Upgrade Your Physics PDF
Upgrade Your Physics PDF
Upgrade Your Physics PDF
UPGRADE
YOUR PHYSICS
A. C. MACHACEK
Page 1
Introduction
The International Physics Olympiad is an annual international physics
competition for pre-university students. Teams of five from each
participating nation attend, and recently over 60 countries have taken
part. Each nation has its own methods for selecting its team members.
In Britain, this is by means of a series of written and practical exams.
The question paper for the first round is circulated to all secondary
schools.
Once the team has been chosen, it is necessary for its members to
broaden their horizons. The syllabus for the International Physics
Olympiad is larger than that of the British A2-level, and indeed forms a
convenient stepping-stone to first year undergraduate work. For this
reason, training is provided to help the British team bridge the gap.
The British Olympiad Committee recognizes the need for teaching
material to help candidates prepare for the international competition.
Furthermore, this material ought to have greater potential in the hands of
students who wish to develop their physics, even if they have no desire
to take part in the examinations.
It is my hope that these notes make a start in providing for this need.
A.C. Machacek, 2001
About the author: Anton Machacek has served on the British Physics
Olympiad Committee since 1997, and has been involved regularly in
training the British team and in writing Physics Challenge examinations.
He served on the academic committee for the International Physics
Olympiad in Leicester in 2000. Anton is Head of Physics at the Royal
Grammar School, High Wycombe, and is an Academic Visitor in the subdepartment of Atomic and Laser Physics, University of Oxford.
Page 2
Contents
1
1.1
1.2
1.3
1.4
2
2.1
2.2
2.3
2.4
2.5
3
ROTATION........................................................................................................37
4.1
4.2
4.3
4.4
OSCILLATION ................................................................................................53
WAVES & INTERFERENCE .............................................................................55
RAYS.............................................................................................................68
FERMATS PRINCIPLE ....................................................................................69
Page 3
4.5
5
QUESTIONS ...................................................................................................69
6.1
6.2
6.3
6.4
7
8.1
8.2
8.3
8.4
8.5
9
7.1
7.2
7.3
7.4
7.5
8
APPENDIX.......................................................................................................132
9.1
9.2
1 Linear Mechanics
1.1 Motion in a Line
1.1.1
The Fundamentals
1.1.1.1 Kinematics
Mechanics is all about motion. We start with the simplest kind of motion
the motion of small dots or particles. Such a particle is described
completely by its mass (the amount of stuff it contains) and its position.
There is no internal structure to worry about, and as for rotation, even if it
tried it, no-one would see. The most convenient way of labelling the
position is with a vector r showing its position with respect to some
convenient agreed stationary point.
If the particle is moving, its position will change. If its speed and
direction are steady, then we can write its position after time t as
r = s + ut,
where s is the starting point (the position of the particle at t=0) and u as
the change in position each second otherwise known as the velocity. If
the velocity is not constant, then we cant measure it by seeing how far
the object goes in one second, since the velocity will have changed by
then. Rather, we say that u how far the object would go in one second if
the speed or direction remained unchanged that long. In practice, if the
motion remains constant for some small time (called t), and during this
small time, the particles position changes r, then the change in position
if this were maintained for a whole second (otherwise known as the
velocity) is
u = r number of t periods in one second = r t.
Similarly, if the velocity is changing, we define the acceleration as the
change in velocity each second (if the rate of change of acceleration
were constant. Accordingly, our equation for acceleration becomes
a = u t.
Hopefully, it is apparent that as the motion becomes more complex, and
the t periods need to be made shorter and shorter, we end up with the
differential equations linking position, velocity and acceleration:
d
r r u dt
dt
.
d
a u u a dt
dt
u
Page 5
d mu
du
m
ma .
dt
dt
His next assumption tells us more about forces and allows us to define
mass properly. Imagine two bricks are being pulled together by a
strong spring. The brick on the left is being pulled to the right, the brick
on the right is being pulled to the left.
Newton assumed that the force pulling the left brick rightwards is equal
and opposite to the force pulling the right brick leftwards. To use more
mathematical notation, if the force on block no.1 caused by block no.2 is
called f12, then f12=f21. If this were not the case, then if we looked at
the bricks together as a whole object, the two internal forces would not
cancel out, and there would be some left over force which could
accelerate the whole object. 1
It makes sense that if the bricks are identical then they will accelerate
together at the same rate. But what if they are not? This is where
Newtons second law is helpful. If the resultant force on an object of
If you want to prove that this is ridiculous, try lifting a large bucket while standing in it.
Page 6
a 2 m1
and so the more massive block accelerates less. This is the definition
of mass. Using this equation, the mass of any object can be measured
with respect to a standard kilogram. If a mystery mass experiences an
acceleration of 2m/s2 while pushing a standard kilogram in the absence
of other forces, and at the same time the kilogram experiences an
acceleration of 4m/s2 the other way, then the mystery mass must be 2kg.
When we have a group of objects, we have the option of applying
Newtons law to the objects individually or together. If we take a large
group of objects, we find that the total force
Ftotal Fi
i
d
mi u i
dt i
changes the total momentum (just like the individual forces change the
individual momenta). Note the simplification, though there are no fij in
the equation. This is because fij + fji = 0, so when we add up the forces,
the internal forces sum to zero, and the total momentum is only affected
by the external forces Fi.
1.1.1.3 Energy and Power
Work is done (or energy is transferred) when a force moves something.
The amount of work done (or amount of energy transformed) is given by
the dot product of the force and the distance moved.
W = F r = F r cos
(1)
where is the angle between the force vector F and the distance vector
r. This means that if the force is perpendicular to the distance, there is
no work done, no energy is transferred, and no fuel supply is needed.
If the force is constant in time, then equation (1) is all very well and
good, however if the force is changing, we need to break the motion up
into little parts, so that the force is more or less constant for each part.
We may then write, more generally,
W = F r = F r cos
Two useful differential equations can be formed from here.
Page 7
(1a)
W
F,
r
where W is the work done on the object. Accordingly, we can calculate
the force on an object if we know the energy change involved in moving
it. Lets give an example.
An electron (with charge q) is forced through a resistor (of length L) by a
battery of voltage V. As it goes through, it must lose energy qV, since V
is the energy loss per coulomb of charge passing through the resistor.
Therefore, assuming that the force on the electron is constant (which we
assume by the symmetry of the situation), then the force must be given
by W / d = qV / L. If we define the electric field strength to be the force
per coulomb of charge (F/q), then it follows that the electric field strength
E = V/L.
So far, we have ignored the sign of F. It can not have escaped your
attention that things generally fall downwards in the direction of
decreasing [gravitational] energy. In equations (1) and (1a), we used the
vector F to represent the externally applied force we use to drag the
object along. In the case of lifting a hodful of bricks to the top of a wall,
this force will be directed upwards. If we are interested in the force of
gravity G acting on the object (whether we drag it or not), this will be in
the opposite direction. Therefore F = G, and
W = G r,
G
(1b)
W
.
r
In other words, if an object can lose potential energy by moving from one
place to another, there will always be a force trying to push it in this
direction.
1.1.1.5 Power
Another useful equation can be derived if we differentiate (1a) with
respect to time. The rate of working is the power P, and so
W F r
r
F .
t
t
t
As we let the time period tend to zero, r/t becomes the velocity, and so
we have:
Page 8
(2)
where is now best thought of as the angle between force and direction
of motion. Again we see that if the force is perpendicular to the direction
of motion, no power is needed. This makes sense: think of a bike going
round a corner at constant speed. A force is needed to turn the corner thats why you lean into the bend, so that a component of your weight
does the job. However no work is done you dont need to pedal any
harder, and your speed (and hence kinetic energy) does not change.
Equation (2) is also useful for working out the amount of fuel needed if a
working force is to be maintained. Suppose a car engine is combating a
friction force of 200N, and the car is travelling at a steady 30m/s. The
engine power will be 200N 30m/s = 6 kW.
Our equation can also be used to derive the kinetic energy. Think of
starting the object from rest, and calculating the work needed to get it
going at speed U. The force, causing the acceleration, will be F=ma.
The work done is given by
W P dt F v dt m
mv dv
1
2
mv
2 U
0
dv
v dt
dt
mU
1
2
(3)
1.1.2
Changing Masses
the change in speed dv is equal to |dv| cos where dv (note the bold
type) is the vector giving the change in velocity.
Page 9
After
m
U-w
U+u
Notice that the velocity of the burnt fuel is U-w, since w is the speed at
which the combustion gas leaves the rocket (backwards), and we need
to take the rocket speed U into account to find out how fast it is going
relative to the ground.
Conservation of momentum tells us that
Page 10
(4)
m w = M u.
dM
dU
M
(5)
1
dM dU
M
wln M U
M
U final U initial w ln initial
M
final
(6)
T M
u
mw mw
m
M
w w
t
Mt
t
t
(7)
given by the product of the exhaust speed and the rate of burning fuel.
For a rocket of total mass M to take off vertically, T must be greater than
the rockets weight Mg. Therefore for lift off to occur at all we must have
w Mg .
(8)
This explains why heavy hydrocarbon fuels are nearly always used for
the first stage of liquid fuel rockets. In the later stages, where absolute
Page 11
1.1.3
Fictitious Forces
(ii) Centrifugal force is only considered if you are assuming that the
cylinder is at rest (in the cylinders reference frame). On the other hand,
you only have centripetal accelerations if you do treat the cylinder as a
moving object and work in the reference frame of a stationary observer.
This example also shows that fictitious forces generally act in the
opposite direction to the acceleration that is being ignored. Here the
Page 12
(9)
In other words, if working in the reference frame of the lift, you need to
include not only the forces which are really acting on the ball (like
gravity), but also an extra force mA. This extra force is the inertial
force.
Let us continue this line of thought a little further. Suppose the only
force on the ball was gravity. Therefore F=mg. Notice that
F = F mA = m (g-A)
(10)
and therefore if g=A (that is, the lift is falling like a stone, because some
nasty person has cut the cable), F=0. In other words, the ball behaves
as if no force (not even gravity) were acting on it, at least from the point
of view of the unfortunate lift passengers. This is why weightlessness is
experienced in free fall.
Page 13
Fr
GMm
R2
(11)
GMm
GMm
GMm
R F dx R Fr dr R r 2 dr r R R .
E ( R)
GMm
.
R
Page 14
(12)
Fr ( R)
dE ( R)
.
dR
(13)
E ( R, m)
GM
.
m
R
(14)
MG
R2
(15)
g ( R)
1.2.2
dV
.
dR
(16)
Orbital tricks
Page 15
mu 2
Fr AR n
R
mu 2 AR n 1
2
2
(17)
E ( R)
AR n 1
n 1
(18)
if we take the usual convention that E(R) is zero when the force is zero.
Combining equations (17) and (18) gives
mu 2 n 1
E ( R)
2
2
(19)
(20)
so that
This tells us that for circular gravitational orbits (where n=2), the
potential energy is twice as large as the kinetic energy, and is negative.
For elliptical orbits, the equation still holds: but now in terms of the
average 3 kinetic and potential energies. Equation (20) will not hold
instantaneously at all times for non-circular orbits.
1.2.3
Keplers Laws
The motion of the planets in the Solar system was observed extensively
and accurately during the Renaissance, and Kepler formulated three
laws to describe what the astronomers saw. For the Olympiad, you
wont need to be able to derive these laws from the equations of gravity,
but you will need to know them, and use them (without proof).
1.
All planets orbit the Sun in elliptical orbits, with the Sun at one
focus.
By average, we refer to the mean energy in time. In other words, if T is the orbital period,
1
T
A(t )dt .
Page 16
The area traced out by the radius of an orbit in one second is the
same for a planet, whatever stage of its orbit it is in. This is
another way of saying that its angular momentum is constant, and
we shall be looking at this in Chapter 3.
3.
The time period of the orbit is related to the [time mean] average
3
radius of the orbit: T R 2 . It is not too difficult to show that this
is true for circular orbits, but it is also true for elliptic ones.
1.2.4
Large Masses
In our work so far, we have assumed that all masses are very small in
comparison to the distances between them. However, this is not always
the case, as you will often be working with planets, and they are large!
However there are two useful facts about large spheres and spherical
shells. A spherical shell is a shape, like the skin of a balloon, which is
bounded by two concentric spheres of different radius.
1.
2.
These rules only hold if the sphere or shell is of uniform density (strictly
if the density has spherical symmetry).
Therefore let us work out the gravitational force experienced by a miner
down a very very very deep hole, who is half way to the centre of the
Earth. We can ignore the mass above him, and therefore only count the
bit below him. This is half the radius of the Earth, and therefore has one
eighth of its mass (assuming the Earth has uniform density which it
doesnt). Therefore the M in equation (11) has been reduced by a factor
of eight. Also the miner is twice as close to the centre (R has halved),
and therefore by the inverse-square law, we would expect each kilogram
of Earth to attract him four times as strongly. Combining the factors of
1/8 and 4, we arrive at the conclusion that he experiences a gravitational
field that at the Earths surface, that is 4.9 N/kg.
1.3.1
(21)
Upthrust = V g
Volume V
Mass M
Fluid
Density
Weight = M g
Therefore, things float if their overall density (total mass / total volume) is
less than the density of the fluid. Notice that the overall density may not
be equal to the actual density of the material. To give an example a ship
is made of metal, but contains air, and is therefore able to float because
its overall density is reduced by the air, and is therefore lower than the
density of water. Puncture the hull, and the air is no longer held in place.
Therefore the density of the ship = the density of the steel, and the ship
sinks.
For an object that is floating on the surface of a fluid (like a ship on the
ocean), the upthrust and weight must be equal otherwise it would rise
Page 18
1.3.2
Under Pressure
What is the pressure in a fluid? This must depend on how deep you are,
because the deeper you are, the greater weight of fluid you are
supporting. We can think of the pressure (=Force/Area) as the weight of
a square prism of fluid above a horizontal square metre marked out in
the depths.
Pressure = Weight of fluid over 1m2 square
= g Density Volume of fluid over 1m2
= g Density Depth Cross sectional area of fluid (1m2)
Pressure = g Density Depth
(22)
1.3.3
Continuity
Page 19
10 cm
5 cm
1.3.4
Bernoullis Equation
Something odd is going on in that pipe. As the water squeezes into the
smaller radius, it speeds up. That means that its kinetic energy is
increasing. Where is it getting the energy from? The answer is that it
can only do so if the pressure in the narrower pipe is lower than in the
wider pipe. That way there is an unbalanced force on the fluid in the
cone-shaped part speeding it up. Lets follow a cubic metre of water
through the system to work out how far the pressure drops.
The fluid in the larger pipe pushes the fluid in the cone to the right. The
force = pressure area = PL AL. A cubic metre of fluid occupies length
1/AL in the pipe, where AL is the cross sectional area of the pipe to the
left of the constriction. Accordingly, the work done by the fluid in the
wider pipe on the fluid in the cone in pushing the cubic metre through is
Force Distance = PL AL 1/AL = PL. However this cubic metre does
work PR AR 1/AR = PR in getting out the other side. Thus the net
energy gain of the cubic metre is PL PR, and this must equal the
change in the cubic metres kinetic energy uR2/2 uL2/2.
Page 20
1.3.5
Equation (23) is also useful in the context of electric currents, and can be
adapted into the so-called flow equation. Let us suppose that the fluid
contains charged particles. Suppose that there are N of these particles
per cubic metre of fluid, and each particle has a charge of q coulombs,
then:
Current = Flow rate of charge (charge / second)
= Charge per cubic metre (C/m3) flow rate (m3/s)
= N q Area Speed .
(24)
Among other things, this equation shows why the free electrons in a
semiconducting material travel faster than those in a metal. If the
semiconductor is in series with the metal, the current in both must be the
same. However, the free charge density N is much smaller in the
semiconductor, so the speed must be greater to compensate.
1.4 Questions
1.
2.
3.
A trolley can move up and down a track. Its potential energy is given by
V = k x4, where x is the distance of the trolley from the centre of the
track. Derive an expression for the force exerted on the trolley at any
point. +
4.
5.
Use arguments similar to equation (3) to prove that the kinetic energy is
still given by 12 mu 2 even when the force which has caused the
acceleration from rest has not been applied uniformly in a constant
direction. +
6.
Page 21
8.
9.
10. Architectural models can not be properly tested for strength because
they appear to be stronger than the real thing. To see why, consider a
half-scale model of a building made out of the same materials. The
weight is 1/8 of the real building, but the columns are the cross
sectional area. Accordingly the stress on the columns is half of that in
the full size building, and accordingly the model can withstand much
more severe load before collapsing. To correct for this, a 1:300
architectural model is put on the end of a centrifuge arm of radius 10m
which is spun around.
The spinning simulates an increased
gravitational force which allows the model to be accurately tested. How
many times will the centrifuge go round each minute?
11. Consider an incompressible fluid flowing from a 15cm diameter pipe into
a 5cm diameter pipe. If the velocity and pressure before the constriction
are 1m/s and 10 000 N/m2, calculate the velocity and pressure in the
constricted pipe. Neglect the effects of viscosity and turbulent flow. To
work out the new pressure, remember that the increase in speed
involves an increase of kinetic energy, and this energy must come from
somewhere so there will be a drop in pressure.
12. Calculate the orbital time period T of a satellite skimming the surface of a
planet with radius R and made of a material with density . Calculate
the orbital speed for an astronaut skimming the surface of a comet with a
10km radius.
13. The alcohol percentage in wine can be determined from its density. A
very light glass test tube (of cross sectional area 0.5cm2) has 5g of lead
pellets fixed to the bottom. You place the tube in the wine, lead first, and
it floats with the open end of the tube above the surface of the wine.
You can read the % alcohol from markings on the side of the tube.
Calculate how far above the lead the 0% and 100% marks should be
placed. The density of water is 1.00g/cm3, while that of ethanol is
Page 22
Page 23
2 Fast Physics
Imagine a summers day. You are sunbathing by the side of a busy
motorway while you wait for a pickup truck to rescue your car, which has
broken down. All of a sudden, an irresponsible person throws a used
drinks can out of their car window, and it heads in your direction. To
make things worse, they were speeding at the time. Ouch.
The faster the car was going, the more it will hurt when the can hits you.
This is because the can automatically takes up the speed the car was
travelling at. Suppose the irresponsible person could throw the can at
10mph, and their car is going at 80mph. The speed of the can, as you
see it, is 90mph if it was thrown forwards, and 70mph if it was thrown
backwards.
To sum this up,
Velocity as measured by you = Velocity of car + Velocity of throwing
where we use velocities rather than speeds so that the directionality can
be taken into account.
So far, this probably seems very obvious. However, lets extend the
logic a bit further. Rather than a car, let us have a star, and in place of
the drinks can, a beam of light. Many stars travel towards us at high
speeds, and emit light as they do so. We can measure the speed of this
light in a laboratory on Earth, and compare it with the speed of ordinary
light made in a stationary light bulb. And the worrying thing is that the
two speeds are the same.
No matter how hard we try to change it, light always goes at the same
speed. 4 This tells us that although our ideas of adding velocities are
nice and straightforward, they are also wrong. In short, there is a
problem with the Newtonian picture of motion. This problem is most
obvious in the case of light, but it also occurs when anything else starts
travelling very quickly.
While this is not the way Einstein approached the problem, it is our way
into one of his early theories the Special Theory of Relativity and it is
part of the Olympiad syllabus.
Before we go further and talk about what does happen when things go
fast, please be aware of one thing. These observations will seem very
4
Light does travel different speeds in different materials. However if the measurement is
made in the same material (say, air or vacuum) the speed registered will always be the same,
no matter what we do with the source.
Page 24
2.
3.
4.
We say non-accelerating for a good reason. If the laboratory were accelerating, you would
feel the inertial force, and thus you would be able to measure this acceleration, and indeed
adjust the laboratorys motion until it were zero. However there is no equivalent way of
measuring absolute speed.
Page 25
2.2.1
A metre stick comes hurtling towards you at high speed. With a clever
arrangement of cameras and timers, you are able to measure its length
as it passes you. If the sticks length is perpendicular to the direction of
travel, you still measure the length as 1 metre.
However, if the stick is parallel to its motion, it will seem shorter to you.
If we call the sticks actual length (as the stick sees it) as L0, and the
apparent length (as you measure it) La, we find
L a L0 1
u2
,
c2
(1)
where u is the speed of the metre stick relative to the observer. The
object in the square root appears frequently in relativistic work, and to
make our equations more concise, we write
1
1 u c
(2)
2.2.2
L0
(3)
A second observation is that if a clock whizzes past you, and you use
another clever arrangement of timers and cameras to watch it, it will
appear (to you) to be going slowly.
We may state this mathematically. Let T0 be a time interval as
measured by our (stationary) clock, and let Ta be the time interval as we
see it measured by the whizzing clock.
Ta
2.2.3
T0
(4)
Equations (3) and (4) are consistent you cant have one without the
other. To see why this is the case, let us suppose that Andrew and Betty
both have excellent clocks and metre sticks, and they wish to measure
their relative speed as they pass each other. They must agree on the
relative speed. Andrew times how long it takes Betty to travel along his
metre stick, and Betty does the same.
Page 26
According to the muon, it still has a woefully short life, but the tube
(which is whizzing past) is shorter so it appears to get further along
in the 2s.
For the two calculations to agree, the clock slowing must be at the
same rate as the tube shrinkage.
2.2.4
The most useful observation of them all, as far as the Olympiad syllabus
is concerned is this: if someone throws a 1kg bag of sugar at you at high
speed, and you (somehow) manage to measure its mass as it passes,
you will register more than 1kg.
If the actual mass of the object is M0, and the apparent mass is Ma, we
find that
Ma M0 .
(5)
The actual mass is usually called the rest mass in other words the
mass as measured by an observer who is at rest with respect to the
object.
Note that slow and short are placed in quotation marks. Bettys clock and metre stick are
not defective however to Andrew they appear to be.
Page 27
Let us pause for a moment to see why. Suppose the object concerned
is an electron in a particle accelerator (electrons currently hold the speed
record on Earth for the fastest humanly accelerated objects). It starts at
rest with a mass of about 10-30 kg. We turn on a large, constant electric
field, and the electron starts to move relative to the accelerator.
However, as it gets close to the speed of light, it starts to appear more
massive. Therefore since our electric field (hence accelerating force) is
constant, the electrons acceleration decreases. In fact, the acceleration
tends to zero as time passes, although it never reaches zero exactly
after a finite time. We are never able to persuade the electron to break
the light-barrier, since when u c , , and the apparent mass
becomes very large (so the object becomes impossible to accelerate any
further).
Please note that this does not mean that faster-than-light speeds can
never be obtained. If we accelerate one electron to 0.6c Eastwards, and
another to 0.6c Westwards, the approach speed of the two electrons is
clearly superlumic (1.2c) as we measure it with Earth-bound
speedometers. However, even in this case we find that the velocity of
one of the electrons as measured by the other is still less than the speed
of light. This is a consequence of our first observation namely that
relative velocities do not add in a simple way when the objects are
moving quickly.
In fact the approach speed, as the electrons see it, is 0.882c. If you
want to perform these calculations, the formula turns out to be
u AC
u AB u BC
1 u AB u BC c 2
(6)
d
momentum
dt
Page 28
(7)
2.3.1
Momentum
p m0 u
(7)
Notice that when you use momentum conservation in collisions, you will
have to watch the factors. Since these are functions of the speed u,
they will change if the speed changes.
2.3.2
Force
d
d
d
p m0 u u
.
dt
dt
dt
(8)
In the case where the speed is not changing, will stay constant, and
the equation reduces to the much more straightforward F=m0a. One
example is the motion of an electron in a magnetic field.
2.3.3
Kinetic Energy
If you wish to derive this yourself, here are the stages you need. Firstly, differentiate with
respect to u to convince yourself that
3u
d
u
du c 2 1 u 2 c 2 3 2
c2
du c 2
.
d 3 u
Using this result, the derivation can be completed (see over the page):
Page 29
K 1m 0 c 2 .
(9)
E K m0 c 2 m0 c 2 .
2.3.4
(10)
A Relativistic Toolkit
We can derive a very useful relationship from (10), (7) and the definition
of :
E 2 p 2 c 2 2 m02 c 2 c 2 v 2
v 2
m c 1 .
c
2 4
m0 c
2
2
0
(11)
This is useful, since it relates E and p without involving the nasty factor.
Another equation which has no gammas in it can be derived by dividing
momentum by total energy:
p mo u
u
2,
2
E m0 c
c
(12)
which is useful if you know the momentum and total energy, and wish to
know the speed.
2.3.5
Tackling problems
If you have to solve a collision type problem, avoid using speeds at all
costs. If you insist on having speeds in your equations, you will also
have gammas, and therefore headaches. So use the momenta and
energies of the individual particles in your equations instead. Put more
bluntly, you should write lots of ps, and Es, but no us. Use the
d
du
dt m 0 u
dt
dt
dt
du
c2
c2
d m 0 u u 2 d m 0 u d m0 c 2
m 0 u u
d
u
u
K Fdx Fu dt m 0 u 2
Page 30
Finally, if the question asks you for the final speeds, use (12) to calculate
them from the momenta and energies.
2.4.1
B x
A t
1 A2
.
vA
Page 32
t Cc D
Cc D c Ac B
Cc 2 B since A D
1 - A2 2
c vA
vA
1 A 2 c 2 v 2 A 2
1
A
v2
1 2
c
This concludes our reasoning, and gives the Lorentz transforms (after a
little algebra to evaluate C) as:
x x vt
xv
t t 2
c
x x vt
x v
t t 2
c
v 2
1
c
Page 33
2.4.2
L.
2.4.3
T.
Four Vectors
The Lorentz transforms show you how to work out the relationships
between the (t,x,y,z) co-ordinates measured in different frames of
reference. We describe anything that transforms in the same way as a
four vector, although strictly speaking we use (ct,x,y,z) so that all the
components of the vector have the same units. Three other examples of
four vectors are:
(c, ux, uy, uz) is called the four velocity of an object, and is the
derivative of (ct, x,y,z) with respect to the proper time . Proper
time is the time elapsed as measured in the rest frame of the object
t=.
It also turns out that the dot product of any two four-vectors is frameinvariant in other words all observers will agree on its value. The dot
product of two four-vectors is slightly different to the conventional dot
product, as shown below:
ct , x, y, z ct , x, y, z x 2 y 2 z 2 ct 2 .
Notice that we subtract the product of the first elements.
The dot product of the position four vector with the wave four vector
gives
ct , x, y, z
c , k x , k y , k z k r t .
Now this is the phase of the wave, and since all observers must agree
whether a particular point is a peak, a trough or somewhere in between,
then the phase must be an invariant quantity. Accordingly, since
(ct,x,y,z) makes this invariant when dotted with (/c,kx,ky,kz), it follows
that (/c,kx,ky,kz) must be a four vector too.
2.5 Questions
1. Work out the relativistic factor for speeds of 1%, 50%, 90% and 99% of
the speed of light.
2. Work out the speeds needed to give factors of 1.0, 1.1, 2.0, 10.0.
3. A muon travels at 90% of the speed of light down a pipe in a particle
accelerator at a steady speed. How far would you expect it to travel in 2s
(a) without taking relativity into account, and (b) taking relativity into
account?
4. A particle with rest mass m and momentum p collides with a stationary
particle of mass M. The result is a single new particle of rest mass R.
Calculate R in terms of p and M.
Page 35
Page 36
3 Rotation
Rotational motion is all around us [groan] from the acts of subatomic
particles, to the motion of galaxies. Calculations involving rotations are
no harder than linear mechanics; however the quantities we shall be
talking about will be unfamiliar at first. Having already studied linear
mechanics, you will be at a tremendous advantage, since we shall find
that each rule in linear mechanics has its rotational equivalent.
3.1 Angle
In linear mechanics, the most fundamental measurement is the position
of the particle. The equivalent base of all rotational analysis is angle: the
question How far has the car moved? being exchanged for How far
has the wheel gone round? a question which can only be answered
by giving an angle. In mechanics, the radian is used for measuring
angles. While you may be more familiar with the degree, the radian has
many advantages.
We shall start, then by defining what we mean by a radian. Consider a
sector of a circle, as in the diagram; and let the circle have a radius r.
The length of the arc, that is the curved line in the sector, is clearly
related to the angle. If the angle were made twice as large, the arc
length would also double.
Can we use arc length to measure the angle? Not as it stands, since we
havent taken into account the radius of the circle. Even for a fixed angle
(say 30), the arc will be longer on a larger circle. We therefore define
the angle (in radians) as the arc length divided by the
Arc length = r if is
circle radius. Alternatively you might say that the angle
measured in rad ians
in radians is equal to the length of the arc of a unit circle
(that is a circle of 1m radius) that is cut by the angle.
r
Before getting too involved with radians, however, we must work out a
conversion factor so that angles in degrees can be expressed in radians.
To do this, remember that a full circle (360) has a circumference or arc
length of 2r. So 360=2 rad. Therefore, 1 radian is equivalent to
(360/2) = (180/).
Page 37
Here we are not including the centripetal acceleration which is directed towards the centre of
the rotation.
Page 38
mass I by
Why force radius? We can use a virtual work argument (as in section 1.1.1.4) to help us.
Suppose a tangential force F is applied at radius r. When the object moves round by angle ,
it moves a distance d = r, and the work done by the force = Fd = Fr = Fr angular force
angular distance. Now since energy must be the same sort of thing with rotational motion
as linear, the rotational equivalent of force must be Fr.
Page 39
rF sin
F
r2 r2 m
a sin r
a
where we have used the fact that the mass m will be the ratio of the
force F to the linear acceleration a, as dictated by Newtons Second
Law. This formula can also be used for solid objects, however in this
case, the radius r will be the perpendicular distance from the mass to the
axis. The total angular mass of the object is calculated by adding up
the I = m r2 from each of the points it is made from.
Usually this angular mass is called the moment of inertia of the object.
Notice that it doesnt just depend on the mass, but also on the distance
from the point to the centre. Therefore the moment of inertia of an object
depends on the axis it is spun round.
An object may have a high angular inertia, therefore, for two reasons.
Either it is heavy in its own right; or for a lighter object, the mass is a
long way from the axis.
Thus we see that, like in linear motion, the time derivative of angular
momentum is angular force or torque. Two of the important facts that
stem from this statement are:
1.
If there is no torque C, the angular momentum will not change.
Notice that radial forces have C = 0, and therefore will not change the
angular momentum. This result may seem unimportant but think of the
planets in their orbits round the Sun. The tremendous force exerted on
them by the Suns gravity is radial, and therefore does not change their
angular momenta even a smidgen. We can therefore calculate the
velocity of planets at different parts of their orbits using the fact that the
angular momentum will remain the same. This principle also holds when
Page 40
v sin
mvr sin
r
10
Two cautions. Firstly, in a rocket, the mass of the rocket does decrease as the burnt fuel is
chucked out the back, however the total mass does not change. Therefore F=dp/dt=ma still
works, we just need to be careful that the force F acts on (and only on) the stuff included in
the mass m. A complication does arise when objects start travelling at a good fraction of the
speed of light but this is dealt with in the section on Special Relativity.
Page 41
We see that the kinetic energy is given by half the angular mass
angular velocity squared which is a direct equivalent with the half mass
speed2 of linear motion.
Symbol
Unit
Definition
Other equations
Angular velocity
rad/s
= d/dt
r = v sin
Angular
acceleration
rad/s2
= d/dt
r = a sin
Torque
Nm
C = F r sin
Moment of inertia
kg m2
I=C/
I = m r2
Angular
momentum
kg m2 /s
L=I
L = m v r sin
Rot. Kinetic
Energy
K = I 2 / 2
K = m (v sin )2
Page 42
3.10
This section involves much more advanced mathematics, and you will be
able to get by in Olympiad problems perfectly well without it. However, if
you like vectors and matrices, read on...
So far we have just considered rotations in one plane that of the paper.
In general, of course, rotations can occur about any axis, and to describe
this three dimensional situation, we use vectors. With velocity v,
momentum p and force F, there is an obvious direction the direction of
motion, or the direction of the push. With rotation, the direction is less
clear.
Imagine a clock face on this paper, with the minute hand rotating
clockwise. What direction do we associate with this motion? Up
towards 12 oclock because the hand sometimes points that way?
Towards 3 oclock because the hand sometimes points that way? Both
are equally ridiculous. In fact the only way of choosing a direction that
will always apply is to assign the rotation direction perpendicular to the
clock face the direction in which the hands never point.
This has not resolved our difficulty completely. Should the arrow point
upwards out of the paper, or down into it? After thought we realise that
one should be used for clockwise and one for anti-clockwise motion, but
which way? There is no way of proceeding based on logic we just
have to accept a convention. The custom is to say that for a clockwise
rotation, the direction is down away from us, and for anticlockwise
rotation, the direction is up towards us.
Various aides-memoire have been presented for this my favourite is to
consider a screw. When turned clockwise it moves away from you:
when turned anticlockwise it moves towards you. For this reason the
convention is sometimes called the
right hand screw rule.
r
r sin
With this convention established, we
Page 43
11
We use to represent the angle between r and , to distinguish it from the angle between
r and v, which is of course a right angle for a strict rotation.
12
This intentionally does not include the centripetal acceleration, as before. If you aim to
calculate this a from the former equation v = r, then you get a = dv/dt = d(r)/dt = r +
v = r + (r) = r + (r.) r 2. The final two terms in this equation deal with the
centripetal acceleration. However in real situations, the centripetal force is usually provided
by internal or reaction forces, so often problems are simplified by not including it.
Page 44
mr 2 mr r
r 2 x2
m xy
xz
xy
r y2
yz
xz
yz
r2 z2
y2 z2
m xy
xz
xy
x z
yz
xz
yz
x 2 y 2
C m yz
x2 y2
which is a little better. Notice that it is still pretty nasty in that the torque
required to cause this z-rotation acceleration is not necessarily in the zdirection! Another consequence of this is that the angular momentum L
is not necessarily parallel to the angular velocity . However for many
objects, we rotate them about an axis of symmetry. In this case the xz
and yz terms become zero when summed for all the masses in the
object, and what we are left with is the mass multiplied by the distance
from the axis to the masses (that is x2 + y2). Alternatively, for a flat
object (called a lamina) which has no thickness in the z direction, the xz
and yz terms are zero anyway, because z=0.
At this point, you are perfectly justified in saying yuk and sticking to twodimensional problems. However this result we have just looked at has
interesting consequences. When a 3-d object has little symmetry, it can
roll around in some very odd ways. Some of the asteroids and planetary
moons in our Solar System are cases in point.
The moment of inertia can also be obtained from the rotational
momentum, however, the form is identical to that worked out above from
Newtons second law, as shown here.
L r p mr v
mr r
mr 2 mr r
12 r mv
12 r p
12 L 12 I
For the cases where I can be simplified, this reduces to the familiar form
K = I 2/2.
3.11
When a system is rotating, it often makes sense to use polar coordinates. In other words, we characterise position by its distance from
the centre of rotation (i.e. the radius r) and by the angle it has turned
through.
Conversion between these co-ordinates and our usual
Cartesian (x,y) form are given by simple trigonometry:
x r cos
y r sin
(1)
r
r
Page 46
(2)
where we have used the dot above a letter to mean time derivative of.
Now if the particle does not change its , then the direction r will not
change either, and we have a velocity given simply by r r . We next
consider the case when r doesnt change, and the particle goes in a
circle around the origin. In this case, our formula would say that the
d r
. We know from section 3.2 that in this case, the
velocity was r
dt
speed is given by r, that is r , so the velocity will be r . In order to
make this agree with our equation for dr/dt, we would need to say that
d
r .
dt
(3)
Does this make sense? If you think about it for a moment, you should
find that it does. Look at the diagram below. Here the angle has
changed a small amount . The old and new r vectors are shown, and
form two sides of an isosceles triangle, the angle between them being
. Given that the sides r have length 1, the length of the third side is
going to be approximately equal to (with the approximation getting
better the smaller is). Notice also that the third side the vector
corresponding to rnew rold is pointing in the direction of . This allows
us to justify statement (1).
r
rnew
rold
d
r .
dt
Page 47
(4)
(5)
Now suppose that a force acting on the particle (with mass m), had a
radial component Fr, and a tangential component F. We could then
write
Fr m r r 2
.
F m r 2r
(6)
2.
If the force is purely radial (we call this a central force), like gravitational
attraction, then F= 0. It follows that
0 mr 2mr
mr 2 2mrr ,
d
mr 2
dt
(7)
3.12
When you are dealing with a rigid body, things are simplified in that it
can only do two things move in a line and rotate. If forces Fi are
applied to positions ri on a solid object free to move, its motion is
completely described by
13
This assumes that the angular acceleration is a simple speeding up or slowing down of an
existing rotation. If and are not parallel, the situation is more complex.
Page 49
d 2R
Ftotal
dt 2
and the point R moves as if it were a single point of mass M being acted
on by the total force. This position R is called the centre of mass.
Given that we already know that R does not have any rotational motion,
this must be the centre of rotation, and we can use the equation from
section 3.10 to show that the rate of change of angular momentum of the
object about this point, d(I)/dt, is equal to the total torque (ri R)Fi
acting on the body about the point R. Given that the masses dont
change, we may write
d
mi u i Fi f ij
dt
j
mi a i Fi f ij
mi ri a i ri Fi ri f ij
j
m r a r F r f
i i
ij
ij
The final term sums to zero since fij+fji=0, and the internal forces
between two particles must either constitute a repulsion, an attraction or
the two forces must occur at the same place. In any of these cases fij
(rirj) = 0.
If we now express the positions ri in terms of the centre of mass position
R and a relative position ri, where ri = R + ri (so ai = A+ai), then
Page 50
A a i R ri Fi
mi R a i mi ri A mi ri a i R Fi ri Fi
m R r
i
Rm a
i
0 mi ri a i R Fi ri Fi
m r
i i
a i ri Fi
ri u i ri Fi
dt
d
d
L I ri Fi C
dt
dt
3.13
Questions
1. A car has wheels with radius 30cm. The car travels 42km. By what angle
have the wheels rotated during the journey? Make sure that you give your
answer in radians and in degrees.
2. Why does the gravitational attraction to the Sun not change the angular
momentum of the Earth?
3. Calculate the speed of a satellite orbiting the Earth at a distance of 42
000km from the Earths centre.
4. A space agency plans to build a spacecraft in the form of a cylinder 50m in
radius. The cylinder will be spun so that astronauts inside can walk on the
inside of the curved surface as if in a gravitational field of 9.8 N/kg.
Calculate the angular velocity needed to achieve this.
5. A television company wants to put a satellite into a 42 000km radius orbit
round the Earth. The satellite is launched into a circular low-Earth orbit
200km above the Earths surface, and a rocket motor then speeds it up. It
then coasts until it is in the 42000km orbit with the correct speed. How
fast does it need to be going in the low-Earth orbit in order to coast up to
the correct position and speed?
6. Estimate the gain in angular velocity when an ice-skater draws her hands
in towards her body.
Page 51
Page 52
F Ax Bx 2 Cx 3 ,
(1)
where the minus sign indicates that the force acts in the opposite
direction to the disturbance. If x is small enough, x2 and x3 will be so
small that they can be neglected. We then have a restoring force
proportional to the displacement x.
Just because the system has a force acting to restore the equilibrium,
this does not mean that it will return to x=0 immediately. All systems
have some inertia, or reluctance to act quickly. For a literal force, this
inertia is the mass of the system and we know that the acceleration
caused by a force (F) is given by F/m, where m is the mass. We can
therefore work on equation (1) to find out more:
d 2x
F m 2 Ax
dt
.
2
d x
A
x
m
dt 2
(2)
Am
(3)
T 2
2
T
.
1
f
T 2
Page 53
(4)
4.1.1
Non-linearity
4.1.2
Energy
(5)
At the moment when the system passes through its equilibrium (x=0)
point, all the energy is in kinetic form. Therefore the total energy is
E K x x 0 12 mu 2 12 m 2 x 02
(6)
1
2
Ax 2 .
(7)
Notice that we did not include the minus sign on the force. This is
because when we work out the work done the force involved is the
force of us pulling the system. This is equal and opposite to the
restoring force of the system, and as such is positive (directed in the
same direction as x).
The total energy is given by the potential energy at the moment when x
has its maximum (i.e. x=x0). Therefore
E E pot x x 0 12 Ax 02 .
(8)
Equations (8) and (6) are in agreement. This can be shown by inserting
the value of from equation (3) into (6).
While we have only demonstrated that energy is proportional to
amplitude squared for an oscillation, it turns out that the same is true for
linked oscillators and hence for waves. The intensity of a wave (joules
of energy transmitted per second) is proportional to the amplitude
squared in exactly the same way.
Intensity of a wave is also related to another wave property its speed.
The intensity is equal to the amount of energy stored on a length u of
wave, where u is the speed. This is because this is the energy that will
pass a point in one second (a length u of wave will pass in this time).
4.2.1
Wave number
Page 55
(13)
kD .
(14)
We can see that this makes sense by combining equations (13) and
(14):
kD
2D
(15)
4.2.2
Wave equations
Page 56
4.2.3
Standing waves
2 cost 12 1 2 coskx 12 2 1
At any time, the peaks and troughs will only occur at the places where
the second cosine is +1 or 1, and so the positions of the peaks and
troughs do not change. This is why this kind of situation is called a
standing wave. While there is motion, described by the first cosine, the
positions of constructive interference between the two counterpropagating waves remain fixed (these are called antinodes), as to the
positions of destructive interference (the nodes).
While there are many situations which involve counter-propagating
waves, this usually is caused by the reflection of waves at boundaries
(like the ends of a guitar string). Accordingly, there is nothing keeping
the phase constants 1 and 2 the same, and so the standing wave
doesnt develop. However if the frequency is just right, then it works, as
indicated in section 4.2.7.5.
4.2.4
Trigonometric Interference
A cos t B cost
A cos t B cos t cos B sin t sin
A B cos cos t B sin sin t
X 0 cos cos t sin sin t
(9)
X 0 cos 0 t
where we define
X0
cos
A B cos 2 B sin 2
A 2 B 2 2 AB cos
A B cos
X0
sin
Page 57
B sin
X0
(10)
(11)
2 A cos 12
and we obtain the familiar result that if the waves are in phase (=0),
the amplitude doubles, and if the waves are radians (half a cycle) out
of phase, we have complete destructive interference.
Equation (10) can be used to provide a more general form of this
statement the minimum resultant amplitude possible is |A-B|, while the
maximum amplitude possible is A+B.
This statement is reminiscent of the triangle inequality, where the length
of one side of a triangle is limited by a similar constraint on the lengths of
the other two sides. This brings us to our second method of working out
interferences: by a graphical method.
4.2.5
Graphic Interference
In the graphic method a vector represents each wave. The length of the
vector gives the amplitude, and the relative orientation of two vectors
indicates their phase relationship. If the phase relationship is zero, the
two vectors are parallel, and the total length is equal to the sum of the
individual lengths. If the two waves are out of phase, the vectors will
be antiparallel, and so will partly (or if A=B, completely) cancel each
other out.
The diagram below shows the addition of two waves, as in the situation
above. Notice that since + = , cos = cos. One application of
the cosine rule gives
X0
A 2 B 2 2 AB cos
X
B
Page 58
(12)
4.2.6
The results of the last section allow us to determine the amplitude once
we know the phase difference between the two waves. Usually the two
waves have come from a common source, but have travelled different
distances to reach the point. Let us suppose that the difference in
distances is D this is sometimes called the path difference. What will
be?
To find out, we use the wave number k. The phase difference is given
by
kD
2D
(14)
4.2.7
To obser vation
point
D = d sin
A
C
Hard reflections occur when, at the boundary, the wave passes into
a sterner material. At these reflections, a peak (before the
reflection) becomes a trough (afterwards) and vice-versa. This is
usually stated as a phase difference is added to the wave by the
reflection. These mean the same thing since cos cos .
To visualize this imagine that you are holding one end of a rope,
and a friend sends a wave down the rope towards you. You keep
your hand still. At your hand, the incoming and outgoing waves
interfere, but must sum to zero (after all, your hand is not moving,
so neither can the end of the rope). Therefore if the incoming wave
is above the rope, the outgoing wave must be below. In this way,
peak becomes trough and vice-versa.
Soft reflections, on the other hand, are where the boundary is from
the sterner material. At these reflections, a peak remains a peak,
and there is no phase difference to be added.
Page 60
From
To
Reflection
Light
Light
Air
Water / Glass
Hard
Light
Water / Glass
Air
Soft
Light
Lower
index
Sound
Solid / liquid
Air
Soft
Sound
Air
Solid / liquid
Hard
Wave on string
Hard
Wave on string
Soft
Hard
D 2t cos n
D 2t cos 12 n
(16)
(17)
Page 61
(18)
X 0 cos1t
where
X0
cos
A B cos t 2 B sin t 2
A 2 B 2 2 AB cos t
(19)
A B cos t
X0
4.2.8
The diagram shows a section of a crystal. Light (in this case, X-rays) is
bouncing off the layers of atoms. There are certain special angles for
which all the reflections are in phase, and interfere constructively.
Looking at the small triangle in the diagram, we see that the extra path
travelled by the wave bouncing off the second layer of atoms is
D 2d sin .
(20)
Page 63
To obser vation
point
Dmax = W sin
(21)
Make sure you remember that W is the width of the gap, and that this
formula is for destructive interference.
This formula is only valid (as in the diagram) when the observer is so far
away that the two rays drawn are effectively parallel. Alternatively the
formula works perfectly when it is applied to an optical system that is
focused correctly, for then the image is at infinity.
4.2.8.4 Resolution of two objects
How far away do you have to get from your best friend before they look
like Cyclops? No offence but how far away do you have to be before
you cant tell that theyve got two eyes rather than one? The results of
diffraction can help us work this out. Lets call this critical distance L.
The rays from both eyes come into your eyeball. Let us suppose that
the angle between these rays is , where is small, and that your
friends eyes are a distance s apart. Therefore tan sin s L .
These two rays enter your eye, and spread out (diffract) as a result of
passing through the gap called your pupil. They can only just be
resolved that is noticed as separate when the first minimum of ones
diffraction pattern lines up with the maximum of the other. Therefore
W sin where W is the width of your pupil.
Page 64
L W
.
sW
(22)
For normal light (average wavelength about 500nm), a 5mm pupil, and a
10cm distance between the eyes: you friend looks like Cyclops if you are
more than 1km away! If you used a telescope instead, and the
telescope had a diameter of 10cm, then your friends two eyes can be
distinguished at a distances up to 20km.
4.2.8.5 The Bandwidth Theorem
In the last section, we asked the question, What happens when you add
lots of waves together? However we cheated in that we only
considered waves of the same frequency. What happens if the waves
have different frequencies?
Suppose that we have a large number of waves, with frequencies evenly
spread between f and f+f. The angular frequencies will be spread from
to +, where =2f as in equation (4). Furthermore, imagine that
we set them up so that they all agree in phase at time t=0. They will
never agree again, because they all have different frequencies.
The phases of the waves at some later time t will range from t to
()t.
Initially we have complete constructive interference. After a short time
t, however, we have destructive interference. This will happen when
(as stated in the last section) all phases are equally represented when
the range of phases is a whole multiple of 2. This happens when
t=2. After this, the signal will stay small, with occasional complete
destructive interference.
From this you can reason (if youre imaginative or trusting) that if you
need to give a time signal, which has a duration smaller than t, you
must use a collection of frequencies at least =2/t. This is called the
bandwidth theorem. This can be stated a little differently:
Page 65
t 2
2f t 2 .
f t 1
(23)
f 1
f
c
1
x 1
x 1 .
(24)
(25)
4.2.9
Doppler Effect
(26)
The first line is constructed like this: The travellers start a distance L0
apart, so by the time the signal is received, the distance between them is
L0 + vt1. This distance is covered by waves of speed c in time t1 hence
the right hand side.
The next wave peak will be broadcast at time 1/f one wave cycle later.
At this time, the distance between the two musicians will be
L0 v u T L0 v u / f . This second peak will be received at time
t2, where
L0
vu
1
1
v t 2 c t 2
f
f
f
.
cu
L0
c v t 2
f
(27)
Finally we can work out the time interval elapsed between our listener
hearing the two peaks, and from this the apparent frequency is easy to
determine.
t 2 t1 c v L0 c u L0
f
cu
1
t 2 t1
f
f c v
cv
f f
cu
(28)
From this we see that if v=u, no change is observed. If the two are
approaching, the apparent frequency is high (blue-shifted). If the two are
receding, the apparent frequency is low (red-shifted).
4.2.9.2 Relativistic Doppler Effect
Please note that if either u or v are appreciable fractions of the speed of
light, this formula will give errors, and the relativistic calculation must be
used.
For light only, the relativistic formula is
Page 67
f f
(29)
4.3 Rays
4.3.1
All the discussion so far has centred on the oscillatory nature of waves.
We can predict some of the things waves do without worrying about the
oscillations like reflection and refraction. The diagram below shows
both. We refer to a refractive index of a material, which is defined as
Refractive Index ( n)
(30)
Air has a refractive index of about 1.0003 14 , glass has a refractive index
of about 1.5, and water about 1.3.
n1
n2
First of all, the angle of reflection is equal to the angle of incidence (both
were labelled i in the diagram).
Secondly, the angle of incidence is related to the angle of refraction r by
the formula:
sin i n2
sin r n1
14
(31)
The refractive index also gives a measure of pressure, since n-1 is proportional to pressure.
Page 68
c c 0 0
f nf
n
(32)
D nD
c
c0
(33)
where c0 is the speed of light in vacuum; minimizing the time is the same
as minimizing the product of distance and refractive index. This latter
quantity (nD) is called the optical path. It is possible to prove the laws of
reflection and refraction using this principle. 15
4.5 Questions
1.
A hole is drilled through the Earth from the U.K. to the centre of the Earth
and out of the other side. All the material is sucked out of it, and a 1kg
mass is dropped in at the British end. How much time passes before it
momentarily comes to rest at the Australian end? (NB You may need
some hints from section 1.2.4) +
2.
15
To do this, imagine the plane as a sheet of graph paper, with the boundary along the x-axis.
Suppose that the light starts at point (0,Y), and needs to get to (X,-Y). Now assume that the
light crosses the x-axis at point (x,0). Work out the total optical path travelled along the route,
and then minimize it with respect to x. You should then be able to identify sin i and sin r in the
algebraic soup, and from this, you should be able to finish the proof.
Page 69
Show that Fermats principle allows you to derive the Law of Reflection.
Assume that you have a mirror along the x-axis. Let light start at point
(X1,Y1) and end at point (X2,Y2). Show that the least-time reflected route
is the one which bounces off the mirror where angles of incidence and
reflection will be equal. +
4.
Show that Fermats principle allows you to derive Snells Law. Assume
that you have a material with refractive index 1 for y>0 (that is, above the
x-axis), and refractive index n where y<0. Show that the shortest time
route from point (X1,Y1) to (X2,Y2), where Y1>0 and Y2<0, crosses the
boundary at the point where sin i / sin r = n. +
5.
You are the navigator for a hiking expedition in rough ground. Your
company is very thirsty and tired, and your supplies have run out. There
is a river running East-West which is 4km South of your current position.
Your objective is to reach the base camp (which is 2km South and 6km
West of your current position), stopping off at the river on the way. What
is the quickest route to the camp via the river? +
6.
7.
8.
9.
10. How far away do you need to hold a ruler from your left eye before you
can no longer resolve the millimetre markings? Keep your right eye
covered up during this experiment. Use equation (22) to make an
Page 70
Page 71
5 Hot Physics
This section gives an introduction to the areas of physics known as
thermodynamics and statistical mechanics.
These deal with the
questions What happens when things heat up or cool down? and
Why? respectively.
We start with a statement that will be very familiar but then find that it
leads us into new territory when explored further.
(1)
Page 72
(2)
Work dW provided by
electric compressor
Fridge
dQ2
Page 73
Work dW produced,
drives electrical
generator
Engine
dQ2
It was Carnot who realised that the most efficient heat engine of all was
a reversible heat engine. In other words one that got the same
amount of work out of the heat transfer as would be needed to operate a
perfect fridge to undo its operation.
In order to do this, it is necessary for all the heat transfers (between one
object and another) to take place with as small a temperature difference
as possible. If this is not done, heat will flow from hot objects to cold a
process which could have been used to do work, but wasnt. Therefore
not enough work will be done to enable the fridge to return the heat to
the hot object.
Carnot therefore proposed that the ratio of heat coming in from the hot
object to the heat going out into the cold object has a maximum for this
most-efficient engine. This is because the difference between heat in
and heat out is the work done, and we want to do as much work as
possible. Furthermore, he said that this ratio must be a function of the
temperatures of the hot and cold objects only.
This can be stated as
dQ1
f (T1 , T2 )
dQ2
(3)
where T1 is the temperature of the hot object, and T2 is that of the cold
object. More light can be shed on the problem if we stack two heat
engines in series, with the second taking the heat dQ2 from the first (at
Page 74
dQ2
f (T2 , T3 )
dQ3
5.3.1
dQ1
f (T1 , T3 )
dQ3
f (T1 , T2 )
(4)
g (T1 )
g (T2 )
Thermodynamic Temperature
Q 2 T2
(5)
The kelvin temperature scale obtained using the gas laws satisfies this
definition. For this reason, the kelvin is frequently referred to as the unit
of thermodynamic temperature.
5.3.2
dQ1 dQ2
T
dW
1 2 .
dQ1
dQ1
T1
(6)
300
64% .
840
Page 75
The design of modern large power stations is such that the actual
efficiency is remarkably close to this value.
5.4 Entropy
Now we need to take a step backwards before we can go forwards.
Look back at the definition of thermodynamic temperature in equation
(5). It can be rearranged to state
dQ1 dQ2
T1
T2
dQ1 dQ2
0
T1
T2
REVERSIBLE.
(7)
dQ
0 REVERSIBLE
T
(8)
where the circle on the integral implies that the final position (on a p,V
graph) is the same as where the gas started.
Now suppose that there are two points on the (p,V) graph which are of
interest to us, and we call them A and B. Let us go from A to B and then
back again (using a different route), but only using reversible processes.
We call the first route I, and the second route II. Equation (8) tells us
Page 76
REVERSIBLE
(9)
dQ I B dQ II
A T A T
In other words the integral of dQ/T between the two points A and B is the
same, no matter which reversible route is chosen. This is a very special
property of a function we label dQ/T as a function of state, and call it
the entropy (S).
This means that the current entropy of the gas, like pressure, volume
and temperature, is only a function of the state that the gas is in now
and does not depend on the preparation method.
Let us now return to the physics, and the gas in the piston. What does
irreversibility mean here? We havent lost any energy the First Law
has ensured that. But we have lost usefulness.
Equation (8) tells us that if we come back to where we started, and only
use reversible processes on the way, the total entropy change will be
zero. There is another way of looking at this, from the point of view of a
heat engine.
Page 77
(9)
We find that this equation is also true for irreversible processes. This is
because T, S, U, p and V are all functions of state, and therefore if the
equation is true for reversible processes, it is true for all processes.
However care must be taken when using it for irreversible processes,
since TdS is no longer equal to the heat flow, and pdV is no longer equal
to the work done.
16
(10)
Whats the pipe got to do with it? Remember that we said that change in entropy dS is only
given by dQ/T for reversible processes. The passing of Q joules of heat into the pipe is done
reversibly (at temperature TA), so we can calculate the entropy change. Similarly the passing
of Q joules of heat from pipe to cylinder is done reversibly (at TC), so the calculation is
similarly valid at the other end. However something is going on in the pipe which is not
reversible namely Q joules of heat passing from higher to lower temperature. Therefore we
mustnt apply any dS = dQ/T arguments inside the pipe.
Page 78
5.7.1
Atmospheric Pressure
kT
(11)
5.7.2
The probability that a molecule in the air will have x-component of its
velocity equal to ux is proportional to exp(-mux2/2kT). Here the energy E
is the kinetic energy associated with the x-component of motion.
From this statement, you can work out the mean value of ux2, and find it
to be:
2
x
1
2
3/ 2
1/ 2
1
2
mu x2 12 kT
Page 79
kT
m
(12)
K 12 mu 2 12 m u x2 u y2 u z2 32 kT
(13)
(14)
where R N A k is the gas constant. From this it follows that the molar
heat capacity of a perfect gas 17 , CV 32 R .
5.7.3
Vapour Pressure
The probability that a water molecule in a mug of tea has enough (or
more than enough) energy to leave the liquid is proportional to exp(EL/kT) where EL is the energy required to escape the attractive pull of
the other molecules (latent heat of vaporization per molecule).
5.7.4
These are said to be the five macrostates of the system. We can work
out how likely each one is to occur if the energy is distributed randomly
by counting the ways in which each macrostate could have happened.
17
This is the heat capacity due to linear motion. For a monatomic gas (like helium), this is the
whole story. For other gases, the molecules can rotate or vibrate about their bonds as well,
and therefore the heat capacity will be higher.
Page 80
7 ways
42 ways
21 ways
105 ways
35 ways
We can see that there is one macrostate clearly in the lead where 4
atoms have no energy, 2 atoms have one energy unit, and 1 atom has
two energy units. This macrostate is interesting because there is a
geometric progression in the number of atoms (4,2,1) having each
amount of energy.
It can be shown that the most likely macrostate will always be the one
with (or closest to) a geometric progression of populations. 19 In other
18
where W is the number of ways of setting up the macrostate, n0 is the number of atoms with
no energy, n1 is the number of atoms with one unit of energy, and so on.
Here is one justification for this formula: N! gives the total number of ways of choosing the
atoms in order. The division by n0! prevents us overcounting when we choose the same
atoms to have zero energy in a different order. A similar reason holds for the other terms on
the denominator.
Alternatively, W = the number of ways of choosing n0 atoms from N the number of ways of
choosing the n1 from the remaining (Nn0) the number of ways of choosing the n2 from the
remaining (Nn0n1) and so on. Thus
19
N n0 ! N n0 n1 !
N!
N!
n0 ! N n0 ! n1! N n0 n1 ! n2 ! N n0 n1 n2 !
n0 ! n1! n2 !
To show this, start with a geometric progression (in other words, say ni Af
where f is
some number), and write out the formula for W. Now suppose that one of the atoms with i
units of energy gives a unit to one of the atoms with j units. This means that ni and nj have
gone down by 1, while ni1 and nj+1 have each gone up by one. By comparing the old and
Page 81
(18)
pV=NkT
(19)
where p is the pressure of the gas, V its volume, and T its absolute (or
thermodynamic) temperature. This temperature is measured in kelvins
always. There are two ways of stating the equation: as in (18), where n
represents the number of moles of gas; or as in (19), where N
represents the number of molecules of gas. Clearly N=NAn where NA is
the Avogadro number, and therefore R=NAk.
You can adjust the equation to give you a value for the number density
of molecules. This means the number of molecules per cubic metre, and
is given by N/V = p/kT. The volume of one mole of molecules can also
be worked out by setting n=1 in (18):
new values of W, you can show that the new W is smaller, and that therefore the old
arrangement was the one with the biggest W.
20
The bet gets better as the number of atoms increases. The combination (4,2,1) was the
most popular in our example of N=7, P=4, however if you repeated the exercise with N=700
and P=400, you would find a result near (400,200,100) almost a certainty. In physics we deal
with huge numbers of atoms in matter, so the gambling pays off.
Page 82
RT
p
(20)
You can adjust this equation to give you an expression for the density. If
the mass of one molecule is m, and the mass of a mole of molecules
(the R.M.M.) is M, we have
Mp
Mass
M
Volume RT p RT
N mp mp
A
N A kT kT
(21)
Please note that this is the ideal gas law. Real gases will not always
follow it. This is especially true at high pressures and low temperatures
where the molecules themselves take up a good fraction of the space.
However at room temperature and atmospheric pressure, the Gas Law
is a very good model.
5.8.1
We have already shown (in section 5.7.2) that for a perfect gas, the
internal energy due to linear motion is 32 RT per mole. If this were the
only consideration, then the molar heat capacity would be 32 R . However
there are two complications
5.8.1.1 The conditions of heating
In thermodynamics, you will see molar heat capacities written with
subscripts CP and CV. They both refer to the energy required to heat a
mole of the substance (M kilograms) by one kelvin. However the energy
needed is different depending on whether the volume or the pressure is
kept constant as the heating progresses.
When you heat a gas at constant volume, all the heat going in goes into
the internal energy of the gas (dQV = dU).
When you heat a gas at constant pressure, two things happen. The
temperature goes up, but it also expands. In expanding, it does work on
its surroundings. Therefore the heat put in is increasing both the internal
energy and is also doing work (dQP = dU + p dV).
Given that we know the equation of state for the gas (18), we can work
out the relationship between the constant-pressure and constant-volume
heat capacities. In these equations we shall be considering one mole of
gas.
Page 83
dT
dT
dQ P dU
dV
CP
p
dT
dT
dT
d RT
CV R
CV p
dT p
CV
(22)
The monatomic molecule only has one use for energy going places
fast. Therefore its internal energy is given simply by 32 kT , and so the
molar internal energy is U 32 RT . Therefore, using equation (22), we
can show that CV 32 R and C P CV R 52 R .
A diatomic molecule has other options open to it. The atoms can rotate
about the molecular centre (and have a choice of two axes of rotation).
They can also wiggle back and forth stretching the molecular bond like
a rubber band. At room temperature we find that the vibration does not
have enough energy to kick in, so only the rotation and translation (the
linear motion) affect the internal energy.
Each possible axis of rotation adds 12 kT to the molecular energy, and so
we find that for most diatomic molecules, CV 52 R and C P 72 R .
5.8.1.3 Thermodynamic Gamma
It turns out that the ratio of C P CV crops up frequently in equations, and
5.8.2
Pumping Heat
Page 84
The Gas Law tells us (18) that pV=nRT, and hence that pV is a function
of temperature alone (for a fixed amount of gas). Hence in an isothermal
process
pV const .
ISOTHERMAL
(23)
Using this equation, we can work out how much we need to compress
the gas to remove a certain quantity of heat from it. Alternatively, we
can work out how much we need to let the gas expand in order for it to
absorb a certain quantity of heat. These processes are known as
isothermal compression, and isothermal expansion, respectively.
Suppose that the volume is changed from V1 to V2, the temperature
remaining T. Let us work out the amount of heat absorbed by the gas.
First of all, remember that as the temperature is constant, the internal
energy will be constant, and therefore the First Law may be stated
dQ=pdV. In other words, the total heat entering the gas may be
calculated by integrating pdV from V1 to V2:
Q pdV
V
nRT
V2
dV nRT ln V V 1 nRT ln 2 .
V1
V
(24)
21
While the terms isentropic and adiabatic are synonymous for a perfect gas, care must be
taken when dealing with irreversible processes in more advanced systems. In this context dQ
is not equal to TdS. If dQ=0, the process is said to be adiabatic: if dS=0, the process is
Page 85
nCV dT
(25)
pdV Vdp
dV dp
V
p
ln V ln p C 0
pV e C
. ADIABATIC
(26)
pV const
isentropic. Clearly for a complex system, the two conditions will be different. This arises
because in these systems, the internal energy is not just a function of temperature, but also of
volume or pressure.
Page 86
(27)
T1 V1 1
1
T2 V2 1
5.8.2.3 A Gas Heat Engine
We may now put our isothermal and adiabatic processes together to
make a heat engine. The engine operates on a cycle:
1.
2.
3.
4.
V2
V1
V
ln 4
V3
(28)
V2 Tcold
V4 Thot
V1 Tcold
V3 V4
V
V
3 2
V2 V1
V4 V1
(29)
(30)
where the minus sign reminds us that Q2<0, since this heat was leaving
the gas.
To summarize this process, we have used a perfect gas to move heat
from a hot object to a colder one. In doing this, we notice less heat was
deposited in the cold object than absorbed from the hot one. Where has
it gone? It materialized as useful work when the cylinder was allowed to
expand. Had the piston been connected to a flywheel and generator, we
would have seen this in a more concrete way.
We also notice that we have proved that the kelvin scale of temperature,
as defined by the Gas Law, is a true thermodynamic temperature since
equation (30) is identical to (5).
(31)
P is the power radiated (in watts), A is the surface area of the object (in
m2), and T is the thermodynamic temperature (in K).
5.10
Questions
1.
2.
Mechanical engineers have been keen to build jet engines which run at
higher temperatures. This makes it very difficult and expensive to make
the parts, given that the materials must be strong, even when they are
almost at their melting point. Why are they making life hard for
themselves?
Page 88
Two insulated blocks of steel are identical except that one is at 0C,
while the other is at 100C. They are brought into thermal contact. A
long time later, they are both at the same temperature. Calculate the
final temperature; the energy change and entropy change of each block
if (a) heat flows by conduction from one block to the other, and if (b) heat
flows from one to the other via a reversible heat engine. ++
4.
5.
The amount of energy taken to turn 1kg of liquid water at 100C into 1kg
of steam at the same temperature is 2.26 MJ. This is called the latent
heat of vaporization of water. How much energy does each molecule
need to free itself from the liquid?
6.
7.
Estimate the altitude of the mountaineer in q5. Assume that all of the air
in the atmosphere is at 0C. +
8.
Use the Gas Law to work out the volume of one mole of gas at room
temperature and pressure (25C, 100kPa).
9.
What fraction of the volume of the air in a room is taken up with the
molecules themselves? Make an estimate, assuming that the molecules
are about 1010m in radius.
Page 90
(1)
Q
4 0 R
1
Notice that the symbols are slightly different for gravity field is now
given E rather than g, and in consequence we have to use another letter
for energy hence our choice of U. Here we have put a charge Q at the
origin, and we measure the quantities associated with a small test
charge q at distance R. Notice that we do not have a minus sign in front
this allows like charges to repel rather than attract (whereas in gravity,
positive mass attracts positive mass [and we have never found any
lumps of negative mass if we did this would upset a lot of our
thinking]). Also, in place of the G of gravity, we have the constant
1 4 0 which is about 1020 times bigger. No wonder the theoreticians
talk of gravity as a weaker force!
Now, you may wonder, why the factor of 4? This comes about,
because the equations above are not the nicest way of describing
electrostatic forces. They are based on the Coulomb Force Law, which
is the first equation in (1) however there is another, equivalent, way of
describing the same physics, and it is called Gauss Law of
Electrostatics. This Gauss Law is on the Olympiad syllabus, and you will
find it useful because it will simplify your electrical calculations a lot.
Page 91
6.1.1
Firstly, let us say what the law is. Then we will describe it in words, and
then prove that it is equivalent to the Coulomb Law. Finally, we will
show its usefulness in other calculations.
Q
E dS
S
(2)
What does this mean? Firstly let us look at the individual symbols. S is
a closed surface (that is what the circle on the integral sign means) like
the outer surface of an apple, a table, or a doughnut but not the outer
surface of a bowl (which ends at a rim). dS is a small part of the
surface, with area dS, and is a vector pointing outward, perpendicular to
the surface at that point. Q is the total electric charge contained inside
the surface S. Finally, the vector E is the electric field (in volts per metre)
where the vector points in the direction a positive charge would be
pulled.
The odd looking integral tells us to integrate the dot product of E with dS
(a normal vector to the surface) around the complete surface. This may
sound very foreign, strange, and difficult, but let us give some examples.
First of all, suppose S is a spherical surface of radius R, with one charge
(+Q) at the centre of the surface. Assuming there are no other charges
nearby, the field lines will be straight, and will stream out radially from
the centre. Therefore E will be parallel to dS, and the dot product will
simply be E dS the product of the magnitudes. Now the size of E must
be the same all round the surface, by symmetry. Therefore
E dS E dS E dS E 4R
S
(3)
Q
4 0 R 2
(4)
E dS E dS E dS E 2RL
S
curved S
(5)
curved S
where L is the length of the cylinder, and hence 2RL is its curved
surface area. By Gauss Law, this must be equal to Q 0 L 0 , and
so
E
.
2 0 R
(6)
We notice that for a line charge, the inverse square of the Coulomb
Law has become an inverse, not squared law.
Before moving on, let us make two more points. Firstly, we have not
proved the equivalence of Gauss Law and Coulombs Law we have
only shown that they agree in the case of calculating the field around a
fixed, point charge. However the two can be proved equivalent but the
proof is a bit involved, and is best left to first (or second) year university
courses.
Secondly, let us think about a surface S which is entirely inside the same
piece of metal. E will be zero within a metal, because any non-zero E
(i.e. voltage difference) would cause a current to flow until the E were
zero. Therefore a surface entirely within metal can contain no total
charge!
Impossible, I hear you cry! Let us take a hollow metal sphere, with a
charge +Q at the centre of the cavity. How can the enclosed charge be
zero surely its +Q! Oh, no it isnt. Actually the total enclosed charge
is zero, and this enclosed charge is made up of +Q at the middle of the
cavity, and Q induced on the inside wall of the hollow sphere! If the
sphere is electrically isolated, and began life uncharged, there must be a
+Q charge somewhere on the metal, and it sits on the outer surface of
the sphere.
6.1.2
Capacitors
Area A
Separation L
Voltage
difference V
E dS
middle of the gap counts. The face buried in the metal of the plate has
E=0, while the other four surfaces normals are perpendicular to the field.
Therefore
Q
E dS AE .
(7)
We next work out the voltage. This is not hard, as by analogy from
gravitational work (chapter 1, equation 16)
Page 94
(8)
QL
0 A
Q 0 A
L
V
(9)
I
+Q
-Q
Here we need to take care. Depending on how the circuit has been
drawn, either I=dQ/dt or I=-dQ/dt. That is why it is essential that you
Page 95
dt C
.
dQ
Q
dt
RC
V IR R
(10)
(11)
where Q0 was the initial charge on the capacitor after it was charged up.
Given that the voltage across the capacitor V(t)=Q(t)/C, the time
dependence of the voltage obeys a similar equation. The constant RC is
known as the time constant, and it is the time taken for the voltage (or
charge) to fall by a factor e (approx 2.7).
6.1.2.3 Energy considerations
Next, we need to know how much energy has been stored in a capacitor,
if its voltage is V and its capacitance C. The energy is actually stored in
the electric field between the plates but more of this later.
dQ
d (CV )
dt V
dt
dt
dt
dV
VC
dt CVdV C VdV C 12 V 2
dt
1
2
CV
(12)
Given that the energy stored must be zero when V=0, and there is no
electric field in the gap, this fixes the constant of integration as zero, and
we obtain the fact that energy stored = half the capacitance the square
of the voltage across the gap. You could equally well say that the
energy is given by half the charge multiplied by the voltage.
Before we leave this formula, let us do some conjuring tricks with this
energy, assuming that the capacitor is a simple parallel plate device:
U 12 CV 2
0 AV 2
2L
0 AEL 2
2L
Page 96
12 0 E 2 AL
(13)
Page 97
6.2.1
= L (NQu) B
= (NQL) u B
= QTOT u B.
6.2.2
(14)
mu sin
BQ
(15)
u sin BQ
R
m
(16)
The overall motion is therefore a helix, or corkscrew shape, with the axis
of the corkscrew parallel to the magnetic field. The radius of the helix is
given by R in equation (15), and the pitch (that is, the distance between
successive revolutions), is equal to D = T u cos.
Page 99
6.2.3
0 I r s
dL
4 r s 3
(17)
where (r-s) is the vector that points from the bit of wire to the point at
which we are calculating B. To work out the total field B(r), we integrate
the expression along the wire. In most cases, this can be put more
simply as
dB
0 I sin
dL
4 d 2
Page 100
(18)
0 INR 2
2 R2 D2
32
(19)
B dL
(20)
In other words, if we choose a loop path, and integrate the magnetic flux
density around it, we will find out the current enclosed by that loop.
Let us give an example. This formula is very useful for calculating the
magnetic field in the vicinity of a long straight wire carrying current I.
Suppose we take path L to be a circle of radius R, with the wire at its
centre, and with the wire perpendicular to the plane of the circle, as
shown in the diagram.
Current I
We know that B points round the wire, and therefore that B is parallel
with dL (dL being the vector length of a small part of the path).
Page 101
B dL B dL B dL 2RB
L
(21)
I
B 0
2R
Integration path
B-field direction
Only one of the sides of the rectangle counts in the integration the
one completely inside. Of the other three, one is so far away from the
coil that there is no magnetic field, and the other two are perpendicular
to the B-field lines, so the dot product is zero. The rectangle encloses
nL turns, and hence a current of nLI. Therefore, Amperes Law tells us:
BL 0 nLI
(22)
B 0 nI
22
1
0
dS
B dL I 0 rel E
S
where the surface S is any surface that has as its edge the loop L. This reduces to equation
(20) when E is constant.
Page 102
6.2.4
(23)
The total amount of magnetic field (the flux) is proportional to the current,
and we call the constant of proportionality L the self-inductance (or
inductance for short). Any coil (or wire for that matter) will have an
inductance, and this gives you an idea of how much magnetic field it will
make when a current passes. You might think of an analogy with
capacitors the capacitance gives a measure of how much electric field
a certain charge will cause (since C-1 = V/Q and V is proportional to E).
Now, this magnetic field is important, because a changing current will
cause a changing magnetic field, and this will generate (or induce) a
voltage, and therefore upset the circuit it is in. This is something we
need to understand better but before we do so, let us remind ourselves
of the laws of electromagnetic induction:
6.2.5
Page 103
(24)
d
dt
B dS
S
d
dt
(25)
6.2.5.3 Equivalence?
The two expressions are very similar, as we shall see when we look at
the first from a different perspective. Look at the diagram we have
completed a loop by using a very long wire.
Region of magnetic
field B, pointing
down into paper
End wire moved in
direction of white
arrow. Distance
moved equals ut,
where u is speed.
Page 104
6.2.6
Inductors in circuits
dI
d d
LI L
dt
dt dt
(26)
Equation (26) is the definitive equation for inductors, just like Q=CV was
for capacitors, and V=IR is for a resistor.
Again, we need to bear in mind the comments above, that the voltage is
in the direction needed to oppose the change in current. You will often
see (26) with a minus sign in it for that reason.
Let us now calculate how much energy is stored in the device (actually in
its magnetic field)
U P dt VI dt L
dI
I dt L I dI
dt
1
2
LI 2
(27)
Since it seems sensible for the device to hold no energy when there is
no current and no field, we take the constant of integration to be zero.
6.2.7
Page 106
Force on right
hand wire
Direction of Bfield due to left
hand wire
R
6.2.7.1 A Classical Magnetic calculation
Equation (21) tells us that the magnetic field at a distance R from a
straight infinite wire is
B
0 I
2R
I2
F
IB 0
2R
L
(28)
Page 107
Distance
between
ion cores
From perspective of
ions in second wire,
first wire appears
negati vely charged.
Wires attract.
Distance
between
electrons
appears
contracted
Total field = Field due to ion cores + Field due to loose electrons
0
0
2 0 R 2 0 R
0
1
2 0 R
u2
2
2c
(29)
2 0 R
where the final stage has made use of a Binomial expansion of (-1) to
first order in u/c. Now the total charge of ion cores experiencing this field
per metre is of course 0. Therefore the total attraction of the ion cores
in the second wire to the first wire (electrons & ion cores) is
u2
Fions QE 2
2c
2
I
2
4c 0 R
20
2 0 R
Page 108
(30)
6.3.1
Circuit Analysis
Our aim here is to be able to solve a circuit like the one below. The
circles represent constant-voltage sources (a bit like cells or batteries)
and the linked circles represent constant-current sources. Our aim is to
find voltage difference across each component, and also to work out the
current in each resistor.
R1
V1
R2
R3
I1
R4
D
V2
E
In order to solve the circuit, we use two rules the Kirchoff Laws.
Kirchoffs 1st says that the total current going into a junction is equal to
the total current leaving it. Therefore, at B in the circuit below, we would
say that IBE = IAB + ICB, where IBE means the current flowing from B to E
(through R4).
Kirchoffs 2nd Law is that voltages always add up correctly. In other
words, no matter which route we took from E to B, say, we would agree
on the voltage difference between E and B. In symbols, if VBE means
Page 109
IBE = IAB+ICB;
I1 = ICB+ICD;
ICD = IDE
23
To see why the Law of Conservation of Energy is involved, let us suppose that our coulomb
of charge would lose 5J going from B to E via A, whereas it would lose 3J in going directly.
All it would have to do is go direct from B to E, then back to B via A and it would be back
where it started, having gained 2J of energy! This is not allowed.
Page 110
R1
R2
R3
R4
V1
E
This circuit is easier to analyse as it only has one supply. Supply V1
feeds a circuit with resistance
R1 + {R4 // (R2+R3)}
R1
R4 R2 R3
R4 R2 R3
Page 111
R1
V1
IL1
R2
R4
IL2
R3
I1
IL3
D
V2
E
E to A to B and back to E (loop 1),
E to B to C to E (loop 2), and
E to C to D to E (loop 3).
We call the current in loop 1 loop current number one (IL1), with IL2 and
IL3 representing the currents in the other two loops. We then express all
other currents in terms of the loop currents. Clearly, IAB = IL1, since R1 is
in the first loop alone. Similarly, IBC = IL2, and ICD = IL3.
The current through R4 is more complex, since this resistor is part of two
of the loops. We write IBE = IL1 IL2. Here IL1 is positive, since IBE is in
the same direction as IL1, whereas IL2 (which goes from E to B then on to
C) is in the opposite direction. These designations automatically take
care of Kirchoffs First Law. Notice that by this method, I1 = IL3 IL2.
Each loop now contributes one equation Kirchoffs 2nd law around that
loop. Clearly, if you go all the way round the loop, you must return to the
voltage you started with. Taking the first loop as an example, we have:
0
6.3.2
Alternating Current
CV0 cost
(31)
1
2
CV0 cost 12
24
It may help when writing the equations to notice the pattern: voltage sources count
positively if you go through them from to +, but negatively if you go from + to . The
voltages across resistors (e.g. VBA = I R1) count negatively if you go through them in the same
direction as the current, and positively if you go through them the opposite way to the current.
Page 113
(32)
1
C
(33)
Page 114
V=I XL
Origin
Origin
V=IR
Origin
I
V=I XC
Resistor
Inductor
Capacitor
25
In other words, a resistance is a special kind of impedance with zero phase difference, and
a reactance is a special kind of impedance when the phase difference is 90.
Page 115
6.3.3
2
Vrms
V 2 cos 2 t V02
0
R
R
2R
1
V0
Vrms
2
(34)
Resonance
One further circuit needs a mention, and that is the simple circuit of an
inductor and a capacitor connected together, as shown in the diagram
below. Both the voltage and current for the two components must be the
same, and so with the sign conventions chosen in the diagram:
Direction of positive current I
+Q
-Q
Q
d 2Q
dI
L
L 2
C
dt
dt
2
d Q
1
Q
2
LC
dt
(35)
6.4 Questions
1. Calculate the size of the repulsion force between two electrons 0.1nm
apart.
Page 116
Page 117
7.5 Questions
1. Calculate the wavelength and frequency of the quantum associated with a
60g ball travelling at 40m/s. Why dont we observe interference effects
with balls such as this?
2. Blue light has a wavelength of approximately 400nm, while red light has a
wavelength of approximately 650nm. Calculate the energies of photons of
blue and red light (a) in joules (b) in electron-volts (eV). One electron-volt
is equal to 1.6021019 J.
3. Work out the wavelengths of light emitted when electrons from the n=5, 4,
and 3 levels descend to the n=2 level. Why do you think that these
transitions were more important in the historical development of atomic
theory than the more fundamental transitions going down to the n=1
level?
4. Calculate the energies of the n=1, 2, 3, 4 and 10 levels for an ionized
helium atom (a helium nucleus with a single electron).
5. Calculate the size of a muonic hydrogen atom in comparison with a normal
hydrogen atom. A muonic hydrogen atom has a muon rather than an
electron moving near a proton. The muon has a charge equal to that of an
electron, but its mass is 207 times greater.
6. In this question, you will make an estimate for the size of a hydrogen atom.
Suppose that the atoms radius is r. Then the uncertainty in the electrons
position is 2r. Use the uncertainty principle to work out the uncertainty in
its momentum, and from this work out its typical kinetic energy, in terms of
r. The electrons typical electrostatic energy is e2/40r2. Find the value
of r which minimizes the total energy of the electron.
7. Calculate the energy liberated in the fusion reaction 21 H 31 H 42 He 01 n .
The masses of the particles are given in the table in unified mass units (u).
1u = 1.660431027 kg.
28
When analysing the collision, you find that you can not satisfy momentum and energy
conservation at the same time if only one photon is produced.
Page 124
7 Small Physics
The rules, or laws, of classical mechanics break down typically in three
cases. We have seen that when things start going quickly, we need to
take special relativity into account. Another form of relativity the
general theory is needed when things get very heavy, and the
gravitational fields are strong. The third exception is very mysterious
and occurs often when we deal with very small objects like atoms and
electrons. This is the realm of quantum physics, and many of its
discoverers expressed horror or puzzlement at its conclusions and
philosophy.
Having said that, there is no need to be frightened. While there is much
we do not understand, a set of principles have been set up which allow
us to perform accurate calculations. Furthermore, those calculations
agree with experiment to a high degree of accuracy. The development
of the transistor, hospital scanner, and many other useful devices testify
to this. The situation is analogous to a lion-tamer who can get the lion to
jump through a hoop, though she doesnt know what is going on inside
the lion.
Page 118
(1)
where h is the Planck constant, and has a value of 6.6310-34 Js. You
may also come across the constant h-bar h 2 , which can be used
in place of h if you wish to express your frequency as an angular
frequency in radians per second.
The momentum per particle (in the particle picture) is related to the
wavelength of the wave (in the wave picture) by the relationship
h
.
Wavelength (m)
(2)
7.2 Uncertainty
The bridge between wave and particle causes interesting conclusions.
We have seen in the chapter on Waves that a wave can have a welldefined frequency or duration (in time), but not both. This was
expressed in the bandwidth theorem:
f t 1 .
(3)
In other words, only something that lasts a long time can have a very
well known energy.
Let us have an example. Suppose a nucleus is unstable (radioactive),
with a half-life T. Seeing as the emission of the radiation is a process
that typically takes a time T, the energy of the alpha particle (or
whatever) has an inherent uncertainty of E h T . If we were watching
a spectrometer, monitoring the radiation emitted, we would expect to see
a spread of energies showing this level of uncertainty.
The bandwidth theorem also has something to say about wavelength:
1
x 1 .
Page 119
(4)
This is frequently stated as, You cant know both the momentum and
the position of a particle accurately. It might be better stated as, Since
it is a bit like a wave, it can not have both a well defined position and
momentum.
We can use this to make an estimate for the speed of an electron in an
atom. Atoms have a size of about 10-10 m. Therefore, for an electron in
an atom, x 10 10 m . So, using equation (4), p 10 23 kg m/s . Given
the electron mass of about 10-30 kg, this gives us a speed of about
107 m/s about a tenth the speed of light!
7.3 Atoms
Putting things classically for a moment, the electron orbits the nucleus.
While a quantum mechanic thinks this description very crude, we shall
use it as a starting point.
Now, lets imagine the electron as a wave. For the sake of visualization,
think of it as a transverse wave on a string that goes round the nucleus,
at a distance R from it. If the electron wave is to make sense, the string
must join up to form a complete circle. Therefore the circumference
must contain a whole number of wavelengths.
2R n
2R
nh
p
nh
pR
n
2
L n
(5)
2 I 2mR 2
,
Ze 2
Potential Energy
4 0 R
Kinetic Energy
(6)
(7)
where E is the total energy of the electron. We may use this information
to eliminate the radius in equations (6), obtaining:
E
Z 2e4m
24 0
1
.
n2
(8)
When dealing with atoms, the S.I. units can be frustrating. A more
convenient unit for atomic energies is the electron-volt. This is the
energy required to move an electron through a potential difference of
one volt, and as such it is equal to about 1.6010-19 J. In these units,
equation (8) can be re-written:
Z2
E 2 13.6 eV .
n
(9)
This form should be remembered. It will help you to gain a feel for the
energies an electron can have in an atom, and as a result, it will help you
spot errors more quickly.
26
The kinetic energy is calculated using the relationships derived in chapter 3. If you do not
wish to go in there, a simpler derivation can be employed. L=mvR, where v is the speed.
Therefore the kinetic energy mv2/2 = L2/(2mR2).
27
Hydrogen, that is, and hydrogen-like ions: which are atoms that have had all the electrons
removed apart from one.
Page 121
(10)
7.4.1
Types of radiation
Alpha decay: in which a helium nucleus (two protons and two neutrons)
is ejected from the unstable nucleus.
Beta decay: in which some weird nucleonic processes go on. In all beta
decays, the total number of nucleons (sometimes called the mass
number) remains constant.
In - decay (the most common), a neutron turns into a proton and an
electron. The electron is ejected at speed from the nucleus.
There are two other forms of beta radiation. In + decay, a proton turns
into a neutron and an anti-electron (or positron). The positron flies out of
the nucleus, and annihilates the nearest electron it sees.
The
annihilation process produces two gamma rays.
The other permutation is electron capture () in which an electron is
captured from an inner (low n) orbit, and reacts with a proton to make a
neutron. This phenomenon is detected when another electron descends
to fill the gap left by the captive and gives out an X-ray photon as it
does so.
Page 122
7.4.2
Radioactive decay
(11)
I N 0 e t I 0 e t
exp T
ln 12 T
ln 2 T
ln 2
T
(12)
7.4.3
Nuclear Reactions
Now for the final technique: You will need to be able to calculate the
energy released in a nuclear reaction. For this, add up the mass you
started with, and add up the mass at the end. Some mass will have
gone missing. Remembering that mass and energy are basically the
same thing the lost mass is the energy released from the nuclei.
Page 123
H
H
4
He
n
3
2.014102
3.016049
4.002604
1.008665
Page 125
8 Practical Physics
8.1 Errors, and how to make them29
Every dog has its day, every silver lining has its cloud, and every
measurement has its error.
If you doubt this, take (sorry borrow with permission) a school metre
stick, and try and measure the length of a corridor in your school. Try
and measure it to the nearest centimetre. Then measure it again.
Unless you cheated by choosing a short corridor, you should find that
the measurements are different. Whats gone wrong?
Nothing has gone wrong. No measurement is exact, and if you take a
series of readings, you will find that they cluster around the true value.
This spread of readings is called random error and will be determined
by the instrument you use and the observation technique. To be more
precise and polite, this kind of error is usually called uncertainty, as this
word doesnt imply any mistake or incompetence on the part of the
scientist.
So, whenever you write down a measurement, you should also write
down its uncertainty. This can be expressed in two ways absolute and
relative.
The absolute uncertainty gives the size of the spread of readings. You
might conclude that your corridor was (12.30.2)m long. In other words,
your measurements are usually within 20cm of 12.3m. In this case the
absolute uncertainty is 20cm.
The absolute uncertainty only gives part of the story. A 10cm error in the
length of a curtain track implies sloppy work. A 10cm error in the total
length of the M1 motorway is an impressive measurement. To make this
clearer, we often state errors (or uncertainties) in percentage form and
this is called relative uncertainty. The relative uncertainty in the length of
the corridor is
29
A mathematician would probably be appalled at some of the statements I make. The study
of errors and uncertainties is embedded in statistics, which is a well-established discipline.
There are many refinements to the results I quote which are needed to satisfy the rigour of a
professional statistician. However, the thing about uncertainties in measurements is that
quoting them to more than one significant figure is missing the point, and therefore our
methods only need to be accurate to this degree. If you are doing statistics and you want to
take things more seriously, then you will understand (2) from the addition of variances; and
you will realise that in 8.2.1 we really ought to be adding variances not errors. You will also
appreciate that (2) ought to have an (n-1) in the denominator to take into account the
difference between population and sample statistics, and that our section 8.2.2 is a form of
the Binomial theorem to first order.
Page 126
Relative Error
1.6% 2% .
Measuremen t
12.3 m
(1)
x
n
(2)
Therefore, the more measurements you take, the more accurate the
work. Notice that if you wish to halve the uncertainty, you need to take
four times as many readings. This is subject to one proviso:
Measurements also have a resolution.
This is the smallest
distinguishable difference that the measuring device (including the
technique) can detect. For a simple length measurement with a metre
ruler, the resolution is probably 1mm. However if, by years of practice
with a magnifying lens, you could divide millimetres into tenths by eye,
you would have a resolution of 0.1mm using the same metre stick. That
is why we say that the resolution depends on the technique as well as on
the apparatus.
The uncertainty of a measurement can never be less than the resolution.
This is the proviso we mentioned below equation (2). Why should this
be the case? Let us have a parable.
Many years ago, the great nation of China had an emperor. The masses
of the population were not permitted to see him. One day, a citizen had
30
This result will be proved in any statistics textbook. To give a brief justification the more
readings you take, the more likely you are to have some high readings cancelling out some
low readings when you take the average.
Page 127
8.2.1
31
Of course, there is a good chance that the errors will partly cancel out, and so our method
of estimating the overall error is pessimistic. Nevertheless, this kind of error analysis is good
enough for most experiments after all it is better to overestimate your errors. If you want to
do more careful analysis, then you work on the principle that if the absolute uncertainties in a
Page 128
8.2.2
set of measurements are A, B, C, then the absolute uncertainty in the sum (or in any of the
differences) is given by A B C . This result comes from statistics, where we find
that the variance (the square of the standard deviation) of a sum is equal to the sum of the
variances of the two measurements.
2
32
The conclusions of this paragraph can be justified using calculus. If measurement x has
y
y
dy x
y x
x
x
An x n 1
n
n ,
y
x y
x
dx y
Page 129
.
g ( x, y ) A h ( x , y ) B
We can then plot g(x,y) against h(x,y), and obtain the parameters A and
B from the gradient and intercept of the line. Furthermore, the presence
of the straight line on the graph assures us that our function f was a
good guess. We shall now look at the most common examples.
8.4.1
So we plot (ln y) on the vertical axis, and (x) on the horizontal. The yintercept gives ln A, and the gradient gives B.
8.4.2
Page 130
8.4.3
Power laws
y Ax B .
The
y Ax B
ln y ln A ln x B .
ln y ln A B ln x
Here we plot (ln y) against (ln x), and find the power (B) as the gradient
of the line. The A value can be inferred from the y-intercept, which is
equal to ln A.
8.4.4
Other forms
This looks even worse, doesnt it? But remember that it is x and y that
are known. If we plot ( y x ) on the vertical, and (x5/2) on the horizontal,
a straight line appears, and we can read A and B from the y-intercept
and gradient respectively.
8.5 Questions
1. Work out the relative uncertainty when a 5V battery is measured to the
nearest 0.2V.
2. If I dont want to have to correct my watch more than once a week, and I
never want my watch to be more than 1s from the correct time, calculate
the necessary maximum relative uncertainty of the electronic oscillator
which I can tolerate.
3. My two-storey house is 7.050.02m tall. The ground floor is 3.20.01m
tall. How tall is the first floor?
4. I want to measure the resistance of a resistor. My voltmeter can read up
to 5V, with an absolute uncertainty of 0.1V. My ammeter can read up to
1A with an absolute uncertainty of 0.02A. Assuming that my resistor is
approximately 10, calculate the absolute uncertainty of the resistance I
measure using the formula R=V/I. Assume that I choose the current to
make the relative uncertainty as small as possible.
Page 131
9 Appendix
9.1 Multiplying Vectors
Physics is riddled with quantities which have both magnitude and
direction velocity, acceleration, displacement, force, momentum,
angular velocity, torque, and electric field to name but eight. When
describing these, it is very useful to use vector notation. At best this
saves us writing out separate equations for each of the components.
While the addition and subtraction of vectors is reasonably
straightforward (you add, or subtract, the components to get the
components of the result), multiplication is more tricky.
You can think of a vector as a little arrow. You can add them by stacking
them nose-to-tail, or subtract them by stacking them nose-to-nose. But
how do you go about multiplying them? It is not obvious!
To cut a long story short, you cant do it unambiguously. However there
are two vector operators which involve multiplication and are useful in
physics. Ordinary multiplication is commonly written with either the
cross () or dot (), so when it comes to vectors we call our two different
multiplication processes the dot product and cross product to
distinguish them.
These are the closest we get to performing
multiplication with vectors.
9.1.1
(1)
Notice that the dot product of two vectors is itself a scalar. Note that
when we talk about the square of a vector, we mean its scalar product
with itself. Since in this case, =0, this is the same as the square of the
vectors magnitude.
Page 132
9.1.2
(2)
(3)
j k
a b c
u v w
Page 134
and this agrees with the units of the left hand side. Now this may all
seem pretty obvious, but it gives us a useful procedure for checking
whether our working is along the right lines. If, during your calculations,
you find yourself adding a charge of 3C to a distance of 6m to get a
result of 9N; or you multiply a speed of 13m/s by a time of 40s and get a
current of 520A; then in either case you must have made a mistake!
We can also use the principle that units must balance to guess the form
of an equation we do not know how to derive. For example, you may
guess that the time period of a simple pendulum might depend on the
length of the pendulum L, the strength of the local gravity g and the
mass of the pendulum bob m. Now
L is measured in m
g is measured in N/kg or m/s2
m is measured in kg,
and we want a time period, which will be measured in s.
V C C
s.
A V A
Yes, you get seconds. Therefore, it should come as little surprise to you
that if you double the resistance of a capacitor-resistor network, it will
take twice as long to charge or discharge. Furthermore, you have
worked this out without recourse to calculus or the tedious electrical
details of section 6.1.2.2.
Page 135
33
By independent we mean that no one unit can be derived entirely from a combination of the
others. For example, m, s, kg and J would be no good as a set of four since J can already be
expressed in terms of the others J = kg (m/s)2, and hence we have ambiguity arising as to
how we express quantities.
34
OK, if you want to do work where there are electric currents and temperatures as well as
mechanical quantities, you might need to go up to five (an extra one for temperature).
35
Indeed, quantities with no units are usually said to have the dimensions of the number
one. Thus [angle] = 1. It follows that angular velocity has dimensions of [angle][time] = 1T
= T1. This comes from the property 1 has in being the unity operator for multiplication.
Page 136