and
For the earth is a vector pointing along the polar axis towards the north pole. Its magnitude is
divided by the sidereal day (the rotation period relative to the fixed stars, which is less than that with
respect to the sun by one part in ). That is

s s

Any position vector that is fixed in the rotating body can be written as the difference of two position
vectors

hence

(4.5)
Section 4.5 General Motion of the Coordinate System 93

This means that the vector does not have to start at the origin of the reference system to yield (4.4). In
particular if rot rot rot are Cartesian unit vectors fixed to the rotating body, which can represent a
moving coordinate system, then

rot rot rot rot rot rot (4.6)

If, for example is in the rot direction, then

rot rot rot rot rot

In this case the body is rotating about an axis parallel to the rot axis, and rot and rot change direction, so
equation (4.6) is valid for the instantaneous directions of the unit vectors.
Now consider an arbitrary position vector expressed in terms of the rotating coordinate system

rot rot rot

Then

rot rot rot rot rot rot


f

rot rot rot rot rot rot

(4.7)
m

where f
and m
are used to indicate time derivatives of scalar quantities in the fixed and moving
reference systems respectively. We shall employ this convention in the remainder of this chapter .
m
indicates only the derivatives of the scalar components of the vector decomposed in the moving coordinate
system.

4.5 General Motion of the Coordinate System


If the moving coordinate system undergoes both translation and rotation relative to the inertial system, it
follows that

f f f

(4.8)
m

In summary, the velocity of a moving particle as measured in an inertial reference system can be expressed
as a sum of three terms:
the velocity of the particle measured in the moving coordinate system, m
the apparent rotational velocity that the particle has due to the rotation of the moving reference
system,
the linear velocity of the origin of the moving reference system relative to the inertial system.

Note that in the discussion is measured in the inertial reference system and describes the change
in orientation of the moving coordinate system as measured in the inertial system. The change of the
direction (orientation) of a unit vector fixed in the moving system does not depend on the origin of the
moving coordinate system, but only on the angular velocity which measures the change in orientation of
3 The time derivative of a scalar quantity does not depend on the reference frame, but the time derivative of a vector does.
94 Chapter 4 Noninertial Reference Frames

the moving coordinate system Naturally the angular velocity vector can be expressed in terms of the unit
vectors of the moving coordinate system.

Equation (4.7) was derived for position vectors, but it can be shown (see exercise 1) that the time
derivative of any vector in the fixed (inertial) reference system can be written as

f m

In particular

f m

(4.9)
m

The acceleration relative to an inertial observer is then given by

f m f f f f

(4.10)
m m m

The first term on the right­hand side is the acceleration of the particle measured in the moving system and
is the acceleration that will be measured by an observer in the moving system. The next three terms are
rotational terms for the particle as seen in the fixed system. The last term is the acceleration of the moving
origin measured in the fixed system.

m
Coriolis acceleration
m
transverse acceleration
centripetal acceleration

The centripetal acceleration is perpendicular to the axis of rotation as may be seen by writing it as (see
exercise (1.4))

or deduced directly from the definition of the vector cross product.

 x  x r)
v =  xr

Figure 4.3. Centripital acceleration


Section 4.5 General Motion of the Coordinate System 95

Examples
1. A wheel of radius rolls along the ground without slipping with constant forward speed Find the
acceleration, relative to the ground, of any point on the rim.

Choose a coordinate system fixed to the rotating wheel, with the origin at the centre and the ­axis
passing through the point on the rim, Fig. (4).

Figure 4.4. Rotating coordinate system fixed to a rolling wheel.

We have

m m

The angular velocity of the rotating coordinate system is given by

The only term that survives in (4.10) is the centripetal acceleration, hence

Thus has magnitude and is always directed towards the centre of the wheel.

2. A bicycle travels with constant speed around a track of radius What is the acceleration relative to
the ground of the highest point on one of the wheels, each of radius if the wheels do not slip?

Choose a coordinate system with origin at the centre of the wheel with the ­axis horizontal and
pointing towards the centre of the curvature of the track. We do not let the moving coordinate system
rotate with the wheel, Fig. (5). Let the ­axis point vertically upwards. The moving coordinate system
now has an angular velocity
96 Chapter 4 Noninertial Reference Frames

and acceleration

relative to

Figure 4.5. Wheel rolling along curved track.

Since each point on the wheel is moving in a circle of radius with respect to the moving origin, the
acceleration in the moving reference frame has magnitude and is directed towards In the moving
system we therefore have for point at the top of the wheel

The velocity of this point in the moving system is

so the Coriolis acceleration is

Since the angular velocity is constant, the transverse acceleration is zero. The centripetal
acceleration is also zero since

The total acceleration of the highest point of the wheel relative to the ground at a given time is

Note that the direction of the unit vector changes with respect to the ground.

4.6 Dynamics of a Particle in a Rotating Coordinate


System
If the primary coordinate system is considered to be an inertial system, the fundamental equation of motion
Section 4.7 Effects of the Earth’s Rotation 97

in that system is

The equation of motion in the moving system can be written as

(4.11)
m m m

The inertial terms are usually referred to as

Coriolis m
Coriolis force
transverse m
transverse force
centrifugal centrifugal force
translation translation force

The term is due to the translation of the origin of the moving coordinate system as discussed
before. The actual contributions of the four terms depend on the coordinate system in which the motion is
described. They arise from the inertial properties of matter rather from interaction with other bodies.
The Coriolis force is present only if the particle is moving in a rotating coordinate system. Its direction is
always perpendicular to the velocity vector of the particle in the moving coordinate system. The Coriolis
force thus seems to deect a moving particle at right angles to its direction of motion.
The transverse term is only present when there is an angular acceleration of the rotating coordinate
system. The force is always perpendicular to the position vector hence the name transverse.
The centrifugal force is the familiar term arising from the rotation about an axis. The force is directed
outward away from the axis of rotation and is perpendicular to that axis of rotation.

4.7 Effects of the Earth’s Rotation


Since the angular speed of the earth’s rotation about the polar axis is only rad s we might
expect the effects of the rotation to be small. Nevertheless, it is the spin of the earth that produced the
equatorial bulge; the equatorial radius as about km greater than the polar radius.

4.7.1 Static Effects. The Plumb Line


We assume for the moment that the centre of the earth is the origin of an inertial reference frame. Consider
the bob at the end of a plumb line, at rest near the surface of the earth. We fix the orientation of our
coordinate system to the surface of the earth with the origin at the position of the bob, so that . The
angular velocity is in the direction of the earth’s axis and is very nearly constant so that . From
(4.11) the result is

The force is given as the sum of two forces: the true gravitational attraction of the earth, and the
tension in the string, which we measure as shown in Fig. 6.
98 Chapter 4 Noninertial Reference Frames

Figure 4.6. Gravitational and centrifugal forces on a particle on the surface of the earth.

We have

If we model the earth as a radially uniform sphere, the vector is directed towards the centre of the
earth. If we take the centre of the earth as the origin of the fixed reference frame, the acceleration is
just the centripetal acceleration of our moving origin:

where is the radius of the earth and is the geocentric latitude. Since and
points perpendicularly away from the earth’s axis, the plumb line does not point towards the centre of the
earth, but deviates by a small angle as shown in Fig. 7.

Figure 4.7. Vector diagram defining


Section 4.7 Effects of the Earth’s Rotation 99

Applying the sine rule, we have

or, since is small

The angle vanishes at the equator and at the poles The maximum deviation from
the true vertical (direction towards the centre of the earth) is for where

rad.

Approximating the earth as spherical, the horizontal and vertical components of are

The magnitude of the centrifugal force can be found from and

s km
mm s

This is about of At the pole we expect

and at the equator

so

pole equator
mm s

The actual measured value is mm s The discrepancy arises from the fact that the earth is not a
perfect sphere, but rather more spheroidal in shape, attened at the poles.

4.7.2 Dynamic Effects


4.7.2.1 Motion of a Projectile

Let us fix the origin of our moving reference system near the surface of the earth and that of the fixed
system at the centre of the earth. As before we assume that the centre of the earth can be taken as a point
in an inertial reference frame.
The equation of motion for a particle moving near the surface of the earth can then be written as

app
m m

where app represents the applied forces other than gravity. From the previous section, the term
can be called and the equation of motion becomes

app
m m
100 Chapter 4 Noninertial Reference Frames

Let us consider the motion of a projectile. We neglect air resistance, so app . As discussed in the
previous section, the term is small and as a first approximation can be neglected. The
equation of motion reduces to

(4.12)
m m

where the last term is the Coriolis force.


To solve (4.12) we choose the directions of the moving coordinate axes with the ­axis in the direction
of a plumb line, the ­axis pointing east and the ­axis pointing north, as in Fig. 8.

y z

O x

Figure 4.8. Coordinate axes for analyzing projectile motion.

With this choice of axes, we have to a good approximation that

and

(4.13)

Therefore

(4.14)
m m m m
Section 4.7 Effects of the Earth’s Rotation 101

The equation of motion can be written in component form as

(4.15)
m m m

(4.16)
m m

(4.17)
m m

Upon integration with respect to we get

(4.18)
m m

(4.19)
m m

(4.20)
m m

where , and m m
are the initial components of the position and velocity
m
vectors respectively. Ignoring terms involving we can substitute the expressions for the components of
the velocity in the equations of motion to get, for example,

m m m

Integrating again yields

m m m m

and therefore

m m m

The other components yield

(4.21)
m m

(4.22)
m m

Case1 : Assume that the particle is dropped from rest from the origin of the moving system. Then

The ­component tells us that the particle drifts to the east.

Case 2: Consider a particle fired with a high velocity in a horizontal direction to the east. Then

m m m
102 Chapter 4 Noninertial Reference Frames

and

(4.23)
(4.24)
(4.25)

which means that the particle drifts to the right of its direction of travel if launched in the northern
hemisphere and to the left when launched in the southern hemisphere.

4.7.2.2 Coriolis forces and the weather

The magnitude of the Coriolis force per unit mass is less than ( Coriolis m
which
for a velocity of km h is about . Although small it nevertheless has a significant impact
on meteorological phenomena. Wind is an air mass in motion, and in the absence of Coriolis forces the
motion would be along the pressure gradient from high to low pressure, and therefore perpendicular to the
isobars. However, the Coriolis forces deect the wind. At equilibrium the wind patterns are stationary
and the velocity remains constant. The net force on the air mass must be zero. In such cases the Coriolis
force must be equal and opposite to the pressure gradient force, which requires that the wind direction be
parallel to the isobars. A region of low pressure with roughly concentric isobars is known as a cyclone.
As a consequence of Coriolis forces a cyclone circulates in a counterclockwise direction in the northern
hemisphere and clockwise in the southern hemisphere.
Near the equator heating causes air to rise and colder air ows towards the equator to replace it. The
Coriolis effect causes the winds to deviate to the west and we get north­east trade winds in the northern
hemisphere and south­east trade winds in the southern hemisphere.

Deection of wind from the direction of the pressure gradient by Coriolis forces (shown for the northern
hemisphere.)

4.7.3 The Foucault Pendulum


A Foucault pendulum is a pendulum free to swing in any direction and arranged to be perfectly symmetric
so that its periods of oscillation in all directions are equal. (It should be long and sufficiently heavy so
that it will swing freely for several hours despite air resistance.) If the amplitude is small, the pendulum
equation can be approximated by the equation of two dimensional simple harmonic motion.
We use the coordinate system shown in Fig. 8. The vertical component of the Coriolis force is negligible
since it is merely a small correction to The important components are the horizontal ones. For small
amplitudes the bob moves in an almost horizontal plane, so that With these approximations
m
the equation of motion for the and coordinates are

(4.26)
m m

(4.27)
m m
Section 4.7 Effects of the Earth’s Rotation 103

or in vector notation

m m
where is the length of the pendulum We can find a solution by combining the two equations: define the
complex quantity

then

m m m

m m

were and We look for solutions of the form

The roots are

where

The general solution is

In particular, if we set we get

or in terms of components
(4.28)
(4.29)
Initially the oscillation is in the direction. As time progresses, the amplitude decreases while
the amplitude increases. When the relationship reverses. The solution represents an
oscillation of amplitude in a plane rotating with angular speed and the period At
the poles the angular speed of the amplitude is while at the equator it is infinite. At the poles it takes 24
hours to complete a revolution while at a latitude of o it takes about hours.
At the poles viewed in the inertial reference frame, it is clear that the pendulum must swing in a fixed
direction in space while the earth rotates beneath it. Thus relative to the earth, the oscillation plane must
rotate around the vertical (polar axis) with angular speed
104 Chapter 4 Noninertial Reference Frames

Exercises
1. Derive equation (4.7) using rectangular coordinates.
2. Find the velocity relative to an inertial reference frame (in which the centre of the earth is at rest) of
a point on the earth’s equator. An aircraft is ying above the equator at km h relative to the
earth Assuming that it ies straight and at a constant altitude of km above the surface, what is its
speed relative to the inertial frame (a) if it ies due north, (b) if it ies west, (c) if it ies east?
Use the methods discussed in this section.
[ ms ; ms ms ]
3. Estimate the centrifugal force (per unit mass) due to the earth’s revolution about the sun and compare it
to the centrifugal force (per unit mass) due to the earth’s rotation about its axis. (Take the ratio of the
two forces.)
[ ]
4. A bug crawls outward at constant speed (relative to the wheel) on the top surface along a spoke of a
wheel which is rotating at a constant angular velocity about a vertical axis. Treat the axis as a inertial
reference system.
(a) Find all the forces, real and inertial, acting on the bug.
(b) How far can the bug crawl before it starts to slip, given the coefficient of friction between bug and
spoke is
(c) If the bug does not stay on the top surface, can it crawl any further?
5. A cockroach crawls with constant speed in a circular path of radius on a turntable rotating at
a constant angular speed The circular path is concentric with the centre of the turntable. The
coefficient of static friction is and the mass of the cockroach is Assume that the centre of the
turntable is an inertial reference point.
(a) What are the real and inertial forces acting on the cockroach?
(b) How fast, relative to the turntable, can the cockroach crawl before it starts to slip when
i. it crawls in the direction of rotation,
ii. it crawls opposite to the direction of rotation?
(c) For a given and what is the maximum value of for which the bug can crawl in the direction
of rotation without slipping?
Hint: Use cylindrical coordinates [
6. Consider a cylindrical container on the surface of the earth, filled with a liquid and spinning about a
vertical axis with constant angular velocity . Ignore cohesive forces between atoms, but take into
account that a molecule on the surface experiences a normal force due to the presence of the other
molecules in the liquid. Assume that the axis of rotation is at rest in an inertial reference frame. Let the
­axis of a Cartesian coordinate system point in the vertical direction and show that the surface of the
liquid has a shape given by the paraboloid of revolution

constant.

Hint: Allow the system to come to equilibrium and consider the forces acting on a particle on the
surface. Use cylindrical coordinates. Consider potential energy of a particle on the surface in the
moving system.
7. The earth rotates to the east, yet a projectile launched from the surface of the earth tends to drift to the
east. Explain why you think this is correct (or not).
8. A bead of mass moves in a smooth circular tube of radius The plane of the tube is vertical and the
tube rotates about its vertical diameter. Find the equations of motion.
105

Chapter 5
Lagrangian and Hamiltonian
Mechanics
5.1 Generalised Coordinates and Degrees of Freedom
For a system of particles we need coordinates to specify the positions of all the particles, the
configuration of the system, at a given time. If constraints are imposed, the number may be less. For
example, for a rigid body we only need to specify the coordinates of the centre of mass and the orientation
relative to the centre of mass. In this case only six coordinates such as the position of the centre of mass
and the Euler angles are necessary to determine the position and orientation of the body.
The degrees of freedom of a system is the minimum number of independent quantities required to specify
the position and orientation of a system. We shall designate the minimum number of coordinates to
specify the position of a system by the symbols

called generalised coordinates. A given coordinate may be an angle or a distance. For a sphere
constrained to roll on a perfectly rough plane, five coordinates are required, two for the position of the
centre of mass, and three for the orientation of the sphere. If the sphere rolls without slipping, at least two
coordinates must change, so not all the coordinates can change independently, the number of degrees of
freedom is 4. If the sphere is allowed to slip on the surface, there will be five degrees of freedom.
For a point particle the Cartesian coordinates can be expressed as functions of the generalised
coordinates:
one degree of freedom­motion on a curve
two degrees of freedom­motion on a surface

three degrees of freedom­motion in space

As an example consider the motion of a particle constrained to move in a plane. Choose polar coordinates

Then

and

giving the corresponding small changes in and for small changes in and

In the general case, suppose that the ’s change from their initial values to the neighbouring
106 Chapter 5 Lagrangian and Hamiltonian Mechanics

values The corresponding changes in the Cartesian coordinates are

The partial derivatives are functions of the ’s.


For a system consisting of particles with generalised coordinates we have the general relation

(5.1)

where represents any of the Cartesian coordinates and any of the generalised coordinates. So
we map the notation In the new notation,
for example, is equivalent to the coordinate of particle 2. We do this to make writing expressions
more compact.
We will also use a corresponding subscript notation for the masses of particles. The mass of particle 1,
will be denoted by and to correspond to the Cartesian coordinates for particle 1.
The mass of particle 2 will be written as and and so on. You will soon get used to the notation
which we will only use in the formal introduction to Lagrangian and Hamiltonian Mechanics.

5.2 Generalised Forces


When a particle undergoes a displacement under the action of a force the work done by the force on a
single particle is

(5.2)

The last expression holds for any number of particles, with the appropriate components of the forces. In
terms of generalised coordinates

and by reversing the order of the summation, we have

This can be written as

(5.3)
Section 5.3 Lagrangian Equations of Motion 107

where

(5.4)

The quantity is called the generalised force associated with the coordinate Since the product
has the dimensions of work, has the dimensions of force if is a distance and the dimensions of
torque if is an angle. Note that is the partial derivative with respect to of the work done on the
system.
It is frequently best to find from the context rather than from the definition. For example, if the
system is a rigid body that turns through an angle about a given axis, then the associated generalised
force is the component of the torque of all forces about that axis ,

5.2.1 Generalised Forces for Conservative Systems


For a conservative system the force acting on a particle can be expressed in terms of the potential function
:

The generalised force then becomes

(5.5)

For example in polar coordinates, and For a central force where


is independent of and

Exercise 13 Prove that Eq. (5.5) is correct.

5.3 Lagrangian Equations of Motion


In this section we shall reformulate Newton’s second law of motion in terms of the generalised coordinates
and velocities. The new set of equations are called the Lagrangian Equations of Motion and contains no
new information, but allows us to approach the description of the motion of a system of particles in an
alternate manner.
The kinetic energy of particles can be written as

(5.6)

and therefore

is simply a different way of writing Newton’s equation of motion. Note that we write the equation of
108 Chapter 5 Lagrangian and Hamiltonian Mechanics

motion as scalar equations. In terms of generalised coordinates we have

(5.7)

We have included a possible explicit dependence on time in the equations. The are also functions of the
and possibly time:

From Eq. (5.7) it follows that

(5.8)

Multiply Eq. (5.8) by and differentiate with respect to :

(5.9)

or

(5.10)

where we use the fact that the order of differentiation with respect to and or can be reversed. Next
multiply by and set to get

(5.11)

Summing over we find

(5.12)

and finally from the definition of generalised forces we get

(5.13)

These are the differential equations of motion of the generalised coordinates and are known as the
Lagrange equations of motion.
For conservative forces, all the and we can write

(5.14)

We define the Lagrangian function as


Section 5.4 Some Applications of Lagrange’s Equations 109

where it is understood that and are expressed in terms of the generalised coordinates. Since

we have

and

For a conservative system, from Eq. (5.14), Lagrange’s equations can be written as

(5.15)

If part of the generalised forces are non­conservative, say , and part comes from a conservative
potential function , then we can write

(5.16)

With the Lagrangian still defined as the general differential equations of motion can be written
as

(5.17)

This form of the equations are convenient when dissipative forces such as friction are present.
We see that Lagrange’s equations yield second order differential equations.

5.4 Some Applications of Lagrange’s Equations


The general procedure is :
Find a suitable set of coordinates to represent the configuration of the system.
Obtain the kinetic energy as a function of these coordinates and their time derivatives.
If the system is conservative, find the potential energy in terms of the generalised coordinates,
and/or, if the system has non­conservative forces, find the generalised forces
Apply the differential equations of motion, Equations (5.13), (5.15) or (5.17).

5.4.1 Harmonic Oscillator


Consider a one­dimensional harmonic oscillator with linear damping (proportional to velocity). The
system is non­conservative. If is the displacement coordinate, then the Lagrangian function is

in which is the mass and the stiffness parameter. Hence

The system is non­conservative, and the generalised non­conservative force is

The Lagrangian equation of motion, Eq. (5.17) reads


110 Chapter 5 Lagrangian and Hamiltonian Mechanics

or

This is the familiar equation of the damped harmonic oscillator.

5.4.2 Single Particle in a Central Field


A particle moves in a plane under a central force. Let

The relevant partial derivatives are as follows:

The equations of motion, Eq. (5.15), are therefore

These are identical to the equations found before.

5.4.3 Particle Sliding on a Movable Inclined Plane


A block is free to slide on a smooth inclined plane, which in turn is free to slide on a smooth horizontal
surface.

Figure 5.1. Sliding block on sliding wedge.

There are two degrees of freedom, so we need two coordinates to specify the configuration. Choose the
horizontal displacement of the wedge, and the displacement of the block, along the inclined plane.
From the velocity diagram

and the kinetic energy is


Section 5.4 Some Applications of Lagrange’s Equations 111

The potential energy of the system is

constant

and

constant

The equations of motion are

which yields two equations:

(5.18)
(5.19)

We can eliminate from these equations

(5.20)

We can find an expression for by eliminating

Written in a differenc form

Note that is negative and is positive as expected.


An interesting question is what path does the block follow? Let represent the position
of the block as measured in a coordinate system fixed the smooth horizontal surface. The horizontal
component of the acceleration of the block is

The vertical component of the acceleration of the block is


112 Chapter 5 Lagrangian and Hamiltonian Mechanics

The acceleration of the block is constant and therefore the block moves in a straight line if the system
(block plus wedge) is released from rest.

Analysis of forces and reactions yield the same result, but is much more tedious.

Exercise 14 Show that you find the same results if you analyse the system using forces and reactions. In
other words, when you apply Newton’s laws of motion to the system.

5.5 Generalised Momenta


Consider the motion of a single particle. The kinetic energy is

Now rather than defining momentum components of the particle as we could define as the
quantity

If the system is conservative we can write this as

In the case of a system described by the generalised coordinates

the quantities

(5.21)

are defined as the generalised momenta.


Lagrange’s equations for a conservative system can be written as

(5.22)

5.6 Hamilton’s Variational Principle and Euler’s


Equations
Hamilton proposed this principle in 1834. It states that the trajectory that the system follows
from an initial configuration at time to a final configuration
at time is that for which the integral (for a conservative system)

assumes an extreme value (is an extremum for the path of motion).


is the Lagrangian function of the system. The inte­
gral is called the action.
Section 5.5 Generalised Momenta 113

In other words, Hamilton’s principle states that out of all possible ways a system can change in a given
finite time interval, the motion will occur in a manner that makes the above integral a maximum or
a minimum. If this is true, the actual path taken according to Hamilton’s principle must be consistent with
Newton’s Laws of motion.
Let be a trajectory from to Let
be a neighbouring trajectory also from to This means that the variations (changes in the
coordinates) are small and that
If is evaluated at an extremum, it means that if we change any of the functions by an infinitesimal
amount from the extreme value, the change in the action integral must be zero. The difference between the
action integrals for two trajectories is

The variation in is defined as the change in to first order in

(5.23)

Figure 5.2. Illustrating the variation of .

We can represent the family of functions as

where the and are the functions that yield the extremum and is any function that
vanishes at and

Then we can set


114 Chapter 5 Lagrangian and Hamiltonian Mechanics

We now have

where the last equation follows from So

(5.24)

But the are arbitrary independent functions, and we can set all but one equal to zero at a time. Hence
the integrand must be zero term by term and

for all

These are the familiar Lagrange equations.


Hamilton’s principle can be generalised to non­conservative forces by replacing by where
is the work done by all forces, including non­conservative forces.

5.7 Homogeneous Functions and the Kinetic Energy

Euler’s Theorem
A function is homogeneous of degree if
(5.25)

If is a homogeneous function of degree then

(5.26)

Proof:
Let then
(5.27)
(from definition of homogeneity).
Differentiate with respect to

(5.28)

(5.29)

Multiply on both sides by


Section 5.8 Conservation Laws 115

Homogeneous Functions and the Kinetic Energy


Recall that for a system with degrees of freedom, the position vectors can be expressed in terms of the
generalised coordinates

Therefore

and hence

Note: The coefficients depend on the ’s. In particular, if no explicit


time dependence, is quadratic in the ’s.
So the kinetic energy has a part that is homogeneous of degree 2 in a part that is homogeneous of
degree 1, and a part that is homogeneous of degree 0,

5.8 Conservation Laws


Conservation laws result from symmetry of the system.

5.8.1 Energy Conservation


This results from the homogeneity in time. If the Lagrangian of the system does not explicitly depend on
time, then

Consider a conservative system, then


116 Chapter 5 Lagrangian and Hamiltonian Mechanics

Hence

(5.30)

and the quantity

the energy of the system, is conserved. By Euler’s theorem

and

For simple dynamic systems is quadratic in the ’s,

and then

5.8.2 Momentum Conservation


Consider the case when the Lagrangian is invariant under any parallel displacement of the system. That is
the Lagrangian is invariant under

where is an infinitesimal constant displacement and Hence

Hence, since is an arbitrary constant, we can set and leave non­zero to get

and similarly for the other components. Since , the vector

the total momentum of the system, is conserved when the Lagrangian is invariant under infinitesimal
constant displacements.
Section 5.8 Conservation Laws 117

Example
A system of free particles.

is in invariant under , hence

is conserved.

Conservation of angular momentum


Consider the case when the Lagrangian is invariant under rotation about an axis parallel to the unit vector
That is the Lagrangian is invariant under

where is an infinitesimal constant.

(5.31)

since

(5.32)

From

and since

it follows that

and the component of the angular momentum in the direction of is conserved.

5.8.3 Cyclic or Ignorable Coordinates


A generalised coordinate that is not contained in the Lagrangian is called a cyclic or ignorable coordinate.
118 Chapter 5 Lagrangian and Hamiltonian Mechanics

Suppose in particular that one of the coordinates, say is not explicitly contained in Then

(5.33)

and

constant (5.34)

The generalised momentum associated with an ignorable coordinate is therefore a constant of the system.

Example
In the problem of the particle sliding down the smooth inclined plane, we found that the coordinate the
position of the wedge, was not contained in the Lagrangian function. Thus is an ignorable coordinate and

constant.

We can see, as a matter of fact, that is the total horizontal component of the linear momentum of the
system, and since there are no external horizontal forces acting in the system, the horizontal component of
the linear momentum is conserved.
Another example is the case of a particle in a central field. In polar coordinates

In this case is an ignorable coordinate and

constant

Here is the scalar component of the angular momentum.

5.9 The Hamiltonian Function


The Hamiltonian of a system is defined as

(5.35)

where is the Lagrangian of the system,

Example 1
Particle in a potential.
Section 5.9 The Hamiltonian Function 119

Hence

If we write

we can write the Hamiltonian as

This is the form of the Hamiltonian that is suitable for generating the corresponding quantum mechanical
Hamiltonian operator.

Example 2
Particle in a central potential. Motion in a plane, use plane polar coordinates.

Hence

Note that is a cyclic coordinate and therefore is constant. Of course is simply the angular
momentum of the system.

Note that in both cases the Hamiltonian is equal to the energy of the system. This is no coincidence. In
section 5.8 we saw from symmetry considerations that

is conserved if the Lagrangian does not depend on time explicitly. This is precisely the Hamiltonian
function. It follows that if the Hamiltonian is not an explicit function of time, it is a constant of motion
which implies that the total energy of the system is conserved.
120 Chapter 5 Lagrangian and Hamiltonian Mechanics

By Euler’s theorem

and since

the Hamiltonian can also be written as

5.10 Hamilton’s Equations


Suppose now that the equations

have been solved for the ’s in terms of the ’s and ’s:

We can now express as a function of the ’s and ’s :

(5.36)

Then

Hence

(5.37)

This gives us a set of first order differential equations in and are known as
Hamilton’s conical equations of motion. They consist of first order differential equations in contrast to
Lagrange’s equations where we found second order differential equations.
Each pair is said to form a canonical conjugate pair. So is the conjugate momentum of
Section 5.11 Poisson Brackets 121

Example 1
Particle in a potential.

The first set of equations simply reproduce the relationship between velocity and momentum. Using the
first set of equations, the second set can be written as

which, of course, is NII.

Example 2
If the variable does not appear in the Hamiltonian, that is is an ignorable coordinate, then

and the conjugate momentum, is conserved.


Particle in a central potential. Motion in a plane, use plane polar coordinates.

Here is an ignorable coordinate and the conjugate momentum

angular momentum,

is conserved.

5.11 Poisson Brackets


Let

Then we define the Poisson bracket of and as

(5.38)

Note
122 Chapter 5 Lagrangian and Hamiltonian Mechanics

Then

(5.39)
Note:

If does not explicitly depend on time, and then is constant as discussed in section 5.10.
For independent of time, hence constant
A formalism of quantum mechanics exist in which

PB commutator bracket

where is Planck’s constant.


Exercises 123

Exercises
Use Lagrange’s method in the following unless stated otherwise.
1. Find the Lagrangian equations of motion and the differential equation of motion for each degree of
freedom for a particle in a uniform gravitational field.
2. Find the general Lagrangian equations of motion and the differential equation of motion for each
degree of freedom for a particle in cylindrical coordinates: Use

3. Find the general Lagrangian equations of motion and the differential equation of motion for each
degree of freedom for a particle in spherical coordinates: Use

Problems
4. Find the acceleration of the centre of mass of a solid uniform sphere constrained to roll in a straight
line down a fixed inclined plane that makes an angle with the horizontal.
(a) a perfectly rough surface (no sliding), [ ]
(b) rolling and slipping. [ ]
[Show that the kinetic energy is equal to the kinetic energy of the motion of the centre of mass plus the
kinetic energy of rotation about the centre of mass, rotation where is the moment of inertia
(see chapter 6) and the angular velocity component about the axis of symmetry of the sphere. Refer
to chapter one to find a general expression for the kinetic energy.]
5. Two blocks of mass and are connected by an inextensible cord of length . The first block is
placed on a smooth horizontal table, the other block hangs over the edge over a frictionless pulley of
negligible mass. Find the acceleration and the motion of the system (starting from rest). Assume that
the horizontal part of the cord is taught.
(a) The mass of the cord is negligible, is a constant and an appropriately chosen
generalised coordinate ]
(b) The cord has mass . [ and are constants ]
(c) Show that in the limit that the solution of (b) reduces to the solution of (a).
6. Atwood’s Machine. A mechanical system known as Atwood’s machine consists of two weights of mass
and connected by a light inextensible cord of length which passes over a pulley of radius .
The moment of inertia of the pulley is and friction in the axle is negligible. The string does not slip.
124 Chapter 5 Lagrangian and Hamiltonian Mechanics

(a) Show that the kinetic energy of the system is given by

(b) Show that the potential energy can be expressed as

constant

where
(c) Show that the Lagrangian of the system is

(d) Show that

7. Set up the equations of motion for the double Atwood machine shown below. Ignore the mass of the
pulleys and the strings and assume that the strings do not slip. Assume that the strings are inextensible.

8. Re­examine the block­on­wedge in section 5.4.3, but now include friction between the block and
wedge. Find the path that the block follows if the block­wedge system is released from rest.
9. A particle slides on a smooth inclined plane whose inclination is increasing at a constant rate If
at time , at which time the particle starts from rest, find the subsequent motion of the
particle. Assume that the particle moves in a vertical plane.
[ , and are constants ]
10. .
Exercises 125

(a) Obtain the Lagrangian of a simple pendulum of mass and length with a second mass at
the point of support. Mass is free to move on a horizontal track lying in the plane in which
moves. Ignore the mass of the string that connects and
(b) Show that the equations of motion are:

(c) For small oscillations, show that

constant

and hence that describes a simple harmonic motion. Calculate the period of the oscillations.
[ ]
11. A particle of mass moves on a smooth horizontal table and is attached to a second particle of mass
by an inextensible string of length which passes through a small hole in the table so that
hangs vertically. If are polar coordinates of with respect to show that
(a)
1
1 2 ( is a constant)

(b)
1
1 2 constant

Hint: Show that constant


(c) If at and show that moves in a circle on the table.
12. A bead of mass moves in a smooth circular tube of radius the plane of the tube being vertical.
The tube rotates about its vertical diameter.
(a) Show that the Lagrangian is

where is the angle that the radius to the bead makes with the upward vertical, the angular
velocity of the tube and the moment of inertia of the tube about the axis of rotation.
(b) Find two conserved quantities and interpret their meanings.
(c) i. Show that if angular velocity is a constant at then

constant

Hint: Consider the bead as a separate system.


ii. What does constant imply for the forces acting on the system?
iii. What is physical quantity does the expression in (i) represent? Does it surprise you that this
quantity is conserved?
13. A truck of total mass (including wheels and axles) has two axles carrying wheels of radius the
moment of inertia of each axle and its wheels being Show that the acceleration of the truck when
rolling without slipping down a plane inclined at an angle with the horizontal is
126 Chapter 5 Lagrangian and Hamiltonian Mechanics

14. A uniform solid sphere of mass rolls down the rough surface inclined at an angle to the horizontal
of a wedge of mass that is free to move on a smooth horizontal surface. If the moment of inertia of
the sphere about the centre of mass is where is the radius of the sphere, show that the
acceleration of the wedge is

15. Find the differential equations of motion for an ”elastic pendulum”: a particle of mass attached to an
elastic string of stiffness and un­stretched length Assume that the pendulum is constrained to
move in a vertical plane. Ignore mass of elastic.
16. The point of support of a simple pendulum is elevated at a constant acceleration Find the differential
equation of motion for small oscillations of the pendulum. Show that the period of the pendulum is

where is the length of the pendulum All masses apart from the mass of the pendulum bob are
negligible.
17. The point of support of a simple pendulum is moved horizontally with constant acceleration Find the
equation of motion.
18. A particle of mass is constrained to move on the inside surface of a smooth cone whose axis points
upwards and whose semi­vertical angle is Use spherical coordinates.
(a) Find the equations of motion.
(b) Show that the particle will oscillate between two horizontal circles.
Hint: Use spherical coordinates with constant. Show that has two roots that
define the limits between which the particle must remain.

19. Use Hamilton’s principle to show that two Lagrangians differing by a total time derivative yield the
same equations of motion. Hence, in classical mechanics, the Lagrangian of a system is defined up to
a total time derivative.
20. Write down the Hamiltonian for
(a) a simple Atwood machine,
(b) a simple pendulum,
(c) a projectile in a uniform gravitational field.
21. Consider a particle of mass moving in one dimension under the action of a linear restoring force.
Obtain the Hamiltonian of the system and solve the equations of motion.
22. A particle of mass is constrained to move on the inside surface of a smooth cone whose axis points
upwards and whose semi­vertical angle is Use spherical coordinates and obtain the Hamiltonian of
the system. Show that circular motion is possible and determine the corresponding angular velocity.
23. (a) Show that if a coordinate is ignorable, the conjugate momentum is conserved, by computing the
Poisson bracket.
Exercises 127

(b) Show that the equations

are equivalent to Hamilton’s equations of motion.


(c) Show that

[ is the Kronecker delta function: and if ]


(d) Show that

24. Consider the Hamiltonian of a one dimensional simple harmonic oscillator.

Define

(a) Show that

(b) Use the results of 23(c) and 23(d) to show that

(c) Show that

(d) Show that this is in agreement with the solution to the harmonic oscillator’s equation of motion.
25. Consider a symmetric top of mass moving about a fixed point. Let the distance from the fixed point
to the centre of mass be
(a) Show that the kinetic energy of the system, in terms of Euler angles, is

and that the Lagrangian is

(b) Use the fact that and are cyclic coordinates to show that

constant

constant

(c) Show that the kinetic energy is a quadratic function of the generalised momenta and hence that the
Hamiltonian is

Also show that the total energy is a constant of motion.


(d) Show that
128 Chapter 5 Lagrangian and Hamiltonian Mechanics

which can be thought of as a one dimensional energy equation

eff

with

eff

(e) Show that


eff

as or and hence that it must have a minimum for


(f) Describe the general type of motion, namely
i. When eff remains fixed corresponding to a steady precession and for
eff oscillates between a minimum and a maximum
ii. If is outside the range then the axis precesses round the vertical
and wobbles up and down between and
iii. If then the axis moves in loops.
iv. If then the loops shrink to cusps.
129

Appendix A
Coordinate Systems and Kinematic
Variables
1.1 Coordinate Systems and Vectors
To specify a point in three dimensional space we need an origin and a system of three axes. Any
point can then be expressed in terms of three coordinates within this reference system and any vector
as a sum of three products of scalars and unit vectors . The only
restriction is that the unit vectors must not be coplanar and no two unit vectors must be parallel. It is
convenient to choose the axes to be mutually perpendicular and to choose the directions of the unit vectors
to form a right­handed triad:

(A.1)

Then

(A.2)

1.1.1 Cartesian Coordinate System


In the Cartesian or rectangular coordinate system any point can be represented by the three coordinates
( and any vector can be expressed in terms of its scalar components along the mutually perpendicular
and axes and the three unit vectors .

(A.3)

Figure 1.1. Cartesian and spherical coordinate systems.


130 Appendix A Coordinate Systems and Kinematic Variables

1.1.2 Change of Coordinate System


Consider a vector

expressed in terms of a right­handed triad Relative to a new triad the same vector
can be expressed as

Now we have

(A.4)

The scalar products are called the coefficients of transformation. There is a similar expression for
the scalar components of the unprimed vector in terms of the primed scalar components of the primed
vector:

(A.5)

The transformation can conveniently be expressed in matrix form, for example

(A.6)

and

(A.7)

The matrix is called the transformation matrix.


Note that rotating the coordinate system (axes) is equivalent to a rotation of the vector in the unprimed
system in the opposite sense.

Figure 1.2. Rotation of the primed axes through an anle about the ­axis.

The transformation matrix for a rotation of the primed coordinate system through an angle around the
131

­axis follows from the figure:

All other scalar products are zero. So the transformation matrix is

It is evident from the above (convince yourself that it is true) that a transformation about a different axis,
say the ­axis through an angle , will be given be the matrix

A combination of a rotation around the ­axis through an angle followed by a rotation around the new
­axis through an angle is given by the matrix product

1.1.3 Spherical Coordinate System


Any vector in a given representation can be written as

In a Cartesian coordinate system the terms are the direction


cosines of the vector and and are the direction angles. Note that is a unit
vector in the direction of
In the spherical coordinate system each point can be represented the three coordinates Any
vector can be written as a product of a radial distance and the radial vector
associated with that point
(A.8)
Two more unit vectors are required to complete the representation of a right­handed triad, and as
shown in Figure 1. The unit vectors in the spherical coordinate system can be obtained from the Cartesian
system by a rotation of the rectangular unit vectors around the ­axis through an angle followed by a
rotation around the new ­axis through an angle . So (see Figure 1):

(A.9)

1.1.4 Cylindrical Coordinates


In cylindrical coordinates any point is specified by the coordinates and any vector as
132 Appendix A Coordinate Systems and Kinematic Variables

Figure 1.3. Unit vectors for cylindrical coordinates.

A third unit vector, is required to complete the right­handed triad and These unit vectors can
be generated by rotating the Cartesian unit vectors through and angle about the ­xis.

(A.10)

1.2 Position Vector of a Particle


In a given reference system the position of a particle can be specified by a vector, the position vector,
which, in rectangular coordinates, is normally given by the expression:

The the vector and its scalar components of the position vector of a moving particle are functions of time

1.3 Velocity Vector


The velocity vector is given by the derivative of the position vector with respect to time:

(A.11)

where the dots indicate differentiation with respect to time.


Let be the position at and the position vector after a time interval . The quotient
is a vector parallel to the displacement in the time interval In the limit
Thus expresses the direction and the rate of motion. The velocity vector is always tangent
to the path of motion.
Section 1.4 Acceleration Vector 133

Figure 1.4. Displacement vector of a moving particle.

The magnitude of velocity is speed. In Cartesian coordinates this reduces to:

The scalar distance along the path is

Then

1.4 Acceleration Vector


The time derivative of the velocity is called the acceleration

In rectangular coordinates
(A.12)
The double dot indicates the second derivative with respect to time. The acceleration vector is tangent to
the path the velocity vector follows as a function of time.

1.5 Vector Integration


Suppose the time derivative of the position vector is given by

(A.13)
134 Appendix A Coordinate Systems and Kinematic Variables

then it is possible to integrate with respect to to get

(A.14)

where is a vector constant. The position vector can be obtained from the velocity vector to within
integration constants. If we know what the integrals are and what the position at any time is, we can
determine the position at any other time.

1.6 Relative Displacement, Velocity and Acceleration


Let two particles have position vectors and respectively. The displacement of the second with respect
to the first is Velocity of the second with respect to the first is given by

(A.15)

which is called the relative velocity. The relative acceleration is

(A.16)

The relative vectors can be interpreted as the result of measurements made by an observer located at the
position
Note that in general

(A.17)

1.7 Derivatives of Products of Vectors


Let be a scalar, then

(A.18)

Similar arguments show that

(A.19)

(A.20)

1.8 Tangential and Normal Components of Acceleration


Any vector can be expressed as the product of its magnitude and a unit vector giving its direction. The
Section 1.6 Relative Displacement, Velocity and Acceleration 135

velocity is always tangent to the path of the particle, and can be expressed in terms of its speed and the
unit vector that gives the direction of travel:

is called the unit tangent vector. The speed and direction may change with time so

(A.21)

Figure 1.5. Unit tangent and unit normal vectors.

A particle is initially at point at time and at time later it is at with unit tangent vectors and
respectively. The directions of the unit vectors differ by an angle where is the angle swept out by
the particle in time (Verify this.) For small angles

becomes tangent to the circle centred at and hence is perpendicular to the direction in the limit that
and the distance from to approach zero. is of constant magnitude and hence it follows
that

(A.22)

is of magnitude unity and direction perpendicular to is the unit normal vector. Now

(A.23)

where

is the radius of curvature of the path of the moving particle at the point With this result

(A.24)

The acceleration has a component

(A.25)
136 Appendix A Coordinate Systems and Kinematic Variables

in the direction of motion and a component

(A.26)

perpendicular to the direction of travel. The normal component is always directed toward the centre of
curvature of the path of motion. Hence it is called the centripetal acceleration.
The time derivative of the speed gives the magnitude of the tangential component of the acceleration.
The magnitude of the acceleration is given by

(A.27)

For example if a particle moves on a circle of radius with constant speed , the acceleration vector
is of magnitude and always points towards the centre of the circle. If the speed is not constant the
acceleration has a tangential component.

1.9 Velocity and Acceleration in Plane Polar Coordinates


For motion in a plane it is sometimes convenient to use plane polar coordinates The position of a
particle can then be expressed as a product of a radial distance, and a unit radial vector

Figure 1.6. Unit vectors for plane polar coordinates.

In terms of Cartesian coordinates


(A.28)
As the particle moves, both and may change as a function of time, thus

As the direction of changes by a small angle the corresponding change of the unit radial vector
is as follows: and the direction of is approximately perpendicular to Let be the
unit vector perpendicular to Then
Section 1.9 Velocity and Acceleration in Plane Polar Coordinates 137

and

so

(A.29)

We also have

and

(A.30)

Equations (A.29) and (A.30) can also easily be derived by taking the derivatives of the unit vectors
expressed in terms of the Cartesian unit vectors.
It follows that
(A.31)
where is the magnitude of the radial component and the magnitude of the transverse component.
We can also derive the expression for the velocity without resorting to geometry. Note that

(A.32)

(A.33)

From equation (A.28)

Combined with equation (A.33) we get


(A.34)
as before.
The acceleration is given by

Using
138 Appendix A Coordinate Systems and Kinematic Variables

we get

(A.35)

The magnitude of the radial component is

and that of the transverse component

(A.36)

This result shows that if the particle moves in a circle of constant radius then and the acceleration
has a magnitude directed inward towards the centre of the circular path. The magnitude of the
transverse component is then

1.10 Velocity and Acceleration in Spherical and


Cylindrical Coordinates
Spherical coordinates:

In spherical coordinates the position is written as for plane polar coordinates:

The direction of is now specified by two angles and Two more unit vectors and complete the
right­handed triad as illustrated in Figure 1.
These unit vectors can be expressed in terms of the Cartesian unit vectors:

By taking time derivatives and using the chain rule as we did for the plane polar unit vectors, we can show
that

(A.37)

(A.38)

Cylindrical coordinates

In cylindrical coordinates, , the position is written as (Figure 3)

and
(A.39)
Section 1.11 Angular Velocity 139

(A.40)

1.11 Angular Velocity


Let a particle undergo a displacement produced by a rotation around an axis defined by the unit vector

Figure 1.7. Displacement produced by rotation. The radius of the circular path is

The particle moves along an arc of a circle of radius where is the angle between and The
magnitude of displacement and the direction is perpendicular to both and Thus the
displacement can be expressed as a vector product

The velocity of the particle is given by

Let be the angular velocity. Then

(A.41)

Now consider an infinitesimal rotation about axis followed by an infinitesimal rotation about
axis

If the angular rotations are both small we can neglect the product and the expression becomes

If the order of the rotations is reversed, we find the same result, so The two rotations are
commutative. If we divide by and take the limit, we get

(A.42)

for the velocity of the particle where and


140 Appendix A Coordinate Systems and Kinematic Variables

Exercises
1. Given two vectors and find
(a)
(b)
(c)
(d)
(e)
(f)
2. Given find
(a)
(b) and
(c) and
(d) and
3. Find the angle between the vectors and
4. Prove the vector identity
5. Show that the magnitude of a vector is unchanged by a rotation.

[Hint: Use the matrix for a rotation about the axis through an angle

6.
(a) Find the velocity,
(b) acceleration,
(c) and speed as a function of time.
(d) Plot the motion in the plane and indicate the velocity and acceleration vectors on the graph.
7.
(a) Find the distance from the origin as a function of time.
(b) Show that the speed is constant.
(c) Show that the acceleration is perpendicular to the velocity.
(d) What kind of motion does describe? Sketch the motion and indicate the position, velocity and
acceleration vectors. ( , and are positive constants)
8. The velocity of a particle is given by where and are constants.
Find the position vector and acceleration vector
9. A particle has position vector and a second particle a position vector
Find
(a) the relative velocity,
(b) magnitude of the relative velocity,
(c) the value of and compare with b.
10. Prove Equation (A.17).
11. A wheel of radius rolls along the ground without slipping with constant forward speed Show that
the velocity of any point on the rim relative to the ground can be expressed as

(A.43)

where the ­axis is horizontal and the ­axis vertical.


Hint: Consider relative motion.
Exercises 141

12. Express the plane polar unit vectors and in terms of the Cartesian unit vectors and derive the
expressions for the velocity acceleration in terms of and and time derivatives of and
13. A particle moves in a spiral path in a plane such that its position in plane polar coordinates is given by

where and are constants. Find the velocity and acceleration as a function of
14. In spherical polar coordinates show that
(a)

(b)

(c)

15. In cylindrical coordinates show that


(a)

(b)

(c)

In cylindrical coordinates show that


(d)
142 Appendix A Coordinate Systems and Kinematic Variables

(e)

(f)

16. A particle moves along a helical path such that its position, in cylindrical coordinates, is given by

Find the speed and magnitude of acceleration as a function of


17. Prove that the velocity is always tangential to the path a particle follows.
18. Show that the magnitude of the tangential component of the acceleration is given by

and the normal component is

19. Prove that and hence that for a moving particle and are perpendicular to each other if the
speed is constant.
[ ]
20. Prove that where is the radius of curvature of the path of the moving particle.
143

Appendix B
Conics
2.1 Cartesian Form
Conic sections or conics refer to curves whose equation in Cartesian coordinates is quadratic in and
The most general conic has an equation of the form
(B.1)
where , and are constants. By choosing the axes appropriately we can always reduce this
equation to a simpler form.
The quadratic part, can always be reduced to the form

(B.2)
by an appropriate rotation of the axes. Let us forget about the original coordinate system and drop the
primes.
Let us assume that and are both non­zero. then we can choose the origin, (adding constants to
and for example ) so as to make and vanish. If is also non­zero, we can move it
to the other side of the equation and divide by to get the standard form of the equation,
(B.3a)
is a degenerate case: if and have the same sign, the only solution is if they have
opposite signs, the equation can be written as equations for two straight lines.
It is not possible for both and in (B.1) to be negative: the equation would then have no solutions.
We distinguish three cases:
Case 1: Both and are positive. Without loss of generality we can assume that Defining
new positive constants and by and we arrive at the canonical form of the
equation

(B.4)

This is the equation of an ellipse.

Here is the semi­major axis and the semi­minor axis. In the special case where we have
a circle of radius
Case 2: and have opposite signs. Again, without loss of generality, we can assume that and
So defining and we get

(B.5)

the equation for a hyperbola. We still call the semi­major axis and the semi­minor axis, but is not
necessarily larger than
144 Appendix B Conics

This curve has two branches on opposite sides of the origin. At large distances, this curve approaches the
two lines asymptotically.
Case 3: The third possibility is that one of the constants or vanishes. Let us say
We have to go back to our original equation where we can shift the origin to eliminate cannot be
zero otherwise will not appear in the equation. We can shift the origin in the ­direction to make zero,
and finally with we get
(B.6)
which is the equation for a parabola.

2.2 Polar Form


Consider the equation in plane polar coordinates

(B.7)

then
(B.8)
Squaring this expression leads to the
(B.9)
Case 1:
We can write (B.9) as

(B.10)
Section 2.2 Polar Form 145

Dividing by leads to

(B.11)

with Equation (B.11) describes an ellipse with origin .


Case 2:
Dividing (B.10) by leads to

(B.12)

where we now have This is a hyperbola centred at


Case 2:
The equation can be written as
(B.13)
which is a parabola with its apex at and directed towards the negative ­direction.
In the expression
(B.14)
is called the eccentricity and the semilatus rectum. In each of the above three cases the origin is at one
focus of the conic. In cases 1 and 2, there is a second focus symmetrically placed on the other side of the
centre; for the parabola the second focus is at infinity.
146 Appendix C

Appendix C

3.1 Elliptic Integrals


The elliptic integral of the first kind is given by the expression

(C.1)

and the elliptic integral of the second kind by

(C.2)

These integrals converge for They are called incomplete if and complete if
The complete integrals have the following series expansions:

(C.3)

3.2 Curvilinear coordinates


Consider a general orthogonal system of coordinates and with unit vectors and The
volume element is then

(C.4)

and the line element is

(C.5)

The gradient, divergence and curl are

grad (C.6)

div (C.7)
Section 3.3 Tensors 147

curl (C.8)

The functions for commonly used coordinate systems are :

Rectangular Coordinates: x, y, z
(C.9)
Cylindrical Coordinates:

(C.10)

Spherical Coordinates:

Parabolic Coordinates: u, v,

v (C.11)

Example: The curl in spherical coordinates is

curl (C.12)

3.3 Tensors
Scalars and vectors are the first two members of a family of objects known collectively as tensors, and are
described by components. In three dimensional space a scalar has components and
a vector has components. Scalars and vectors are called tensors of valence (or rank) and of
valence respectively. Tensors of valence components are called diadics .
Diadics occur in three dimensional space when one vector is related linearly to another vector
according to the equation

(C.13)

The nine elements of the matrix are the components of the tensor T. As an extension of the dot
product of two vectors, we may write (C.13) as

We can take the ordinary dot product with respect to a third vector
148 Appendix C

to obtain a scalar. In general

In fact

(C.14)

where is a transposed tensor, obtained form T by changing the row and column indices :
A tensor is symmetric if T = and antisymmetric (or skew­symmetric) if T = ­
A useful tensor is the unit tensor

(C.15)

3.3.1 Eigenvalues and Diagonalisation of a Symmetric Tensor


Let T be a symmetric tensor. A vector is an eigenvector of T with eigenvalue if

or equivalently The condition for a non­trivial solution is that the determinant of the
coefficients vanish :

(C.16)

This equation is called the secular equation . It is a cubic equation in In general the roots are either
real, or else one is real and the other two from a complex conjugate pair. For a real symmetric tensor
the roots are all real and the eigenvectors can be made orthogonal. Once the eigenvalues are known, the
eigenvectors can be calculated by substituting it back into Eq. (C.16).
To prove this, suppose that and are two eigenvalues and and are the corresponding
eigenvectors. Then

(C.17)
(C.18)

Here we use

The adjoint of the last expression is

or
(C.19)

4 The adjoint is obtained form T by changing the row and column indices and taking the complex conjugate: ! ¤
Section 3.3 Tensors 149

since T is Hermitian (real symmetric). Subtracting (C.19) from (C.17) , we obtain

For this becomes

Since

we conclude that

or is real for all


For and

or

which means that eigenvectors of distinct eigenvalues are orthogonal. For the degenerate case (
the eigenstates are not automatically orthogonal, but can be made orthogonal. If then

so is also an eigenstate of With the appropriate choice of the coefficients, orthogonal


vectors can be generated with the Gram­Schmidt method, for example. This is valid for any number of
degenerate states.
Eigenvector are usually normalised to unity:

(C.20)

3.3.2 Evaluation of Determinants


For a detailed discussion see for example Mathematical Methods for Physicists, G Arfken, Academic Press.
The determinant

can be evaluated according to the Laplacian development of minors as

(C.21)
Index
complete, 146
acceleration fist kind, 70
normal, 136 first kind, 146
tangential, 135 incomplete, 146
amplitude, 35 second kind, 146
angular frequency, 35, 37 energy equation, 74
angular momentum, 11, 70 external forces, 18
angular velocity, 139
antisymmetric, 148 focus, 76, 78, 145
aphelion, 77 force field, 13
apogee, 77 Fourier sum, 48
areal velocity, 71 frame of reference, 1
astronomical unit, 84 frequency, 35
generalised coordinates, 105
beats, 47 generalised force, 107
central force, 63 generalised momenta, 112
centre of mass, 9, 19 geocentric lattitude, 98
centre of mass coordinates, 21 gradient, 146
centrifugal potential, 74 gravitational potential, 66
centripetal acceleration, 136 Green’s function, 51
centripital acceleration, 94 half­width, 45
characteristic time, 7 Hamilton’s equations, 120
coefficients of transformation, 130 harmonic approximation, 32
configuration, 105 harmonic force, 42
conic section, 76 hyperbola, 143
conic sections, 143
conservative force, 4, 13 ignorable coordinate, 117
Coriolus acceleration, 94 impact parameter, 85
Coriolus force, 97 inertia, 2
curl, 146 inertial frames, 1
cyclic coordinate, 117 inertial mass, 2
cyclone, 102 inertial term, 91
initial conditions, 34
degrees of freedom, 105 internal forces, 18
diadics, 147
direction angles, 131 kinetic energy, 4, 13
direction cosines, 131 Lagrangian function, 108
divergence, 146 limit cycle, 59
line integral, 13
eccentricity, 76, 145 linear momentum, 2, 20
eigenvalue, 148 linear mometum, 11
eigenvector, 148
ellipse, 143 mass, length and time, 2
elliptic integral mechanical energy, 14

150
moment of a force, 12 semi­minor axis, 143
semilatus rectum, 76, 79, 145
natural frequency, 39 separable, 17
simple harmonic motion, 29
particle, 1 Simple harmonic oscillator, 35
perigee, 77 speed, 133
perihelion, 77 superposition principle, 34
period, 35 symmetric, 148
periodicity, 29
phase difference, 42 tensors, 147
phase space, 52 terminal velocity, 7
physical quantity, 2 torque, 12
positon vector, 132 total energy, 4
potential, 14 transformation matrix, 130
potential energy, 4, 14 transient, 42, 43
principle of relativity, 1 transposed, 148
transverse acceleration, 94
quality factor, 40, 45 turning point, 75
resonance, 41 uniform sphere, 66
resonance frequency, 43 unit normal vector, 135
right­handed triad, 129 unit tangent vector, 135
rigid body, 1 unit vectors, 129
units, 2
scattering angle, 80
scattering centre, 85 valence, 147
secular equation, 148 velocity vector, 132
self­limiting, 59
semi­major axis, 143 work principle, 13

151

Potrebbero piacerti anche