0% found this document useful (0 votes)
192 views170 pages

Physics Eduqas Revision Guide: Exeter Mathematics School

This document is a revision guide for Pre-U Physics covering various topics in mechanics. It provides definitions and formulas for key concepts like scalars and vectors, kinematics, Newton's laws of motion, and forces. For each topic, it lists the important things candidates should be able to do, such as resolve vectors, draw free body diagrams, calculate moments of forces, construct graphs of motion, and use equations of motion. It also explains concepts like friction, momentum, impulse, and collisions. The revision guide is intended to concisely summarize the essential information needed to study for the Pre-U Physics assessment.

Uploaded by

Joe Rowing
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
192 views170 pages

Physics Eduqas Revision Guide: Exeter Mathematics School

This document is a revision guide for Pre-U Physics covering various topics in mechanics. It provides definitions and formulas for key concepts like scalars and vectors, kinematics, Newton's laws of motion, and forces. For each topic, it lists the important things candidates should be able to do, such as resolve vectors, draw free body diagrams, calculate moments of forces, construct graphs of motion, and use equations of motion. It also explains concepts like friction, momentum, impulse, and collisions. The revision guide is intended to concisely summarize the essential information needed to study for the Pre-U Physics assessment.

Uploaded by

Joe Rowing
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 170

Physics Eduqas Revision

Guide

Exeter Mathematics School

May 18, 2018


v1.0
Contents

Contents i

About this revision guide iii

Structure of Assessment v

Part A 3

1 Mechanics 3

2 Gravitational Fields 9

3 Deformation of Solids 13

4 Energy Concepts 19

5 Electricity 25

6 Waves 31

7 Superposition 41

8 Atomic and Nuclear Processes 57

9 Quantum Ideas 63

i
ii CONTENTS

Part B 73

10 Rotational Mechanics 73

11 Oscillations 77

12 Electric Fields 81

13 Gravitation 87

14 Electromagnetism 99

15 Special Relativity 105

16 Molecular Kinetic Theory 111

17 Nuclear Physics 123

18 The Quantum Atom 131

19 Interpreting Quantum Theory 137

20 Astronomy and cosmology 145

Appendices 157

A Equations 157
About this revision guide

This is a revision guide for Pre-U Physics 2016-18. The specification is avail-
able at http://www.cie.org.uk/images/163265-2016-2018-syllabus.pdf.

This revision guide has been written by the Physics Department at Westmin-
ster School. This is a narrow revision guide, if you are looking for a textbook
try Openstax College Physics or Advanced Physics by Adams and Allday.

The revision guide is written in LATEX, is open source and hosted on Github
(https://github.com/mrpsharp/physics-PreU). Most of the work was done
using SageMathCloud, an excellent online development environment for LATEX,
Sage, Jupyter notebooks and more.

If you find a mistake, or would like to suggest an enhancement, please do


submit bug reports and enhancement requests. Simply click the link and go
to ‘issues’ in order to report your thoughts. If you would like to contribute
to the revision guide, and this would be great, please ensure you read ‘CON-
TRIBUTING.md’ and then you can submit a pull request.

iii
Structure of Assessment

Components Weighting
Paper 1 Multiple Choice 1 hour 30 minutes 20%
Candidates answer 40 multiple-choice questions based on
Parts A and B of the syllabus content.
40 marks
Paper 2 Written Paper 2 hours 30%
Section 1: Candidates answer structured questions based on
Part A of the syllabus content.
Section 2: Candidates answer structured questions related
to pre-released material.
100 marks
Paper 3 Written Paper 3 hours 35%
Section 1: Candidates answer structured questions requiring
short answers or calculations and some longer answers. The
questions are focused on Part B of the syllabus content, but
may also draw on Part A.
Section 2: Candidates answer three questions from a choice
of six. Three questions will have a strong mathematical
focus and three questions will focus on philosophical issues
and/or physics concepts. Learning outcomes marked with
an asterisk (*) will only be assessed in this section.
140 marks
Practical Investigation 20 hours 15%

v
Part A

1
1 Mechanics

Content

• scalars and vectors

• moment of a force

• kinematics

• Newton’s laws of motion

• conservation of linear momentum

• density

• pressure

Candidates should be able to:

Scalars and Vectors

(a) distinguish between scalar and vector quantities and give examples of each
A scalar quantity1 is one which has only a magnitude whereas a vector has
both magnitude and direction. We often use positive and negative values to
indicate direction (e.g. v = −2 ms−1 ) but this does not mean that all negative
values are vectors!

Note that there are different ways of multiplying vectors and scalars. Two
vectors can be multiplied to give a scalar or a vector. For example, work done
is the (scalar) product of force and displacement, both vectors.

1
strictly we are modelling a physical quantity as a mathematical object

3
4 CHAPTER 1. MECHANICS

(b) resolve a vector into two components at right angles to each other by draw-
ing and by calculation

Vectors can be split into two components using trigonometry. The diagram
below shows a velocity vector being split into horizontal and vertical compo-
nents vx and vy .

v = vx + vy
v
vy vx = v cos θ
vy = v sin θ
θ
vx

(c) combine any number of coplanar vectors at any angle to each other by
drawing

Vectors can be added by placing them end to end. The resultant vector is
the one joining the start of the first vector to the end of the final vector. Its
magnitude and direction can be calculated by trigonometry or scale drawing.

v
v3
v1
v2

v = v1 + v2 + v3

Static Equillibrium

(d) calculate the moment of a force and use the conditions for equilibrium to
solve problems (restricted to coplanar forces)

Kinematics
FORCES 5

(e) construct displacement-time and velocity-time graphs for uniformly accel-


erated motion

(f) identify and use the physical quantities derived from the gradients of displacement–
time and areas and gradients of velocity–time graphs, including cases of non-
uniform acceleration

The quantities are given in the table below:

gradient area
displacement-time velocity –
velocity-time acceleration displacement

If the graph is non-linear then the gradient of a tangent must be taken. Note
that areas below the axis in a velocity-time graph represent negative displace-
ment.

(g) recall and use:


∆x
v=
∆t
∆v
a=
∆t

(h) recognise and use the kinematic equations for motion in one dimension
with constant acceleration:

(i) recognise and make use of the independence of vertical and horizontal mo-
tion of a projectile moving freely under gravity

Forces

(j) recognise that internal forces on a collection of objects sum to zero vecto-
rially

This is as a result of Newton’s Third Law.

(k) recall and interpret statements of Newton’s laws of motion

1. An object will remain at rest, or continue at a constant velocity, unless


a resultant force acts upon it.
2. F = ma, where F is the vector sum of the forces acting on the body.
Or, alternatively F = dp
dt (see below).
6 CHAPTER 1. MECHANICS

3.

(l) recall and use F = ma in situations where mass is constant

Remember that F is the resultant force acting on the body.

(m) understand the effect of kinetic friction and static friction

(n) use Fk = µk N and Fs = µs N , where N is the normal contact force and µk


and µs are the coefficients of kinetic friction and static friction, respectively

Friction occurs between two objects when they are pushed together by a nor-
mal force. A useful model is that the maximum size of the frictional force is
proportional to the normal force. There is usually a difference between the
constant of proportionality when the two surfaces are stationary compared to
each other (static friction) compared to when they are sliding past each other
(kinetic friction). It is usually the case that µk < µs .

An interesting result is that blocks of different masses should take the same
distance to slide to a halt:

1 2 1 2 1 2
2 mv 2 mv 2 mv v2
s= = = =
F µk N µk mg 2µk g

(o) recall and use the independent effects of perpendicular components of a


force

As with velocities, forces can be split into two perpendicular components and
their effects considered independently.

(p) recall and use p = mv and apply the principle of conservation of linear
momentum to problems in one dimension

(q) distinguish between elastic and inelastic collisions

(r) relate resultant force to rate of change of momentum in situations where


mass is constant and recall and use F = ∆P ∆t

(s) recall and use the relationship impulse = change in momentum

Multiplying both sides of the equation above by time gives:

F ∆t = ∆P

The quantity on the left hand side is the impulse.


FORCES 7

(t) recall and use the fact that the area under a force-time graph is equal to
the impulse

Using calculus to solve differential version of Newton’s Second Law above


gives: Z t1
∆P = F dt
t0

The right-hand side of this equation represents the area under a force-time
graph.

(u) apply the principle of conservation of linear momentum to problems in two


dimensions

(v) recall and use density = mass / volume

(w) recall and use pressure = normal force / area

(x) recall and use p = ρgh for pressure due to a liquid. These are GCSE
equations and should present no problems.
2 Gravitational Fields

(a) recall and use the fact that the gravitational field strength g is equal to the
force per unit mass and hence that weight W = mg

A field is a region where a particle experiences a force. If this is applied to


gravitation, then we can say that a gravitational field is a region where a
mass experiences a force.

You can only tell if a field exists when it exerts a force on something. It is a
way of envisaging (seeing in your mind’s eye) the size and the direction of the
force that would be exerted on a particle when placed in that field.

A gravitational field is produced by anything with mass.

Therefore, a gravitational field is a way of envisaging what would happen to


a mass if it were placed in the field due to another mass.

The field is usually represented by lines which show both the direction and
strength of the field.

The strength of a gravitational field (the field strength) at any point is the
force felt per unit mass at that point. This is a definition.

It can be written as a word equation:

Gravitational field strength at a point (N/kg)= Force felt by mass (measured


in Newtons)/Size of mass (measured in kilograms)

Or in symbols:

F
g=
m

The force, F, felt by any object on the surface of the Earth due to the grav-
itational field strength of the Earth is known as its weight. It is given the

9
10 CHAPTER 2. GRAVITATIONAL FIELDS

symbol W.

This means that we can re-write equation above for the field strength at the
surface of the Earth by putting W instead of F.

W
g=
m

This then rearranges to an equation that you have all seen before:

W = mg

Thus the weight of an object on the surface of the Earth is its mass multiplied
by the gravitational field strength g.

(b) recall that the weight of a body appears to act from its centre of gravity

The centre of gravity of an object is the point where the weight acts or appears
to act.

Thus, when you draw a free-body force diagram for any object in a gravita-
tional field, you draw one arrow from the centre of gravity of the object to
represent the force due to the field. On the Earth this is, of course, the weight
and the arrow points vertically downwards.

(c) sketch the field lines for a uniform gravitational field (such as near the
surface of the Earth)

A uniform field is a field where the field strength is the same at all points in
the field.

This means that for a gravitational field the force felt per unit mass (see
definition) is the same at all points.

The surface of the Earth is a very good approximation to a uniform field.

Therefore if you draw a diagram of the Earth’s gravitational field at the Earth’s
surface over a small area, it will look like Figure 2.1

As you can see, the field lines are parallel and evenly-spaced. This is always
the case for a uniform field.
11

(d) explain the distinction between gravitational field strength and force and
explain the concept that a field has independent properties.

There is a very important distinction to make between gravitational field


strength and force at this point: The field strength at any point is the same
for all bodies in the field and is the force felt per kilogram, but the force is
different and depends on the size of the mass there.

This is best illustrated with an example: If a mass of 60kg is in the Earth’s


gravitational field at the surface of the Earth, then we can calculate the force
acting on it, its weight, using equation (3):

W = mg = 60 kg × 9.8 N kg−1 = 590 N

So the force felt by the 60kg mass is 590N but the field strength for the mass
and for any other mass is 9.8Nkg-1 . So the field strength is fixed by your
position in the field and the size of the mass that is exerting the field, and
nothing else. The force depends on the mass in the field as well.
3 Deformation of Solids

Content

• elastic and plastic behaviour

• stress and strain

Candidates should be able to:

(a) distinguish between elastic and plastic deformation of a material

Elastic deformation is defined as deformation where the sample returns to


its original length when the load is removed. Plastic deformation involved a
permanent change in length of the sample.

(b) recall the terms brittle, ductile, hard, malleable, stiff, strong and tough,
explain their meaning and give examples of materials exhibiting such behaviour

Brittle Brittleness is an indicator of how soon after the yield point a material
fractures. Failure will be through the propagation of cracks. A brittle
material cannot absorb much energy before breaking. For example, glass
and ceramics can be strong but brittle.

Ductile Ductility is a measure of plastic behaviour under tension. It gives


an indication of how easily a material can be drawn into wires i.e. can
withstand large strains without breaking. Copper is highly ductile.

Hard Hardness is a measure of a materials ability to resist impact or scratch-


ing. Diamond is an exceptionally hard material.

Malleable Malleability is a measure of plastic behaviour under compression.


It gives an indication of how easily a material can be worked. Met-

13
14 CHAPTER 3. DEFORMATION OF SOLIDS

als are ductile and hence relatively easy to form into shapes for use in
manufacture.

Stiff Stiffness measures how a material much resists deformation. A measure


of stiffness is the Young Modulus. Glass fibres and steel are both stiff
materials.

Strong A strong material is able to withstand a large stress without failing.

Tough Toughness is a measure of the ability of a material to resist fail-


ure through crack propagation. It is the opposite of brittleness. A
tough material is able to absorb a lot of energy without breaking. Plas-
tics/polymers are often tough.

(c) explain the meaning of, use and calculate tensile/compressive stress, ten-
sile/compressive strain, spring constant, strength, breaking stress, stiffness and
Young modulus

Firstly, compressive forces and deformations are those which reduce the length
of the sample whereas tensile forces act to increase its length.

Stress Stress is defined as the force per unit of cross-sectional area applied
to a material.
F
σ=
A
Stress is measured in pascals (Pa).

Strain Strain is the fractional extension of a material.


x
=
l
where l is the original length.

Spring constant The spring constant k is the force per unit of extension of
a material during its proportional phase of deformation. It is defined by
Hooke’s Law:
F = kx
The spring constant is often used as a measure of stiffness of an object.

Strength Strength is often measured as the maximum stress a material can


withstand before permanent deformation. This is known as the yield
stress.

Breaking stress This is the stress at which the material fails.


15

Young modulus This is a quantitative measure of the stiffness of a mate-


rial, defined as stress per unit of strain in the proportional region the
material’s behaviour.
σ
E=


(d) draw force-extension, force-compression and tensile/compressive stress-


strain graphs, and explain the meaning of the limit of proportionality, elastic
limit, yield point, breaking force and breaking stress

The gradient of a force-extension graph gives the spring constant.

σ
C
D A The limit of
proportionality
B B The yield point
A
C The ultimate stress
(maximium
stress)

D The breaking point

Figure 3.1: Stress-strain curve for a ductile material

Figure 3.1 shows an example stress-strain curve. Note that the limit of pro-
portionality is often a good approximation of the elastic limit of a metal. The
“breaking stress” usually refers to the ultimate stress, i.e. the maximum stress
the material can withstand, rather than the stress at the breaking point.

(e) state Hooke’s law and identify situations in which it is obeyed

F = kx

Hooke’s law is obeyed by an ideal spring and by a sample of metal up to the


limit of proportionality.
16 CHAPTER 3. DEFORMATION OF SOLIDS

(f) account for the stress-strain graphs of metals and polymers in terms of the
microstructure of the material.

Metals

Metals consist of positive ions in a sea of delocalised electrons. During the


elastic phase of deformation the spaces between the ions get larger and smaller.
The metallic bonds resist this change from their equilibrium length and act
like small springs acting to return the spacing to its original length.

An initial expectation of the plastic phase of deformation in a metallic lattice


may be that the planes of ions slip past one another; however an analysis
of the forces required for such movement gives an answer hundreds of times
higher than the measured yield stress. Instead, the plastic deformation of
metals must be explained in terms of dislocations. A dislocation occurs when
there is a gap in the metallic lattice. Dislocations occur naturally in materials
and enable plastic deformation to occur through the breaking of individual
bonds in succession, rather than all at once.

Figure 3.2: The movement of a dislocation

Figure 3.2 shows a dislocation moving within a metal which would allow the
metal to deform by moving one atom at a time. As the movement of disloca-
tions is the dominant mode of plastic deformation, changes to the ability of
dislocations to move through the metal have significant effects on it properties.
For example:

Work Hardening As a metal is deformed, the dislocations move through the


structure. Slowly the dislocations reach grain-boundaries or other dis-
17

locations and are no longer able to move. The metal therefore becomes
less ductile and more brittle. This may be a desired property in order
to harden a metal, or the additional brittleness may be undesirable.

Alloying The addition of alloying atoms to the lattice can ‘pin’ a dislocation
in place (as shown in figure 3.3. The metal is therefore no longer able
to deform by the movement of dislocations so the metal has a greater
yield stress and is less ductile. Examples include adding carbon to iron
to produce steel or adding zinc to copper to produce brass.

Figure 3.3: Alloying atom pinning a dislocation

The effect on a stress-strain graph can be seen in figure 3.4 below.

σ
Alloy

Pure metal

Figure 3.4: Stress-strain curve for a ductile material


18 CHAPTER 3. DEFORMATION OF SOLIDS

Polymers

Polymers consist of long chain molecules weakly held together by intermolec-


ular forces. Initially the molecules are likely to be tangled-up together. As
force is applied it is initially difficult to move the polymer chains from this
state (A). As the chains begin to unravel they straighten out by bond rota-
tion, requiring relatively little force for a large increase in strain (B). As the
polymer chains become straight it becomes much more difficult to extend the
material any further without damaging the material (C).

A B C

Figure 3.5: Stress-strain curve for a polymer


4 Energy Concepts

Content

• work

• power

• potential and kinetic energy

• energy conversion and conservation

• specific latent heat

• specific heat capacity

Candidates should be able to

(a) understand and use the concept of work in terms of the product of a force
and a displacement in the direction of that force, including situations where
the force is not along the line of motion

(b) calculate the work done in situations where the force is a function of dis-
placement using the area under a force-displacement graph

Equation ?? applies whenever a constant force acts over a displacement; how-


ever, if the force varies then a different approach is needed. Force a constant
force acting in the same direction as the displacement, it can be seen that the
area under a force-displacement graph is equal to F s, i.e. the work done. This
is generally true and work can be written in an integral form as:
Z
W = F ds (4.1)

19
20 CHAPTER 4. ENERGY CONCEPTS

Area = W

Figure 4.1: Work as area under a graph

(c) understand that a heat engine is a device that is supplied with thermal
energy and converts some of this energy into useful work

A heat engine is a device which uses heat to do work. This is shown schemat-
ically in figure 4.2. The energy for the work done comes from the difference
between Q1 and Q2 . Examples of heat engines include internal combustion
engines, jet engines and steam turbines.

Q1

Q2

Figure 4.2: A heat engine

(d) calculate power from the rate at which work is done or energy is transferred

(e) recall and use P = F v

For a constant force, this equation can be shown from equation ?? and ??:
W s
P = = F = Fv
t t
21

(f) recall and use ∆E = mg∆h for the gravitational potential energy trans-
ferred near the Earth’s surface

This is familiar from GCSE.

(g) recall and use g∆h as change in gravitational potential

Gravitational potential is defined as the energy per unit mass. Hence, the
change in gravitational potential is given by

mg∆h
= g∆h
m

(h) recall and use E = 21 F x for the elastic strain energy in a deformed material
sample obeying Hooke’s law

(i) use the area under a force-extension graph to determine elastic strain en-
ergy

This relies on equating the work done straining an object with the elastic
strain energy stored in the object. Once this is done, the statement follows
from equation 4.1 and figure 4.1.

The area of such a graph when the material obeys Hooke’s Law is 21 F x.

Area = 12 F x
x

Figure 4.3: Work as area under a graph

(j) derive, recall and use E = 12 kx2

This can be arrived at from Hooke’s Law (F = kx) and the definition of
work in equation 4.1, noticing that the extension of the spring is equal to the
displacement of the object.
1
Z Z x
W = F ds = kx dx = kx2
0 2
22 CHAPTER 4. ENERGY CONCEPTS

This integration could equally be done by substituting F = kx into the ex-


pression for elastic strain energy derived above from the graph.

(k) derive, recall and use E = 12 mv 2 for the kinetic energy of a body

Consider the work done accelerating an object from rest to a velocity v. Using
the equations for uniform acceleration with u = 0 we can see that a = vt and
s = v2 · t so:

v v 1
W = F s = mas = m t = mv 2
t 2 2

Since this work has gone into the kinetic energy of the object this formula
gives us this kinetic energy.

(l) apply the principle of conservation of energy to solve problems

The principle of conservation of energy states that energy cannot be created


or destroyed, only transferred between different forms.

(m) recall and use

useful energy (or power) out


% efficiency = × 100
total energy (or power) in

This is familiar from GCSE.

(n) recognise and use ∆E = mc∆θ, where c is the specific heat capacity

The specific heat capacity is defined as the energy required to heat 1 kg of a


substance by 1 ◦C.
23

Example Question
A kettle with a power rating of 2 kW heats 500 g of water from 15 ◦C
to boiling. If the kettle is 80% efficient, calculate the time take for the
water to boil.
The specific heat capacity of water is 4200 J kg−1 ◦C−1
Answer
Total heat energy required by the water:

∆E = 0.5 kg × 4200 J kg−1 ◦C−1 × 85 ◦C = 178.5 kJ

Useful power provided by the kettle:

P = 0.8 × 2 kW = 1.6 kW
Therefore the time taken is:
∆E 178.5 kJ
t= = = 112 s
P 1.6 kW

(o) recognise and use ∆E = mL, where L is the specific latent heat of fusion
or of vaporisation

When a substance changes state it releases or absorbs energy. This energy is


known as the latent heat. The specific latent heat is the energy absorbed or
released when 1 kg of the substance changes state.
5 Electricity

(a) discuss electrical phenomena in terms of electric charge

(b) describe electric current as the rate of flow of charge and recall and use
I = ∆Q/∆t

Electric current is the flow of charge. The current is defined as the rate of
flow of charge. Conventional current flows from positive to negative. This
current can consist of positive charges flowing from positive to negative or,
more usually, negative charges flowing from negative to positive. The total
current depends on the charge carrier density, the cross-sectional area, the
charge on the carrier and the drift velocity of the carriers.

(c) understand potential difference in terms of energy transfer and recall and
use VQ = W

When a charge moves through an electric field it gains of loses potential energy.
The energy change per unit charge is defined as the potential difference.

(d) recall and use the fact that resistance is defined by R = V/I and use this to
calculate resistance variation for a variety of voltage-current characteristics

As well as measuring resistance directly it can be found from a graph of V


against I. Note: the resistance is defined as V/I at all points and is not equal
to the gradient of a V/I graph except in the case that V is proportional to I

(e) define and use the concepts of emf and internal resistance and distinguish
between emf and terminal potential difference

(f) derive, recall and use E = I(R + r ) and E = V + Ir

A real cell or battery can be represented by a cell circuit symbol in series with
a resistor. This resistance represents the internal resistance of the cell and the
fixed, theoretical potential difference across the cell symbol is the emf of the
cell (the electromotive force provided). When a voltmeter is connected across
the cell the terminal potential difference is measured.

25
26 CHAPTER 5. ELECTRICITY

E
r

Figure 5.1: Terminal potential difference

In order to relate the terminal potential difference to the internal character-


istics of the cell we must subtract the potential difference across the internal
resistance from the emf. In symbols this gives:

V = E − Ir

which can be re-arranged to give the formula above.

If our real cell is connected into a circuit with a load resistance R, the circuit
in figure 5.2 is produced.

E
r

Figure 5.2: A loaded real cell

Now, the terminal potential difference must be equal to the potential difference
across the load resistor, R.

IR = E = Ir

Which can be re-arranged to give the second equation in the specification.

(g) recall and use P = VI and W = VIt, and derive and use P = I2 R

Power is defined as the energy transferred per unit of time. In the case of
electrical power this is the product of current (charge per unit time) and
potential difference (energy per unit charge). Given a constant voltage and
current, the energy transferred (work done) is given by P = Wt = IV t
27

(h) recall and use R = ρl/A

This formula allows the calculation of a regular sample of material. In this


formula ρ is the resistivity, l is the length of the sample and A its cross-
sectional area. Typical resistivities for conductors are of the order of 10−8 Ω m
and above 109 Ω m for insulators. Semi-conductors lie between these values.

(i) recall the formula for the combined resistance of two or more resistors in
series and use it to solve problems RT = R1 + R2 + . . .

(j) recall the formula for the combined resistance of two or more resistors in
parallel and use it to solve problems R1T = R11 + R12 + . . .

This are fairly simple to derive from Kirchoff’s Laws (see below).

(k) recall Kirchhoff’s first and second laws and apply them to circuits con-
taining no more than two supply components and no more than two linked
loops

Kirchoff’s First Law The current that flows into any junction is equal to
the current which flows out.

Kirchoff’s Second Law The sum of the emfs around any closed loop of
a circuit must equal the sum of the potential differences across any
components. It is important to note that direction matters and if the
loop crosses emfs or components against the flow of current they must
be subtracted.

These two laws, and the definition of resistance, are the most useful tools
in circuit analysis. The important skill is to work methodically through the
circuit applying the laws, rather than attempting to solve the circuit all in
one go.
28 CHAPTER 5. ELECTRICITY

Example Question
Calculate the current through the 3 V cell.

6.0 V 3.0 V
Ix Iy
10.0 Ω

2.0 Ω 0.5 Ω

Answer
We can use Kirchoff’s Second Law to derive two expressions linking
Ix and Iy . The first is formed by creating a loop consisting of the two
branches containing cells.

6 − 3 = 2Ix − 0.5Iy

Note that I am using a clockwise loop so the signs of the two compo-
nents in the ‘Y’ branch have negative signs.
The second expression is now arrived at by using the outermost loop of
the circuit (and using Kirchoff’s First Law to get the current through
the 10 Ω resistor):
6 = 10(Ix + Iy ) + 2Ix
These two equations can now be used to solve for Iy , giving

Iy = −0.923 A

Note the sign of Iy is negative, this means that current is flowing in


the opposite direction to the arrow shown. This means that the cell
is charging.

(l) appreciate that Kirchhoff’s first and second laws are a consequence of the
conservation of charge and energy, respectively

The charge flowing into or out of a junction in a given time, t, is given by


Q = It. Given that a junction can neither store nor create charge, Kirchoff’s
First Law follows directly.

Each charge carrier can only take one loop around the circuit. Once it returns
to its original position its energy must be equal to the amount it had when
it left. The charge carrier gains energy passing through cells and loses it
passing through components. Since the sum of these energies must be zero
29

and W = qV , the sum of emfs must equal the sum of potential differences
across components.

(m) use the idea of the potential divider to calculate potential differences and
resistances

When two resistors are in series with a battery we say that the circuit is a
potential divider.

R1

R2

Figure 5.3: A potential divider

Since there are no junctions in the series circuit we can know that the current
is the same in all parts of the circuit and that the total resistance is R1 + R2 .
The p.d. across resistor 1 is therefore given by:
V R1
V1 = IR1 = R1 = V
R1 + R2 R1 + R2
In other words, the ratio of p.d.s in the circuit is equal to the ratios of the
resistances.

This can also be extended to the ratios between the components as they share
the same current.

V1 V2 R1 V1
I1 = I2 =⇒ = =⇒ =
R1 R2 R2 V2
30 CHAPTER 5. ELECTRICITY

Example Question
Calculate the resistance of the bulb in the circuit below.

9V

800 Ω 3V

Answer
The potential difference across the bulb must be 6 V by Kirchoff’s
Second Law. Therefore:

3V 800 Ω
=
6V R
6V
=⇒ R = 800 Ω × = 1600 Ω
3V
6 Waves

Content

• progressive waves

• longitudinal and transverse waves

• electromagnetic spectrum

• polarisation

• refraction

Candidates should be able to:

(a) understand and use the terms displacement, amplitude, intensity, fre-
quency, period, speed and wavelength

All waves consist of oscillations. The oscillations could be of particles, for


example in a sound wave, or of an electromagnetic field, as in a light wave.

The following terms are used to describe properties of waves:

• displacement: This is a measurement of the distance and direction


away from the equilibrium position.

• amplitude: The maximum displacement of the oscillation, represented


by A.

• intensity: The power of the wave per unit area, represented by I. The
unit of intensity is W m−2 .

• frequency: The number of oscillations per second, represented by f .

31
32 CHAPTER 6. WAVES

• period: The time taken for one oscillation, represented by T .

• speed: The speed of a wave represented by v. This will depend on the


medium through which the wave is travelling.

• wavelength: The distance over which a wave’s shape repeats, repre-


sented by λ.

(b) recall and apply f = 1


T to a variety of situations not limited to waves

This equation follows from the definition of the frequency and time period of
a wave. Remember to use Hertz as the unit for frequency and seconds as the
unit for period.

(c) recall and use the wave equation v = f λ

distancetravelled
speed =
timetaken

For a wave, the distance travelled in one time period, T , is the wavelength, λ.
Therefore we can write

λ
v=
T

Then, using the equation f = T,


1
we can write:

v = fλ

This is known as the wave equation and can be applied to all waves. The
frequency of the wave generally depends on the source of the wave or how it
is produced and the speed depends on the medium through which the wave is
travelling.

(d) recall that a sound wave is a longitudinal wave which can be described in
terms of the displacement of molecules or changes in pressure

When a sound wave travels through a material, the collisions of molecules


are parallel to the direction of travel. Energy is transferred through these
collisions and the speed of the sound wave will depend on factors such as the
density of the material and the temperature.
33

When a sound wave is viewed on an oscilloscope, it looks as though the oscilla-


tions are perpendicular to the direction of travel, as in a transverse wave. The
y-axis can represent either the displacement of molecules (still in the parallel
direction) from their equilibrium position, or the difference in pressure.

figs/chapt-6/soundwave.JPG

Figure 6.1: Sound wave in air (credit: hyperphysics)

(e) recall that light waves are transverse electromagnetic waves, and that all
electromagnetic waves travel at the same speed in a vacuum

(f) recall the major divisions of the electromagnetic spectrum in order of wave-
length, and the range of wavelengths of the visible spectrum

Electromagnetic waves are transverse waves where the oscillations are per-
pendicular to the direction of travel. In all electromagnetic waves there are
actually two waves oscillating perpendicular to each other and to the direction
34 CHAPTER 6. WAVES

of travel. One is an oscillating magnetic field; the other an oscillation electric


field.

figs/chapt-6/emwave.png

Figure 6.2: oscillations in an electromagnetic wave

The electromagnetic spectrum is the name for the arrangement and classifi-
cation of electromagnetic waves in order of their wavelengths or frequencies.

The electromagnetic spectrum is shown below in order of increasing wave-


length.

You can see that the visible light spectrum makes up a small part of the
electromagnetic spectrum, with wavelengths between 400 - 700 nm.

(g) recall that the intensity of a wave is directly proportional to the square of
its amplitude

If the amplitude of a wave varies sinusoidally, the intensity will vary as sine
squared. Therefore the following expression can be used:

I ∝ A2
35

figs/chapt-6/emspectrum.jpg

Figure 6.3: The electromagnetic spectrum (Credit:miniphysics.com)

(h) use graphs to represent transverse and longitudinal waves, including stand-
ing waves

Note: Standing waves will be covered in Chapter 7 on Superposition

There are two types of graphs used to represent transverse and longitudinal
waves, shown in Figure 6.4. You need to be careful as they look similar.

The first graph plots the motion of one part of the wave with time, for example
the motion of one water molecule as a water wave goes by. The x-axis on this
graph can give you the time period of the wave.

The second graph is a snapshot of a section of the wave at one particular


36 CHAPTER 6. WAVES

instant in time. On this graph the wavelength can be measured from the
x-axis.
y

T
y λ

Figure 6.4: Two graphs of a wave

(i) explain what is meant by a plane-polarised wave

(j) recall Malus’ Law (I ∝ cos2 θ) and use it to calculate the amplitude and
intensity of transmission through a polarising filter

A plane-polarised wave is one where there is only one allowed direction of os-
cillation. This is only applicable to transverse waves where there are multiple
allowed modes of oscillation which are all perpendicular to the direction of
travel. A longitudinal wave cannot be polarised as there is already only one
direction of oscillation - the direction parallel to that of travel. All electro-
magnetic waves can be polarised.

Consider visible light as an example of a polarised wave. There are 4 ways in


which light can be polarised.

• Transmission: A polarising filter can be used to polarise light. A filter


is made up of chains of molecules that will absorb one direction of oscil-
lation of the light wave, therefore only letting through the perpendicular
direction. Note that this ’one’ direction is a simplificiation as it encom-
passes oscillations in both the electric and magnetic fields. The axis of
37

transmission of a filter is the direction of oscillation that the filter will


let through.

figs/chapt-6/polarisedwave.JPG

Figure 6.5: diagram showing the operation of a polarising filter (Credit: isaac-
physics)

As unpolarised light passes through a polaroid filter, its intensity will


drop of 50% of what it originally was. If light that is already polarised is
incident on a filter with a perpendicular axis of transmission, none will
pass through. If light that is already polarised is incident on a filter with
a parallel axis of transmission, then all of the light will pass through.
For cases other than parallel or perpendicular, Malus’ Law can be used.
Malus’ Law can be used to work out how the intensity of polarised light
changes as it passes through a polaroid filter. The angle θ is the angle
between the direction of polarisation of the incident light and the axis
of transmission of the polaroid. If you start with unpolarised light, θ is
the angle between the two polaroids.
Malus’ Law states that the intensity of the transmitted light is propor-
tional to the square of cos θ.

I ∝ cos2 θ
38 CHAPTER 6. WAVES

If the incident intensity is I0 , then we can write Malu’s Law as:

I = I0 cos2 θ

Note that if you are dealing with amplitude instead of intensity then you
must take the square root to give cos θ.

• Reflection: Light can be partially polarised on reflection from certain


non metallic surfaces, such as water. The reflected light will be polarised
parallel to the surface. This is why polaroid sunglasses are useful as they
can cut out the glare from water or roads.

• Refraction: Light can be partially polarised, often in two perpendic-


ular directions, when passing through some materials, such as calcite.
Specific details will always be given to you in a question.

• Scattering: Light from the Sun scatters of molecules in our atmosphere


and is partially polarised depending on the direction that you are looking
at the sky. Again, specific details will always be provided in a question.

(k) recognise and use the expression for refractive index


sin θ1 v1
n= =
sin θ2 v2

When a wave crosses a boundary which involves a change in speed, refraction


occurs. This concept should be familiar from GCSE.

For light, the refractive index of a medium is the ratio of the speed of light in
a vacuum, c, to the speed of light in the medium, v.
c
n=
v

Therefore the refractive index of a material is always greater than one.

If a wave now crosses a boundary between material 1 and material 2, with the
angle of incidence being θ1 and the angle of refraction being θ2 , the following
relationship (Snell’s Law) applies:

n2 sin θ1
=
n1 sin θ2

As the refractive index of a material is inversely proportional to the speed of


light in that material, we know that
39

n2 v1
=
n1 v2

Snell’s Law now becomes

n2 sin θ1 v1
= =
n1 sin θ2 v2

This is the most general form of Snell’s Law. For the specific case where
material 1 is air we can take n1 = 1 as the speed of light in air is so close to
the speed of light in a vacuum. Now, replacing n2 with n, the equation is:

sin θ1 v1
n= =
sin θ2 v2

This is the equation given in the specification. Be careful as it only applies to


the case where material 1 is air and this might not always be the case.

(l) derive and recall sin c = 1


n and use it to solve problems

If we take Snell’s Law for the case where light is travelling from a material of
higher refractive index into a material with lower refractive index, n1 > n2 ,
we know that the light will bend away from the normal with the angle of
refraction, θ2 being larger than the angle of incidence, θ1 . If the angle of
incidence is increased until the angle of refraction is 90◦ , then the angle of
incidence is now called the critical angle, as above this angle, total internal
reflection will occur.

Now we can put this into Snell’s Law. θ1 is now c, the critical angle and θ2 is
now 90◦ .

n2 sin θ1
=
n1 sin θ2
This now becomes:
n2 sin c
=
n1 sin 90◦

As sin 90◦ = 1, the most general equation to find the critical angle is:
n2
= sin c
n1
or
n1 1
=
n2 sin c
40 CHAPTER 6. WAVES

In the specification, the equation is given for the specific case where material
2 is air, therefore n2 can be taken to be 1. This gives the equation:
1
sin c =
n

(m) recall that optical fibres use total internal reflection to transmit signals

(n) recall that, in general, waves are partially transmitted and partially re-
flected at an interface between media.

Should be familiar from GCSE.


7 Superposition

Content

• phase difference

• diffraction

• interference

• standing waves

Candidates should be able to:

(a) explain and use the concepts of coherence, path difference, superposition
and phase

(b) understand the origin of phase difference and path difference, and calculate
phase differences from path differences

(c) understand how the phase of a wave varies with time and position

These terms are all used when considering more than one wave.

The phase of a wave is related to how far through an oscillation a wave is. This
is expressed in radians or degrees, where one complete oscillation corresponds
to 360◦ or 2π radians.

The phase difference between two waves is more useful than the phase of one
wave. This refers to the fraction of an oscillation by which one wave ’leads’
or ’lags’ behind another. If the phase difference is 2nπ, where n is an integer,
then two waves are said to be in phase and if the phase difference is (2n − 1)π
then the waves are completely out of phase.

41
42 CHAPTER 7. SUPERPOSITION

Two waves are said to be coherent if they have a constant phase difference.
Most often, this is a phase difference of zero, which means that the waves
are in phase, but this does not always have to be the case. For interference
patterns to occur, coherence is often a necessary condition.

If two waves, from two different sources, meet at a point, the path difference
is the difference in distance travelled between the two waves. To calculate
the path difference the smaller distance should be taken away from the larger
distance. Path difference is normally expressed as a multiple of wavelength,
as this then allows the phase difference to be calculated easily.

When two or more waves meet at a point, superposition will occur. This means
that the displacements of the individual waves add up to give a resultant
displacement. If the two waves are in phase, then constructive interference
will occur and if they are out of phase then destructive interference will occur.

Path difference Phase difference Superposition


nλ 2nπ constructive interference
(n + 12 )λ (2n − 1)π destructive interference

(d) determine the resultant amplitude when two waves superpose, making use
of phasor diagrams

When two waves superpose, the resultant displacement at any point is the
vector sum of the individual displacements.

Phasors are rotating arrows that can be used to describe waves. A phasor
arrow rotates anticlockwise and one full oscillation of the wave corresponds to
one complete oscillation of the phasor arrow. The length of the phasor arrow
corresponds to the amplitude of the wave.

If we have two waves that superpose, their individual phasor arrows at a


particular point can be added up as vectors as shown in the diagram below.

(e) explain what is meant by a standing wave, how such a wave can be formed,
and identify nodes and antinodes

A standing wave arises from a combination of reflection and interference. Con-


sider the set up below.

The vibration generator leads to a progressive wave travelling to the right.


This waves reflects off the fixed end and so there are now two waves on the
string, travelling in opposite directions.
43

figs/chapt-7/superposition.JPG

Figure 7.1: Graph showing superposition of two waves

These two waves superpose and, in some cases (conditions discussed below), a
standing wave can form on the string. If these conditions are met, there will be
points where the two waves always meet in phase and interfere constructively.
These are called antinodes. The points where the two waves always meet out
of phase and interfere destructively are called nodes.

Unlike a progressive wave, all points on a stationary wave do not have the
same amplitude. The amplitude is at a minimum (often zero) at a node and
at a maximum at an antinode.

If the length of the string is L, and the speed of waves on the string is v,
we can work out the frequencies of the various modes of oscillation. The
lowest frequency (first mode) shown in the diagram is called the fundamental
frequency. You can see that half of a wavelength fits on the string. Therefore
we can write:
λ
=L
2
λ = 2L

Putting this into the wave equation gives an expression for the fundamental
44 CHAPTER 7. SUPERPOSITION

figs/chapt-7/phasor.JPG

Figure 7.2: The circle on the left shows the rotating phasors that correspond
to this sine wave

frequency:
v = f λ = f.2L
v
f=
2L

The other modes of oscillation can be worked out in similar ways, by looking
at the relationship between L and λ and substituting into the wave equation.
For a given string under a certain tension the speed is constant. The frequency
is the frequency of the vibration generator which can be changed to give the
different modes of oscillation.

The boundary conditions for this string were that both ends had to be nodes.
45

Figure 7.3: The sum of two phasors placed end to end

figs/chapt-7/melde.JPG

Figure 7.4: Vibration generator connected to a horizontal string under tension

In other cases where standing waves occur, for example sound waves, the
boundary conditions could be different. Closed ends of tubes are always nodes
and open ends are always antinodes.

(f) understand that a complex wave may be regarded as a superposition of


sinusoidal waves of appropriate amplitudes, frequencies and phases

If more than two waves superpose, the resultant wave can get very compli-
cated! This means that any waveform can always be broken down into sinu-
soidal waves. With combinations of sinusoidal waves of various frequencies,
amplitudes and phase differences, any waveform can be made.
46 CHAPTER 7. SUPERPOSITION

figs/chapt-7/string.JPG

Figure 7.5: modes of oscillation on a string

(g) recall that waves can be diffracted and that substantial diffraction occurs
when the size of the gap or obstacle is comparable to the wavelength

(h) recall qualitatively the diffraction patterns for a slit, a circular hole and a
straight edge

When waves pass through an opening, or around a barrier, diffraction occurs


and the waves can change in direction and spread out. Diffraction is most sig-
nificant when the size of the gap or obstacle is comparable to the wavelength.
For example, sound waves will diffract through open doors as they have wave-
lengths of similar orders of magnitudes to the size of the door. However, light
will not diffract as much through a door as the wavelength of light is many
orders of magnitude smaller than the size of the door.

When waves diffract, diffraction patterns will be formed due to interference


of the waves. Specific cases will be discussed below. Here are some examples
of diffraction patterns:

A slit
47

figs/chapt-7/sound.JPG

Figure 7.6: Standing waves in a tube credit:Yonsei Phylab

When a wave passes through a slit, and the wave is observed a certain distance
away from the slit, a diffraction pattern consisting of points of constructive
interference and destructive interference will be formed. Visible patterns will
occur when the size of the slit is comparable to the wavelength. For example,
with visible light and a very narrow slit, the following pattern will be observed.

There is a central maximum which is twice the width of the maxima on either
side and the pattern is symmetrical.

A circular hole

The diffraction pattern for a circular hole is similar to a single slit, except that
instead of fringes, the pattern consists of rings.

Although the examples given here are for visible light, remember that all
waves can diffract.

A straight edge

A similar pattern of fringes is seen, due to constructive and destructive inter-


48 CHAPTER 7. SUPERPOSITION

figs/chapt-7/sine1.JPG

Figure 7.7: E
xample showing how 2 waves combine to form a more complex wave
(credit:cyberphysics.co.uk)

ference. Here the example given is for radio waves.

(i) recognise and use the equation nλ = bsinθ to locate the positions of de-
structive superposition for single slit diffraction, where b is the width of the
slit

When an electromagnetic wave, such as light, travels through a single slit, it


will diffract. Therefore light from the slit reaches many points on a screen
placed at some distance from the slit.

Consider a point on the screen. Light will reach this point from all points
within the slit. The distances travelled from various points within the slit to
the point on the screen will be different, and so there will be a path and phase
difference. Therefore, as we move along the screen, there will be points of
constructive interference and points of destructive interference.

It can be shown that for a single slit of width b, the points of destructive
interference will be at an angle θ given by the equation nλ = bsinθ Here, n,
49

figs/chapt-7/slit.JPG

Figure 7.8: Diffraction pattern for a slit (credit:hyperphysics)

refers to the order of the minimum being considered.

(j) recognise and use the Rayleigh criterion θ ≈ λb for resolving power of a
single aperture, where b is the width of the aperture

As light travels through an instrument, such as the eye, or a telescope, it


passes through a gap or aperture and will diffract. The diffraction pattern
will depend on the shape of the aperture and the width.

If light from two different objects passes through the aperture, there will
be two diffraction patterns that overlap. The Rayleigh criterion tells us the
minimum angle between the two objects at which it is still possible to see
them as separate. This is when the first diffraction minimum of one pattern
coincides with the central maximum of another.

Working in radians, so that the approximation sin θ ≈ θ, then this minimum


angle can be approximated by:

λ
θ≈
b
50 CHAPTER 7. SUPERPOSITION

figs/chapt-7/slitphoto.jpg

Figure 7.9: Photograph of single slit diffraction pattern

(k) describe the superposition pattern for a diffraction grating and for a double
slit and use the equation d sin θ = nλ to calculate the angles of the principal
maxima

(l) use the equation λ = ax


D for double-slit interference using light

When light, or any other coherent waves, pass through a double slit, a diffrac-
tion pattern consisting of evenly spaced fringes is seen. This is due to the
waves from each slit interfering with each other.

You can see that there is also a single slit envelope, which is much wider than
the double slit pattern.

If the number of slits is increased, the pattern becomes more defined, and the
fringes get narrower. Therefore for a diffraction grating, the pattern consists
of sharp, evenly spaced peaks or bright spots.

If we consider two adjacent slits, you can see that for constructive interference
to occur the following must be true:
51

figs/chapt-7/hole.JPG

Figure 7.10: Circular hole diffraction pattern credit:hyperphysics

d sin θ = nλ (7.1)

d is the distance between adjacent slits and n is the order of the maxima that
is being considered.

Equation 7.1 can be re-arranged to give


sin θ = (7.2)
d

This formula can be used for any waves, where there is diffraction from more
than one slit.
52 CHAPTER 7. SUPERPOSITION

figs/chapt-7/edge.jpg

Figure 7.11: Diffraction at a straight edge credit:University of Alberta

Now consider a double slit pattern, where the screen is now a distance D away
from the slits and the slit separation is a instead of d. Bright fringes will be
equally spaced on the screen and we can call the fringe separation x.

For the first maxima, n = 1, the equation from before now becomes:
λ
sin θ =
a
but from the diagram you can see that:
x
tan θ =
D

If θ is small, then the small angle approximation can be used. This is true if
D is much larger than a.

sin θ ≈ tan θ
λ x
=
a D
53

figs/chapt-7/singleslit.JPG

Figure 7.12: Single slit diffraction credit:hyperphysics

ax
λ=
D
54 CHAPTER 7. SUPERPOSITION

figs/chapt-7/double.JPG

Figure 7.13: Double slit diffraction (credit:hyperphysics)


55

figs/chapt-7/diffractioneq.JPG

Figure 7.14: Derivation of grating formula

Figure 7.15: The double slit pattern


8 Atomic and Nuclear
Processes

Content

• the nucleus

• nuclear processes

• probability and radioactive decay

• fission and fusion

Candidates should be able to:

(a) understand the importance of the α-particle scattering experiment in de-


termining the nuclear model

In the famous scattering experiment an alpha source was placed in a lead


container to produce a beam of alpha particles which were directed at a piece
of gold leaf. Given the high energies of these alpha particles it was expected
that they would simply crash through the gold leaf like a bullet through tissue
paper. Instead, it was found that a few alpha particles were scattered through
large angles and a small number came straight back.

57
58 CHAPTER 8. ATOMIC AND NUCLEAR PROCESSES

Figure 8.1: Scattering of alpha particles by gold atoms

This occasional scattering implied that the gold nuclei are very small compared
to the size of the gold atom and contain almost all of the atom’s mass.

(b) describe atomic structure using the nuclear model

The nuclear model has a very small (10−15 m) nucleus which is positively
charged and made up of protons and neutrons. Around this nucleus move the
much lighter, negatively charged electrons.

(c) show an awareness of the existence and main sources of background radi-
ation

The radiation referred to in this part of the specification is ionising radiation.


This is always present and is composed of both natural and artificial sources.
The most common sources are shown in figure 8.2.
59

figs/chapt-8/rad-sources.jpg

Figure 8.2: Background Radiation Sources in the UK (http://www.npl.co.


uk/educate-explore/factsheets/ionising-radiation/)

(d) recognise nuclear radiations (α, β−, γ) from their penetrating power and
ionising ability, and recall the nature of these radiations

Radiation Penetrating Power Ionising Ability Nature


α low high helium-4 nucleus
β medium medium electron
γ high low electromagnetic wave
Table 8.1: Three types of radiation

Table 8.1 shows the properties and nature of three types of radiation. Note
that the penetrating power and ionising ability are inversely proportional as if
a type of radiation is highly ionising it will interact with matter and therefore
be stopped easily.
60 CHAPTER 8. ATOMIC AND NUCLEAR PROCESSES

(e) write and interpret balanced nuclear transformation equations using stan-
dard notation

In nuclear equations the individual species are written as A


Z X where A is the
nucleon number, Z is the proton number and X is the chemical symbol for
the element. The three types of radiation are written as follows:

α 4 He β 0 e γ 0γ
2 −1 0

The electron is given a ‘proton number’ to enable the balancing of nuclear


equations. In beta decay a neutron is transformed into a proton and an
electron, therefore the nucleon number of the daughter does not change but
the proton number increases by one. The beta particle’s proton number of −1
balances this change.

In a balanced equation the sum of the proton numbers on each side must be
equal, as must the sums of the nucleon numbers.

Example Question
Write a balanced nuclear equation for the decay of carbon-14 by beta
minus decay.
Answer

7 N +−1 e
14
6 C → 14 0

A : 14 = 14 + 0
Z :6=7−1

(f) understand and use the terms nucleon number (mass number), proton num-
ber (atomic number), nuclide and isotope

Nucleon number The total number of protons and neutrons in the nucleus.

Proton number The number of protons in the nucleus.

Nuclide A specific number of protons and neutrons which makes a unique


nucleus. Usually defined by a combination of proton number / element
name and mass number. E.g. carbon-14, 146 C.

Isotope Isotopes are nuclei of the same element with different mass numbers,
i.e. the same number of protons but different numbers of neutrons.
61

(g) appreciate the spontaneous and random nature of nuclear decay

Radionuclei decay spontaneously and randomly because of the fundamentally


quantum nature of the process. This means that while one can make predic-
tions about a collection of nuclei using half-life it is impossible to predict when
any particular nucleus will decay.

(h) define and use the concept of activity as the number of decays occurring
per unit time

Activity is defined in this way, but measured by counting the number of ioni-
sation events occurring in a detector per unit time.

(i) understand qualitatively how a constant decay probability leads to the shape
of a radioactive decay curve

A constant probability of decay means that the total rate of decay of a sample
is proportional to the number of nuclei remaining. This means that the number
of nuclei remaining decreases with decreasing rate. The activity is proportional
to the number of nuclei remaining so it also decreases with decreasing rate.

(j) determine the number of nuclei remaining or the activity of a source after
a time which is an integer number of half-lives

One half-life is the time taken for half of a sample to decay. Therefore after
 3
three half-lives the amount remaining is 1
2 N0 where N0 is the original
amount.

(k) understand the terms thermonuclear fusion, induced fission and chain re-
action

Thermonuclear fusion This is the combining of two nuclei to form a new,


heavier nucleus. It is the type of nuclear reaction which occurs in stars
where hydrogen nuclei fuse to form helium. The electrostatic repulsion
between the nuclei is overcome by the thermal energy of the particles
involved, meaning that thermonuclear fusion can only occur at extremely
high temperatures.

Induced fission This is the splitting of a large nucleus into two smaller,
daughter nuclei. This does not occur spontaneously but must be induced
by the collision of a neutron with the nucleus.
62 CHAPTER 8. ATOMIC AND NUCLEAR PROCESSES

Chain reaction When a neutron induces fission in a large nucleus it is usu-


ally the case that as well as the two daughter nuclei, several further
neutrons are released. These neutrons can then go on to cause further
induced fission events in other heavy nuclei. This is known as a chain
reaction.
9 Quantum Ideas

Content

• the photoelectric effect


• the photon
• wave-particle duality

Candidates should be able to:

(a) recall that, for monochromatic light, the number of photoelectrons emit-
ted per second is proportional to the light intensity and that emission occurs
instantaneously

(b) recall that the kinetic energy of photoelectrons varies from zero to a max-
imum, and that the maximum kinetic energy depends on the frequency of the
light, but not on its intensity

(c) recall that photoelectrons are not ejected when the light has a frequency
lower than a certain threshold frequency which varies from metal to metal

(d) understand how the wave description of light fails to account for the ob-
served features of the photoelectric effect and that the photon description is
needed

Each of the above features (a) - (c) is discussed in turn below.

(a) In the wave description of light the energy an electron in the metal
receives energy from the light arrives continuously. An electron there-
fore gradually absorbs enough energy to escape from the surface of the

63
64 CHAPTER 9. QUANTUM IDEAS

metal, something which is not seen in practice. Additionally, increasing


the intensity of a wave corresponds to increasing the amplitude of the
oscillation. This would lead us to expect that increasing the intensity of
light would give photoelectrons of a higher energy, rather than simply
more of them.
The photon description of light accounts of both of these effects. When
light arrives on the surface of the metal a single photon interacts with a
single electron. Assuming the photon gives the electron enough energy to
escape, the electron will leave the metal instantaneously. In the photon
model, the intensity of the light is due to the number of photons arriving
per second. Therefore if the intensity doubles, the number of photons
doubles and the number of photoelectrons doubles.

(b) When the photoelectrons leave the metal some of the photon energy is
used to break away from the metal and the remainder goes into their
kinetic energy. Different photoelectrons have different amounts of en-
ergy, however there is a maximum kinetic energy the photoelectrons are
found to have and this depends only on the frequency of the incoming
radiation.
The wave model of light would allow different electrons to absorb dif-
ferent amounts of energy and therefore this relationship would not be
seen.

(c) The energy from a photon of light is split between the energy required
to escape the metal and the kinetic energy of the photoelectron. If the
photon does not have enough energy to enable to the electron to escape
the metal then no emission of photoelectrons is seen.
The wave model would still allow emission as a single electron could
absorb energy from the wave over a longer period of time. However, since
there are so many electrons in the surface of the metal it is vanishingly
unlikely that a single photoelectron will interact with two photons.

(e) recall that the absorption of a photon of energy can result in the emission
of a photoelectron

As described above.

(f) recognise and use E = hf

The energy of a photon of light is related to the frequency of that radiation


using the equation E = hf where h is Planck’s Constant which has a value of
6.626 × 10−34 J s.
65

(g) understand and use the terms threshold frequency and work function and
recall and use
1
hf = φ + mvmax2
(9.1)
2

Threshold Frequency, f0 . This is the minimum frequency required for the


emission of photoelectrons to occur. This varies from metal to metal.

Work Function, φ. This is the energy required to remove an electron from


the surface of the metal.

Equation 9.1 expresses the sharing of the energy of the photon (hf ) between
the work function (φ) and the kinetic energy of the photoelectron.

(h) understand the use of stopping potential to find the maximum kinetic en-
ergy of photoelectrons and convert energies between joules and electron-volts

We can measure the energy of the photoelectrons by placing a photocell in a


circuit and using a potential difference to stop the flow of electrons.

µA

Figure 9.1: Measuring stopping potential

Figure 9.1 shows a simple set-up to measure the stopping potential for pho-
toelectrons. Monochromatic light is shone on the photocell. Electrons leave
the surface of the metal and travel around the circuit creating a small cur-
rent. The variable power supply is gradually increased until no current flows
through the circuit. At this point none of the electrons leaving the surface of
the metal in the photocell have enough energy to cross the potential differ-
ence and create a current in the circuit. At this point the maximum energy of
66 CHAPTER 9. QUANTUM IDEAS

the photoelectrons can be equated to the energy required to cross a potential


difference V :
1
mv 2 = eV (9.2)
2 max

Since we are measuring the energies of electrons using potential differences, it


makes sense to define a unit of energy in terms of these potential differences.
Hence, 1 electronvolt is defined as the energy transferred by an electron moving
through a potential difference of 1 volt.

1 eV = 1.6 × 10−19 J (9.3)

(i) plot a graph of stopping potential against frequency to determine the Planck
constant, work function and threshold frequency

By repeating the experiment shown in Figure 9.1 for different frequencies of


light and measuring the stopping potential for each frequency we get the graph
shown in Figure 9.2.

stopping
potential
/V

threshold frequency
frequency, / Hz
f0

Figure 9.2: Stopping potential against threshold frequency

Equation 9.1 and 9.2 can be combined with and re-arranged to give
h φ
V = f− (9.4)
e e

Equation 9.4 describes the linear relation seen on the graph in Figure 9.2.
The graph has a gradient of he which enables the determination of the Planck
Constant and the intercept with the x-axis gives the threshold frequency. The
67

work function is simply the energy of a photon with the threshold frequency,
i.e.
φ = hf0 (9.5)

(j) understand the need for a wave model to explain electron diffraction

When electrons are fired at two closely spaced slits the result is entirely unlike
what one would expect from a particle model. A particle model would predict
that each electron would either go through one slit or the other, creating two
regions at which electrons are found as shown in Figure 9.3.

Figure 9.3: Expected pattern for electron diffraction

However, this behaviour is not seen but rather electrons form an interference
pattern similar to that seen for light, as shown in Figure 9.4.

The mystery here is how can single particles form an interference pattern? If
electrons are fired through two slits one at a time then they initially appear to
arrive randomly. As more and more arrive they fill in the interference pattern.
We explain this by saying that the probability of their arrival is determined
by a form of wave-mechanics and as particles arrive they fill in the probability
pattern as shown in figure 9.5.
68 CHAPTER 9. QUANTUM IDEAS

Figure 9.4: Observed pattern for electron diffraction

figs/chapt-9/electron-diff.jpg

Figure 9.5: Build up of the interference pattern of electrons. Credit: Belsazar,


Wikimedia Commons
69

(k) recognise and use


h
λ= (9.6)
p
for the de Broglie wavelength.

In order to explain the wave-like nature of electrons we need to be able to


assign them a wavelength. de Broglie’s thesis is that all particles have a
wavelength which is defined in equation 9.6. This brings wave-particle duality
from photons and electrons to all particles. The natural question here is
why isn’t interference ordinarily observed? The answer is that the de Broglie
wavelength for macroscopic objects is extremely small and therefore these
objects behave as particles, travelling in straight lines and not undergoing
interference.

Example Question
Electrons will be diffracted by crystal lattices if they have a wavelength
of around 0.1 nm. Calculate the speed and energy of such electrons.
Answer
The speed of these electrons can therefore be calculated as follows:
h h
λ= =
p mv
h
v= = 7.27 × 106 m s−1

1
E = mv 2 = 151 eV
2
Part B

71
10 Rotational Mechanics

This chapter contains revision on the topics of:

• kinematics of uniform circular motion

• centripetal acceleration

• moments of inertia

• kinematics of rotational motion

Candidates should be able to:

(a) Define and use the radian

An angle in radians is defined by the length of the arc of circle it subtends


divided by the radius of the circle. Numerically 2π radians is equivalent to
360 degrees.

(b) Understand the concept of angular velocity, and recall and use the equations
v = rω and T = 2πω

v2
(c) Derive, recall and use the equations for centripetal acceleration a = r
and a = rω 2
mv 2
(d) Recall that F = ma applied to circular motion gives resultant F = r

(e) Describe qualitatively the motion of a rigid solid object under the influence
of a single force in terms of linear acceleration and rotational acceleration.

When a rigid solid object has a single force applied to it, it may just result in
linear acceleration, if the force acts through the centre of mass of the object.
However it is more likely that the force will not act through this point, and
therefore also cause some rotational acceleration, causing the object’s angular
velocity to change.

73
74 CHAPTER 10. ROTATIONAL MECHANICS

The remaining sections in this chapter deal with how we describe this ro-
tational acceleration, but it should be noted that they are asterisked sections
so will only form part of section 2 of paper 3.

(f) *Recall and use I = Σmr2 to calculate the moment of inertia of a body
consisting of three or fewer point particles fixed together

The moment of inertia of a body is the rotational analogue of mass, and


basically describes how resistant an object is to angular acceleration when a
torque is applied (in the same way that mass describes how resistant an object
is to accelerating linearly when a force is applied...)

A point mass m at a distance r from an axis of rotation will have Moment of


Inertia I = mr2

More generally the moment of inertia is the sum of all the mr2 values for all
the point masses that make up an object. For a small number of particles this
can just be found by adding the values of individual moments of inertia.

(g) *Use integration to calculate the moment of inertia of a ring, a disk and
a rod

For more complex bodies, the summing of moments of inertia needs to be


done by setting up an integration. This is acheived by summing moments of
inertia for a series of small sections of the object, strips orR rings of thickness
δx say, and using the fact that as δx → 0 the sum Σδx ≡ dx

As an example, consider a disk of radius r , thickness t and density ρ. The


moment of inertia about its central axis could be found by splitting it up into
rings of radius x and thickness δx. See figure 10.3.

Each ring has moment of inertia equal to its mass multiplied by x2 , as all the
particles in it have (approximately) the same distance from the axis.1

Hence Iring = (2πxtρδx)x2 = 2πρx3 δx

1
This can be proved by another, much simpler, sum. If you imagine the ring as the sum
of many small points of mass δm each at the same radius r, the moment of inertia becomes
Σr2 δm which is the same as r2 Σδm and hence I = mr2
75

axis of rotation

r
δx x

Figure 10.1:

To find the moment of inertia of the disk, we sum all the rings and let their
size δx tend to 0.

Idisk = Σ2πρx3 δx

Rr
Idisk = 0 2πρtx3 δx

giving

2πρtr4
Idisk = 4

which is equal to
mr2
Idisk = 2

2
Other objects would be derived in a similar way. You should get I = mr3 for
mr2
a rod rotating about one end, and I = 12 for a rod rotating about its centre
of mass.
76 CHAPTER 10. ROTATIONAL MECHANICS

(h) *Deduce equations for rotational motion by analogy with Newton’s laws for
linear motion, including E = 21 Iω 2 , L = Iω and Γ = I dω
dt

Here is a table outlining the analogies between linear and rotational moton:

Linear Motion Rotational Motion


Mass m Moment of Inertia I
linear velocity v Angular velocity ω
Force F Torque Γ
Linear Momentum p Angular Momentum L
p = mv L = Iω
F = ma = m dv dt Γ = I dω
dt
KElinear = 21 mv 2 KErotational = 21 Iω 2

(i) *Apply the laws of rotational motion to perform kinematic calculations


regarding a rotating object when the moment of inertia is given.

Example Question
The moment of inertia of a large flywheel in a factory is 60 kgm2 .
Calculate how long it would take the flywheel to obtain an angular
velocity of 6.0 rad s−1 when a torque of 24 Nm was applied.
Answer
First use Γ = I dω
dt to find the angular acceleration:

dt = Γ
I = 24
60 = 0.4 rad s−2
Then the time taken will be
T = 0.4
6
= 15 seconds
11 Oscillations

Simple Harmonic Motion

(a) Recall the condition for simple harmonic motion and hence identify situ-
ations in which simple harmonic motion will occur

(b) * show that the condition for simple harmonic motion leads to a differential
equation of the form
d2 x
= −ω 2 x
dt2
and that
x = Acosωt
is a solution to this equation

(c) * use differential calculus to derive the expressions

v = −Aωsinωt

and
a = −Aω 2 cosωt
for simple harmonic motion.

This is achieved very simply by differentiating our expression for displacement


dx dv
with respect to time as velocity = and acceleration =
dt dt
Remember that the differential of cos is -sin.

so
x = Acosωt
dx
v= = −Aωsinωt
dt
dv
a= = −Aω 2 cosωt
dt

77
78 CHAPTER 11. OSCILLATIONS

(d) understand the phase differences between displacement, velocity and ac-
celeration in simple harmonic motion

(e) *recognise and use the expressions x = Acosωt, v = −Aωsinωt, a =


−Aω 2 cosωt and F = –mω 2 x to solve problems

(f) recall and use T = as applied to a simple harmonic oscillator
ω
Angular frequency w is how many radians per second the oscillator goes
through with one complete cycle being 2π radians.
1
So is how many seconds is takes to complete one radian.
ω

and the time for one complete cycle is T =
ω
(g) *show that the total energy of an undamped simple harmonic system is
1
given by E = mA2 ω 2 and recognise that this is a constant
2
(h) distinguish between free, damped and forced oscillations

A free oscillator is one which is set in motion and left to oscillate without any
external forces.

e.g. a pendulum.

A forced oscillation is one where an external periodic force is applied.

e.g. pushing a child on a swing.

Damping is where frictional forces remove energy from the system and the
oscillations die down.

This can be increased intentionally for example adding thick oil to car sus-
pension to prevent you bouncing around every time the car hits a bump.

If a system is lightly damped it will gradually come to rest after a number of


cycles.

An important case is critical damping where the system comes to rest without
overshooting and in the shortest possible time.
SIMPLE HARMONIC MOTION 79

(i) recall how the amplitude of a forced oscillation changes at and around the
natural frequency of a system and describe, qualitatively, how damping affects
resonance.
12 Electric Fields

Content

• concept of an electric field

• uniform electric fields

• capacitance

• electric potential

• electric field of a point charge

Candidates should be able to:

(a) explain what is meant by an electric field and recall and use E = F
q for
electric field strength

An electric field is the region of space in which electrical forces are exerted
on charged bodies. The definition of electric field strength is force per unit
charge, and this is written in symbols as:
F
E=
q
Electric field therefore has units of N C−1 .

(b) recall that applying a potential difference to two parallel plates stores charge
on the plates and produces a uniform electric field in the central region between
them

If an electric potential is created between two parallel plates a distance d


apart then an electric field exists between them. Except at the edges, the field
is uniform. This is shown in figure 12.1 by the fact that the field lines are
parallel.

81
82 CHAPTER 12. ELECTRIC FIELDS

++++++++++++++++++++++++++

- - - - - - - - - - - - - - - - - - - - - - - - - -

Figure 12.1: Uniform field between parallel plates

Note that field lines show the direction in which a force would act on a postive
charge so they are from positive to negative.

(c) derive and use the equations Fd = QV and E = Vd for a charge moving
through a potential difference in a uniform electric field

From the definition potential difference we can say that the work done moving
through a potential difference is

W = QV

and that is equal to the force multiplied by the distance moved,

Fd

If the movement is through a uniform field we can equate these two to give

F d = QV

The definition of electric field strength gives F = QE and therefore substitu-


tion and re-arrangement give
V
E=
d

(d) recall that the charge stored on parallel plates is proportional to the poten-
tial difference between them

This is an application of Gauss’ law which relates the electric field strength
around an object to the charge contained within a surface. It is enough to
recall this fact at Pre-U.
83

Q
(e) recall and use C = V for capacitance

This is the definition of capacitance and should be learnt. It can also be


calculated from the gradient of a graph of Q against V .

(f) recognise and use W = 21 QV for the energy stored by a capacitor, derive
the equation from the area under a graph of charge stored against potential
difference, and derive and use related equations such as W = 12 CV 2

If a capacitor is partially charged with a charge Q then to increase its potential


difference by δV will require a small amount of work δW such that

δW = QδV

If a graph is plotted of Q against V then this can be seen as the area of a small
section. Thus, the energy required to charge a capacitor from uncharged to a
p.d. of V is given by the area of a graph of Q against V from zero to V , hence
1
W = QV
2

We can substitute for each quantity in turn to give the following variations of
the equation:
1 1 1 Q2
W = QV = CV 2 =
2 2 2 C

This can, of course, also be done with integration

1
Z V Z V
W = QdV = CV dV = CV 2
0 0 2

(g) analyse graphs of the variation with time of potential difference, charge
and current for a capacitor discharging through a resistor

When a capacitor discharges through a resistor the potential difference on


the capacitor drives current around the circuit. Since it is this current which
removes charge from the capacitor, it is the case that the rate of change of
charge on the capacitor is proportional to the charge on the capacitor.

dQ V Q
= −I = − = −
dt R RC

This is a first order differential equation and has a solution


84 CHAPTER 12. ELECTRIC FIELDS

t
Q = Q0 e− RC

The substitutions Q = CV and then I = VR can be used to get similar equa-


tions for each of the above. The graph of this change is exponential decay.
The key features are the initial value (e.g. V0 ) and the fact that all three
quantities tend to zero.

(h) define and use the time constant of a discharging capacitor

The time constant, τ , is defines as follows:

τ = RC

This means that in one time constant the current, voltage and charge on a
capacitor have declined to 1e of their initial values.

A common rule-of-thumb from electronics is that a capacitor takes five time


constants to discharge. If we plug this into the equation

t
V = V0 e− RC = V0 e−5 = 0.067 V0

So the voltage across the capacitor has declined to less than 1% of its initial
value.
−t
(i) analyse the discharge of a capacitor using equations of the form x = x0 e RC

Much of this has been covered above. One important point to note is that the
analysis of capacitor decay is usually carried out by plotting a graph of ln V
against time. This changes the equation to give
t
ln V = − + ln V0
RC
Therefore the gradient of the graph is − RC
1
and the intercept ln V0 .

(j) understand that the direction and electric field strength of an electric field
may be represented by field lines (lines of force), and recall the patterns of field
lines that represent uniform and radial electric fields

Field lines are a way of visualising the field. The direction of the field lines
is from north to south and show the direction in which a positively charged
particle will move. The density of field lines represent the strength of the field.
For example in figure 12.2 the region A contains a uniform strong field which
is stronger than the field at B as the field lines are more closely spaced.
85

Figure 12.2: Field around a solenoid

(k) understand electric potential and equipotentials

Electric potential is defined as the energy per unit charge due to an electric
field. Space without an electric field is defined as having zero potential. An
equipotential is a line or surface on which the potential is a constant value.
Therefore no work is done against the electric force moving along an equipo-
tential. On diagrams equipotentials always cross field lines at right angles.

(l) understand the relationship between electric field and potential gradient,
and recall and use E = − dX
dV

The strength of the electric field at any point is equal to the negative of the
potential gradient. This can be seen most easily in a uniform field. If a unit
charge is moved through a distance ∆x within a uniform field E then the work
done, F ∆x = −q∆V (for a unit charge). The negative sign comes from the
fact that if the force is doing work on the charge it must be moving in the
opposite direction to the force on that charge. Thus, the change in energy per
unit charge is given by ∆V = − Fq ∆x. As Fq is the field strength E, this can
be rearranged to give E = − ∆V ∆x . This can be generalised for a non-uniform
field as
dV
E=−
dx
86 CHAPTER 12. ELECTRIC FIELDS

(m) recognise and use


Q1 Q2
F =
4π0 r2
for point charges

The equation above is known as Coulomb’s law and enables the calculation of
the force between two point charges separated by a distance r.
Q
(n) derive and use E = 4πo r2
for the electric field due to a point charge

This is simply from the definition of electric field strength:

F 1 Q1 Q2 Q
E= = =
Q Q2 4π0 r 2 4πo r2

Q1 Q2 Q1 Q2
(o) *use integration to derive W = 4π0 r from F = 4π0 r2
for point charges

Since free space is defined as having zero potential, the electrostatic potential
energy, W , of a particle is equal to the work done by the field moving the
particle from a distance r to infinity. Since work done is equal to F dx it
R

follows: Z ∞
Q1 Q2 Q1 Q2
W = dr =
r 4π0 r 2 4π0 r

You can define this alternatively by thinking about the work done bringing a
charged particle from infinity to a distance r. If you do this it is important to
remember that F acts in the opposite direction to r so the limits of integration
are reversed and the formula has a minus sign, thus reaching the same answer.
Q1 Q2
(p) *recognise and use W = 4π0 r for the electrostatic potential energy for
point charges.

This is simply applying the equation above. This will allow the calculation
of changes in potential energy and possibly transfer to other forms of energy
(e.g. kinetic).
13 Gravitation

(a) state Kepler’s laws of planetary motion Before you learn Kepler’s laws
(which you MUST learn) you should spend some time looking through the
chapter on rotational mechanics and making absolutely sure that you know
how to apply Newton’s laws of motion to an orbiting body. This is vital or
you won’t get the maths in this chapter.

The specification dictates that you need to be able to state Kepler’s Laws.
They are as follows:

1. Planets move in elliptical orbits with the Sun at one focus. (motion in
ellipses is not part of the specification, so don’t worry about the mathe-
matics of this – we approximate to a circle for pre-U)

2. The Sun-planet line sweeps out equal areas in equal times.

3. The orbital period squared of a planet is proportional to its mean dis-


tance from the Sun cubed.

87
88 CHAPTER 13. GRAVITATION

First Law

figs/chapt-13/kepler-1.png

Figure 13.1: Kepler’s First Law

Of course, it looks nothing like this – this is MASSIVELY exaggerated for the
sake of seeing what is going on. The Earth’s orbit around the Sun is very
nearly circular, which is why it took so long for astronomers to realize that it
wasn’t.

Second Law

Figure 13.2 shows the areas swept out in the same amount of time at different
parts of the orbit. As you can see from the diagram, areas x and y are the
same. The reason for this, put simply, is that the objects move more quickly
when they are closer to the Sun and more slowly when they are further away.
89

figs/chapt-13/kepler-2.png

Figure 13.2: Kepler’s Second Law

Third Law

This will be described in more detail below.

(b) recognise and use F = − Gmr12m2

“The gravitational force between two objects is proportional to the product of


their masses and inversely proportional to the square of the distance between
their centres.”

That is a lot of words and it is much more easily explained with an equation:

Gm1 m2
F =−
r2

Where: F is the force, measured in newtons (N)


90 CHAPTER 13. GRAVITATION

G is the universal gravitational constant, which has a value of 6.67x10-11


Nm2 kg-2

m1 and m2 are the two masses measured in kilograms (kg)

r is the distance between their centres, measured in metres (m).

figs/chapt-13/masses.png

Figure 13.3: Two masses

Important things to note:

1. The force is negative. This is because it is always attractive and attrac-


tive forces are always negative, by definition.

2. The force exerted by mass 1 on mass 2 is the same magnitude but


opposite in direction to the force exerted by mass 2 on mass 1. This is a
direct consequence of Newton’s third law. In most cases that we study
the effect of the force on the smaller mass is much greater than it is on
the larger mass and we ignore the effects of the force on the larger mass.

(c) use Newton’s law of gravity and centripetal force to derive r3 ∝ T 2 for a
circular orbit

Now that we know the size of the force acting on a body moving in circular
motion due to the gravitational force acting on it, we can prove Kepler’s third
law:

For a body orbiting another body, the centripetal force is provided by the
gravitational force.

Gm1 m2
F =− (1)
r2

As the body is moving in circular motion, the gravitational force causes the
body to accelerate towards the centre of the circle, as in the diagram:
91

m2 m1
F

Figure 13.4: Two masses and a force

Thus we apply F=ma, with F being the gravitational force, as on the diagram,
and the acceleration being equal to rω 2 .

F = ma

So
m1 m2
G = m1 rω 2
r2

m1 is the object that is orbiting, so it is the mass of m1 that is feeling the


acceleration and thus m1 goes into the right hand side of the equation.

Thus we can cancel m1 and collect the terms in r to give:

Gm2 = r3 ω 2 (2)

But we know from our revision of circular motion that the period of the orbit
is related to the angular velocity by:


ω= (3)
T

Therefore we substitute equation (3) into equation (2) to give:


92 CHAPTER 13. GRAVITATION

3
4π 2 r
Gm2 =
T2

Now re-arrange to make r the subject of the formula:

Gm2 2
r3 = T (4)
4π 2

G, m2 and π are all constants, so we can finally write:

r3 ∝ T 2

You need to be able to do this for your examination, so make sure that you
learn this proof.

(d) understand energy transfer by analysis of the area under a gravitational


force-distance graph

If you apply a force through a distance, you do work on it. This is a definition.

At GCSE, you learnt that it was given by the equation:

Work done = Force x distance

Later you learned that in fact it was the area under the Force-distance graph.

(Of course, it’s a little bit more difficult than that. It is actually given by:

Z ∞
W ork done = F dx (8)
r

You use this integral in other parts of the specification, but not this part.)

Therefore if we want to know the gravitational potential energy gained or


lost by an object in a gravitational field, we look at the area under the force-
distance graph.

The zero of gravitational potential energy is taken as being at infinity, which


makes sense. If the object isn’t in a field, then it isn’t experiencing any force,
so it doesn’t have any GPE.

This does mean, however, that all gravitational potential energies are negative
as they lose GPE as they fall towards a mass.
93

figs/chapt-13/area-energy.png

Figure 13.5: Graviational PE as area of a graph

For example, if you want to know the GPE gained/lost by an object as it


moves from point R1 to point R2 in the field you look at the area under the
force-distance graph (see figure 13.5)

(Generally, we tend to look at the gravitational potential rather than the GPE,
and use the field strength-distance graph, but the specification asks for GPE
and a force-distance graph, so that is what we are looking at!)

(e) derive and use g = − Gm


r2
for the magnitude of the gravitational field strength
due to a point mass

If you look back to chapter 2 on gravitational fields, you will know the defini-
tion of the gravitational field strength. It is given by the equation:

F
g= (5)
m

But we now know how the force is provided by a spherical mass from Newton’s
94 CHAPTER 13. GRAVITATION

Law of Gravitation. It is given by equation (1) as:

Gm1 m2
F = (6)
r2

So we can substitute equation (6) into equation (5) to give us:

Gm
g= (7)
r2

This is the gravitational field strength due to a spherical mass m at a distance


r from its centre and you need to be able to derive this equation.

A graph of the field strength due to a spherical body against distance looks
like this:

figs/chapt-13/g-r.png

Figure 13.6: Field strength due to a sphere

From the centre of the object up to its radius (0 to R) the variation with field
strength inside the body varies linearly if the density is constant. R is the
95

radius of the body, so therefore the distance from the centre. After R it drops
off as 1/r2 .

(f) recall similarities and differences between electric and gravitational fields

figs/chapt-13/table.png

Figure 13.7: Similarities and Differences

(g) recognise and use the equation for gravitational potential energy for point
masses E = − Gmr1 m2

There is an equation for the GPE which you need to be able to use. It is
found from integrating the expression for the force as mentioned earlier, and
it is given by:

Gm1 m2
E=− (9)
r
96 CHAPTER 13. GRAVITATION

(h) calculate escape velocity using the ideas of gravitational potential energy
(or area under a force-distance graph) and energy transfer

We can now use these methods of working out the GPE gained by an object
in a gravitational field to calculate a quantity called the escape velocity.

NB. A lot of people get escape velocity wrong. It is the velocity needed to be
given to an object on the surface of the planet in order for it to escape the
gravitational field of the planet and have zero KE at that time. Once it has
been given this velocity (by, for example, a cannon) no more energy is put
into the system. From that moment on it is a projectile and is constantly
losing KE as it gains GPE.

Of course, this means that when it finally escapes the gravitational field it has
no energy at all, which also means that it has zero overall energy to start with
as well!

Therefore we use the law of conservation of energy to work out what the escape
velocity must be:

Total energy before = Total energy after = 0

Therefore Initial KE + Initial GPE = 0

1 GMm
mv 2 − =0 (10)
2 e R

Where M is the mass of the planet in kg

R is the radius of the planet in m

m is the mass of the projectile in kg

ve is the escape velocity in ms-1

You can cancel the mass of the projectile and re-arrange for ve from equation
(10) to give:

s
2GM
ve = (11)
R

This is the escape velocity and you can work it out for the Earth. You should
get about 12 kms-1 .
97

There is a graphical way of looking at this as well. The GPE gained by the
body as it moves from the surface of the planet, radius R, is given by the area
under the force-distance graph from the surface of the planet to infinity (i.e.
as in Figure 13.5, but with R2 at infinity). If it is launched from the surface
of the planet then that also equals its initial KE.

(i) calculate the distance from the centre of the Earth and the height above its
surface required for a geostationary orbit.

A geostationary orbit is one which stays above the same position on the Earth’s
equator at all times.

This means, therefore, that it has a period of 24 hours.

It is therefore possible to calculate, using equation (4), r for a geostationary


orbit.

N.B. A very common mistake is to say that r is the height of the orbit. This
isn’t the case – it is the radius of the orbit, so it is the distance of the satellite
from the centre of the Earth, not the surface of the Earth.

You should make sure that you can do this. Have a go at working out r, given
the following data:

G = 6.67 x 10-11 Nm2 kg-2

ME = 5.98 x 1024 kg

and remembering that T = 24 hours (but don’t forget to convert to seconds!)

You should have got an answer of 4.23 x 107 m.

If I now tell you that the radius of the Earth is 6.36 x 106 m, you can also
write down the height of the satellite above the surface of the Earth, and this
comes to 3.6 x 106 m.

So a geostationary satellite orbits at a height that is about 6 times greater


than the radius of the Earth.
14 Electromagnetism

Candidates should be able to:

(a) understand and use the terms magnetic flux density, flux and flux linkage

Electromagnetic induction depends crucially on the concept of flux. This is the


product of the magnetic field strength B and the area of a loop perpendicular
to the field, A.
Φ = BA
The unit of flux is the weber, Wb. Given its relation to flux, the field strength
B can also be referred to as the magnetic flux density. It also follows that the
tesla is equivalent to one weber per metre-squared.

When a coil of wire encloses an area of flux we can calculate the flux-linkage
which is the product of the flux through the coil and the number of turns on
the coil, N Φ.

Magnetic flux density can be defined as the force per unit of current in a wire
of unit length in a magnetic field.

(b) understand that magnetic fields are created by electric current

An electric current in a wire creates a magnetic field. The field strength


depends on the size of the current and the distance from the wire. In the case
of a solenoid or electromagnet the field strength is also proportional to the
number of turns on the coil.

(c) recognise and use F = BIl sin θ

This is the equation for the force on a current-carrying wire of length l with
a current of I at an angle θ to a magnetic field of strength B.

99
100 CHAPTER 14. ELECTROMAGNETISM

(d) recognise and use F = BQv sin θ

This is the equation for a particle of charge Q moving at a velocity v at an


angle θ to a magnetic field of strength B. Note that this is sometimes called
the Lorentz force.

(e) use Fleming’s left-hand rule to solve problems

Fleming’s left-hand rule allows the calculation of the direction of the force on
either a charged particle or a current-carrying wire. The geometry is shown
below.

thumb:
force

first finger:
field, B

second finger:
current, I

Figure 14.1: Fleming’s left-hand rule

It is important to note that if you are using the left-hand rule to predict the
direction of force on a moving negatively charged particle then the current is
flowing in the opposite direction to the velocity of the particle.

Note that vector form of the equations explains the left-hand rule and sine
terms as follows:
F = qv × B
F = IL × B
101

(f) explain qualitatively the factors affecting the emf induced across a coil when
there is relative motion between the coil and a permanent magnet or when there
is a change of current in a primary coil linked with it

Faraday’s Law: The induced emf is proportional to the rate of change


of flux linked.

Since the emf depends on the rate of change of flux linkage the size of the emf
is proportional to the number of turns on a coil. The rate of change of flux is
determined by strength of the magnet and how quickly it is moving through
the coil.

(g) recognise and use E = − d(N


dt
Φ)
and explain how it is an expression of
Faraday’s and Lenz’s laws

Since N Φ is the flux linkage then d(N


dt
Φ)
is the rate of change of flux linkage.
Thus Faraday’s law is expressed in this equation.

Lenz’s Law: The direction of the induced emf is such that the current it
causes to flow opposes the change which is producing it.

The negative sign indicates that the induced emf acts to oppose the change in
flux linkage which created it. This may occur by the magnetic field opposing
the motion of a magnet of coil or by the magnetic field opposing a stationary
change in field.

Note that Lenz’s law is a result of conservation of energy and applying it to a


specific situation often requires a discussion of work and energy transfers.

(h) derive, recall and use r = mv


BQ for the radius of curvature of a deflected
charged particle

When a charged particle moves through a uniform magnetic field at a constant


speed it experiences a constant force at right-angles to its velocity. This results
in it moving in circular motion with the Lorentz force providing the centripetal
force.
mv 2
BQv =
r
mv
=⇒ r =
BQ
102 CHAPTER 14. ELECTROMAGNETISM

This allows the calculation of the mass to charge ratio of the particle providing
its velocity can be determined (e.g. from an accelerating potential).
103

(i) explain the Hall effect, and derive and use V = Bvd

When an electron moves through a magnetic field it will experience a force at


right-angles to its motion. If a current flows through a thin film of material
then a negative charge will accumulate on one edge of the film. Figure 14.2
shows this force acting and charge building up.

+ + + + + + + + +

e− v

– – – – – – – – –
Magnetic field into page

Figure 14.2: The Hall Effect

Over time the charges build up on either side of the wager until equilibrium
is reached. At this point the the force due to the electric field on the charge
equals the force due to the Lorentz force on the charge.
V
q = Bqv
d
Vhall = Bvd

This relationship is the basis of a Hall Probe which is used to measure mag-
netic fields. In these circumstances the velocity v is the drift velocity of the
electrons.
15 Special Relativity

This chapter contains revision on the topics of:

• Einstein’s special principle of relativity

• Time dilation

• Length contraction

Candidates should be able to:

(a) recall that Maxwell’s equations describe the electromagnetic field and pre-
dict the existence of electromagnetic waves that travel at the speed of light
(Maxwell’s equations are not required)

Maxwell related electricity and magnetism in a mathematically elegant way,


and in doing so predicted that the speed of electromagnetic waves was:

c= √1
0 µ0 = 3 x 108 ms−1

(b) *recall that analogies with mechanical wave motion led most physicists to
assume that electromagnetic waves must be vibrations in an electromagnetic
medium (the aether) filling absolute space

Just what was it that was oscillating? Most physicists thought there must be
a medium...

(c) *recall that experiments to measure variations in the speed of light caused
by the Earth’s motion through the ether gave null results

...However experiments such as the Michelson-Morley experiment1 showed


that there was no difference in the measurement of the speed of light over
1
http://scienceworld.wolfram.com/physics/Michelson-MorleyExperiment.html

105
106 CHAPTER 15. SPECIAL RELATIVITY

the course of a year - if the Earth was travelling through an aether you would
expect the discrepancy to be measurable. But crucially no differences were
found.

(d) *understand that Einstein’s theory of special relativity dispensed with the
aether and postulated that the speed of light is a universal constant

Einstein put forward the idea that there was no medium, and the speed of
light was fixed.

(e) *state the postulates of Einstein’s special principle of relativity

His two postulates were

• The laws of physics are the same in all inertial rest frames

• The speed of light in free space is a fixed value

From these two postulates, it is possible to deduce the idea that time is not
fixed for all observers, and that distances measured by two observers moving
relative to each other may be different! A whole new avenue of physics was
opened.

(f) *explain how Einstein’s postulates lead to the idea of time dilation and
length contraction that undermines the idea of absolute time and space

Light clock

Imagine a beam of light bouncing up and down between two mirrors, spaced
a distance of 3 x 108 m apart

1 second between bounces


107

Now imagine that you are moving right-to-left relative to the ’light clock’ (or
that the light clock is moving left-to-right relative to you - it doesn’t matter...)

If the speed of light is constant for all observers, as Einstein’s postulates


require it will take longer between bounces, because it has to travel further
due to the extra horizontal component:

direction of motion

>1 second between bounces

q
v2
(g) *recognise and use t0 = q t
2
and l = l 1 − c2
1− v2
c

Example Question
Cosmic radiation constantly bombards the Earth from outer space.
The majority of these cosmic rays are protons which, when they hit
the Earth’s upper atmosphere, create sub-atomic particles called pions,
which then quickly decay into muons. Suppose a cosmic ray proton
hits a molecule of nitrogen in the upper atmosphere at a distance of 50
km above the Earth’s surface. If the pion produced has a velocity of
0.9999c and has an average lifetime (in its own time) of t = 2.60×10–8
s. How far would the pion travel before decaying taking into the effects
of time dilation?
Answer
First we must calculate the time it takes to decay according to a
stationary observer. We do this by calculating the Lorentz factor γ
and multiplying this by the lifetime in its own frame of reference:
γ = q 1 2 = 70.7
1− v2
c

so time taken to decay = 70.7 x 2.60 x 10−8 = 1.83 x 10−6


seconds
Then distance travelled = speed x time = 550 m
108 CHAPTER 15. SPECIAL RELATIVITY

(h) *understand that two events which are simultaneous in one frame of ref-
erence may not be simultaneous in another; explain this in terms of the fun-
damental postulates of relativity and distinguish this from the phenomenon of
time dilation

Imagine a situation in which a light is turned on in the middle of a train


carriage, and light rays travel out in both forward and backwards directions.
From the reference frame of a person riding along with the train, the light

reaches each end of the carriage at the same time. Those events appear to be
simultaneous.
In the reference frame of a stationary observer, the train is moving forward
109

(to the right in this diagram), but the light rays travel at c, so the light ray
travelling backwards will reach the rear end of the carriage first.

In the reference frame of somebody travelling faster than the train, the train
appears to be going backwards, and the forward travelling light ray will reach
the front end first.
The idea of two events being simultaneous is therefore impossible - it depends
on your reference frame.

This effect is a completely different effect to time dilation. The effect we are
dealing with here is called loss of simultaneity.
16 Molecular Kinetic Theory

Content

• absolute scale of temperature

• equation of state

• kinetic theory of gases

• kinetic energy of a molecule

• first law of thermodynamics

• entropy

• second law of thermodynamics

Candidates should be able to:

(a) explain how empirical evidence leads to the gas laws and to the idea of an
absolute scale of temperature

In the seventeenth century Robert Boyle began making quantitative measure-


ments on gases at different pressures. By observing the results of varying the
volumes of gases on their pressures he postulated what has become known as
Boyles’ Law.

The product of the pressure and volume of a sample of gas at constant


termperature remains constant.

Developing a quantitative scale for temperature was a difficult process1 and


it was not until the end of the eighteenth century that a relationship between
1
see Hasok Chang’s Inventing Temperature for more details

111
112 CHAPTER 16. MOLECULAR KINETIC THEORY

pressure and temperature was discovered. Pressure was found to be linearly


related to temperature as measured in the celsius scale. It was found that all
samples of gases had the same intercept with the x-axis and this lead to the
idea of absolute zero, the temperature at which the pressure of the gas would
be zero.
P

x xx
x xx

T ◦C

Absolute Zero

Figure 16.1: Pressure against Temperature of a Gas

The temperature of Absolute Zero was defined as zero kelvin, 0 K with this
being equal to −273 ◦C. This scale is known as the absolute temperature scale.

If the absolute temperature scale is used, then two more empirical laws can
be stated. The first is Charles’ Law:

For gas at a fixed pressure, the volume of a sample of gas is proportional to


its absolute temperature.

The second is Gay-Lussac’s law:

For a fixed volume of gas, the pressure is proportional to the absolute


temperature.

(b) use the units kelvin and degrees Celsius and convert from one to the other

Conversion between kevin and degrees celsius is simply a matter of adding or


subtracting 273 as the two scale share the same size of unit:
0 K = −273 ◦C
113

(c) recognise and use the Avogadro number NA = 6.02 × 1023 mol−1

The Avogadro number is the number of particles (atoms or molecules) which


are present in one mole of the substance. It can also be thought of as the
constant of proportionality between the mass of the particle and the mass of
one mole of the substance.

For example, using carbon-12:


0.012 kg
N= = 6.0302 × 1023
1.99 × 10−26 kg

The number of moles of a substance can be found by:

1. dividing the number of particles by the Avogadro number;

2. dividing the mass of the sample by the molar mass.

(d) recall and use pV = nRT as the equation of state for an ideal gas

This equation includes the following quantities:

Symbol Quantity Standard Unit


p pressure Pa
V volume m3
n no. of moles of the gas N/A
R the molar gas constant J mol−1 K−1
T the absolute temperature K

R, the molar gas constant, has a value of 8.314 J mol−1 K−1 and is equal to
NA k.

(e) describe Brownian motion and explain it in terms of the particle model of
matter

When a small, visible particle such as a pollen grain or smoke particle is


observed under a microscope it is seen to move around in an erratic, random
manner. This is explained by the fact that it is being constantly bombarded
by air molecules whose effects do not quite cancel out. Since the pollen grain
or smoke particle is much more massive than the air molecules it follows that
these must be moving very rapidly. In 1905 Albert Einsein published a detailed
statistical treatment of Brownian motion using the theory of atoms and thus
Brownian motion provides very strong evidence for the atomic hypothesis.
114 CHAPTER 16. MOLECULAR KINETIC THEORY

(f) understand that the kinetic theory model is based on the assumptions that
the particles occupy no volume, that all collisions are elastic, and that there
are no forces between particles until they collide

Notes on these assumptions:

1. 1 m3 of air contains approximately

PV 101 kPa1 m3
N= = = 2.5 × 1025
KT 1.38 × 10−23 J K−1 298 K
particles. If these particles are modeled as spheres with a diameter of
300 pm then the particles take up a fraction of the total volume of around
10−27 .

2. Elastic collisions mean that the kinetic energy of the particles is not lost.

3. The particles travel in straight lines between collisions. Forces between


the particles would cause them to clump together. The lack of forces
between particles also means that all the energy is in the form of kinetic
energy of the particles.

(g) understand that a model will begin to break down when the assumptions on
which it is based are no longer valid, and explain why this applies to kinetic
theory at very high pressures or very high or very low temperatures

High pressures At high pressures particles will be forces together. This


means that the first assumption about negligible volume may break down
as the gaps between particles decreases. In addition, the gas may get
close to the point of condensation and particles may begin to attract
each other.

Very low temperatures Similarly to high pressures, at low temperatures


the molecules may be very close to one another and may begin to con-
dense, implying forces between particles.

Very high temperatures At very high temperatures atoms become a plasma,


i.e. separate into positive ions and free electrons. Under such circum-
stances forces exist between the particles.
115

(h) derive P V = 13 N mhc2 i from first principles to illustrate how the micro-
scopic particle model can account for macroscopic observations

This theory assumes that pressure is caused by the averaging the many elastic
collisions of a number of particles with the walls of the container.

vx

Figure 16.2: A particle in a box

When the particle, i, collides with the right-hand wall of the box it rebounds
with the same y-velocity and a negative x-velocity. The change in momentum
of the particle is therefore ∆p = 2mvx , where m is the mass of the particle.
This is equal to the impulse delivered on the wall during the particle collision.

Impulse = 2mvx

The particle will now travel to the left-hand side of the box and back. The
time taken to do this is equal to 2d/vx . We can therefore think of the impulse
due to a single collision being averaged over this period of time. If this is the
case then the force due to an individual particle is given as:

2mvx mvx 2
fi = =
2d/vx d
116 CHAPTER 16. MOLECULAR KINETIC THEORY

If we take a system of N particles and replace the properties of our particle


with the average properties of the particles in the system we get:
N mhvx 2 i
F = fi =
X

i
d

Note that this includes the expression hvx 2 i - the mean of the squared x-
velocity. The order here is important and the mean velocity of the particles
is zero. The square of a component of velocity is related to the velocity by
pythagoras:
c2 = vx 2 + vy 2 + vz 2
Now we are using many particles we can make the assumption that a particle
is equally likely to be travelling in any direction, therefore
hvx 2 i = hvy 2 i = hvy 2 i
giving
1
hvx 2 i = hc2 i
3
Substitution into the equation for F gives
1 2
3 N mhc i
F =
d
and
1
F N mhc2 i
P = = 3
A Ad
We now note that Ad is equal to the volume of the box and re-arrange to give
1
P V = N mhc2 i
3

(i) recognise and use 12 mhc2 i = 23 kT

The formula for P V derived above can be equated with the formula from the
empirical gas laws (P V = nRT ) to give:
1
N mhc2 i = nRT
3
Since n = N/NA and R = NA k the right-hand side becomes N kT . This is
usually re-arranged to give
1 3
mhc2 i = kT
2 2
as now the left-hand side represents the average kinetic energy of a molecule
in the gas. Since this is an ideal gas and has no potential energies this is
also the total energy per molecule. Hence the link between the macroscopic
quantity of temperature and the microscopic energy per molecule is arrived
at.
117

(j) understand and calculate the root mean square speed for particles in a gas

The root mean square (RMS) is usually calculated by re-arranging the equa-
tion above and square-rooting:

s
q 3kT
hc2 i =
m

(k) understand the concept of internal energy as the sum of potential and
kinetic energies of the molecules

In an ideal gas the internal energy is equal to the sum of kinetic energies of
the molecules, i.e.

1 3
U = N mhc2 i = N kT
2 2

(l) recall and use the first law of thermodynamics expressed in terms of the
change in internal energy, the heating of the system and the work done on the
system

This can be written as:


∆U = ∆Q + ∆W
Note that in some textbooks the first law is written in terms of the work done by
the system, giving ∆W a different sign. If you think in terms of conservation
of energy in the particular example you will be alright.

The Carnot Cycle shown in Firgure 16.3 is commonly used to test the under-
standing for the first law. The cycle begins and ends at point A. This means
that the internal energy of the system at the start and end of the cycle is
the same (as P V = nRT and T is proportional to the internal energy). The
changes at each stage of the cycle are as follows:

1. Isothermal Expansion In stage 1 the gas expands at contant tempera-


ture. Since it is expanding, the gas is doing work on the surroundings
and since it is at constant temperature the internal energy of the gas
remains constant. Therefore the second law implies that the gas must
be absorbing heat. Note that isothermal changes are often indicated by
describing the change as occuring slowly.
2. Adiabatic Expansion In stage 2 the gas expands without transferring
heat to/from the surroundings. Since it is doing work the gas’s internal
energy drops, as does its temperature.
118 CHAPTER 16. MOLECULAR KINETIC THEORY

P
A

Figure 16.3: The Carnot Cycle

3. Isothermal Compression In stage 3 the gas is compressed at constant


temperature. Since work is done on the gas and the internal energy
remains constant it must be the case that heat is lost to the surroundings.

4. Adiabatic Compression Finally, the gas is compressed without heat


transfer to/from the surroundings. Work is done on the gas and there-
fore the internal energy, and termperature, increase.
119

Example Question
Complete the table for the Carnot Cycle shown in figure 16.3.
stage thermal energy work done on the increase in
supplied to the gas / J internal energy of
gas / J the gas / J
1 (a) -936 (b)
2 0 (c) (d)
3 (e) +702 0
4 (f) +844 +844
Answer
In stage 1 there is no change in internal energy of the gas, so (b) equals
zero. In order to satisfy the first law of thermodynamics the thermal
energy supplied must equal the work done by the gas so (a) equals
936 J.
As the total change in internal energy throughout the cycle must be
zero, we can now calculate (d) as being −844 J.
(c) can be calculated using the first law as −844 J.
A similar argument to that used in stage 1 can be used in stage 3 to
give (e) as −702 J.
In stage 4 no energy is transferred by heating therefore (f) is equal to
zero.
The final table is therefore:
stage thermal energy work done on the increase in
supplied to the gas / J internal energy of
gas / J the gas / J
1 +936 -936 0
2 0 -844 -844
3 -702 +702 0
4 0 +844 +844

(m) recognise and use W = p∆V for the work done on or by a gas

This follows from the definition of work done. If the cross-sectional area is a
contant A, then derivation is as follows:

W = F ∆s = P A × ∆s = P ∆V

(n) understand qualitatively how the random distribution of energies leads to


E
the Boltzmann factor e− kT as a measure of the chance of a high energy

When a large number of particles share a fixed amount of energy between


them and are able to transfer energy through random collisions, it turns out
that the most likely distribution of energies follows a specific distribution - the
Boltzmann distribution. One of the features of this distribution is that the
120 CHAPTER 16. MOLECULAR KINETIC THEORY

fraction of particles with an energy greater than a certain amount of energy,


E, is given by the Boltzmann factor, i.e.
NE E
= e− kT
Ntotal

(o) apply the Boltzmann factor to activation processes including rate of reac-
tion, current in a semiconductor and creep in a polymer

From the point above, it follows that the rate of any process which depends
on an activation energy is likely to be proportional to the Boltzmann factor.

Process Depends on Energy E represents


rate of reaction the number of particles the activation energy of
with an energy above the process
the activation energy
current in a semi- the number of free the energy required
conductor charge carriers for a charge carrier to
move the the conduc-
tion band, i.e. the band
gap
creep in a poly- the ability of the poly- the energy required to
mer mer chains to break break the intermolecu-
the weak intermolecular lar bonds between poly-
forces between them mer chains

(p) *describe entropy qualitatively in terms of the dispersal of energy or par-


ticles and realise that entropy is related to the number of ways in which a
particular macroscopic state can be realised

A common way of describing the concept of entropy is to focus on the idea of


disorder. When discussing entropy scientifically it is important to be precise
about what this means. A more precise way of characterising entropy is to
think about the number of ways in which a macroscopic state can be realised.
For example, imagine seven particles in a quantised system with four units of
energy to share between them. Each state can be denoted using the notation
S = {N0 , N1 , N2 , . . .} where N0 is the number of particles with this much
energy.

For example, one particle could have all four units of energy and the rest have
zero. This would be denoted {6, 0, 0, 0, 1}. There are seven different ways in
which this macrostate could come about. The table below shows the different
possible macrostates and the number of ways these could be attained.
121

State No. of microstates


{6, 0, 0, 0, 1} 7
{5, 1, 0, 1} 7 × 6 = 42
{4, 2, 1} 7 × 6 × 5/2 = 105
{3, 4} T ODO

Table 16.1: Macrostates and microstates

As you can see, there are far more ways to achieve the macrostate {4, 2, 1}
than the others. This makes it the state with the highest entropy.

(q) *recall that the second law of thermodynamics states that the entropy of an
isolated system cannot decrease and appreciate that this is related to probability

If the energy available to a system, or a distribution of particles, is dispersed


through random process then over time the most probable configuration will
dominate. As the highest entropy macrostate is also the most probable one it
is highly likely that the entropy will increase with time.

As an example, consider 100 coins which are all heads-up. There is only one
way to arrange for this particular macrostate to occur. There are 1.27 × 1030
possible microstates available and therefore if the coins are tossed it is over-
whelmingly likely that the disorder in the system is going to increase. If this
idea is multiplied up to the billions of particles in even the smallest sample
of matter it is seen that all systems will move towards their highest entropy
state.

The only way to return to the original state of the coins would be to inter-
vene and sort them individually one-by-one, thus breaking the isolation of the
system.

(r) *understand that the second law provides a thermodynamic arrow of time
that distinguishes the future (higher entropy) from the past (lower entropy)

The idea here is that nature tends towards higher entropy states. For ex-
ample, consider the two distributions of gas particles in figure 16.4. In this
case it is overwhelmingly likely that B occured after A. Thus, the second law
of theormodynamics provides an arrow of time despite the fact that individ-
ual particles may be governed by Newton’s laws of motion which are fully
reversible and do not give any information about the direction of time.
122 CHAPTER 16. MOLECULAR KINETIC THEORY

A B

Figure 16.4: Two distributions of particles

(s) *understand that systems in which entropy decreases (e.g. humans) are
not isolated and that when their interactions with the environment are taken
into account their net effect is to increase the entropy of the Universe

Living organisms including humans take in relatively simple molecules and


create highly complex, ordered structures from them. This appears to violate
the second law of thermodynamics and entropy appears to be decreasing.
However, an individual part of a system can have a net increase in entropy if
it is not isolated. In such a scenario the overall entropy of the universe (the
ultimate isolated system) must still decrease. In the case of life on earth, the
whole biosphere can be seen as a complex heat engine extracting work from
the heat energy input from the sun whilst radiating heat into the coldness of
space.

(t) *understand that the second law implies that the Universe started in a state
of low entropy and that some physicists think that this implies it was in a state
of extremely low probability.

If the universe is seen as the ultimate isolated system then it follows that the
second law of thermodynamics should apply. In this case the universe must
be moving from a low entropy state to a high entropy. In isolated systems,
low entropy states are also highly improbably and this leads to the idea that
our universe’s initial state is a highly improbable one. However, this requires
us to accept that we can extend the idea of probability to the initial state of
the universe and that we are not committing a ‘category error’ by considering
the universe as a thermodynamic system in this way.
17 Nuclear Physics

Content

• equations of radioactive decay

• mass excess and nuclear binding energy

• antimatter

• the standard model

Candidates should be able to:

(a) show that the random nature of radioactive decay leads to the differential
equation
dN
= −λN (17.1)
dt
and that
N = N0 e−λt (17.2)
is a solution to this equation.

Radioactive decay is characterised by the fact that the number of nuclei which
disintegrate per unit time is directly proportional to the number of unchanged
nuclei remaining. Since a disintegrating nucleus reduces the number remain-
ing, there is a negative sign in the proportionality. This relationship can be
expressed mathematically as equation 17.1. Where N is the number of nuclei
and λ is called the decay constant, with units s−1 .

Equation 17.1 is a differential equation with respect to time and therefore


the solution to it is a function of time. We can show that equation 17.2 is a
solution to this equation by differentiating it.

123
124 CHAPTER 17. NUCLEAR PHYSICS

N = N0 e−λt
dN
= (−λ) N0 e−λt
dt
= −λN

(b) recall that activity


dN
A=− (17.3)
dt
and show that A = λN and A = A0 e−λt

Every time a radioactive nucleus disintegrates it emits a particle of ionising


radiation. Thus the activity is simply the negative of the rate of change of
the number of unchanged nuclei remaining. Simple substitutions allow the
derivation of the following equations.

dN
A=− N = N0 e−λt
dt
= −(−λN ) −λN = −λN0 e−λt
= λN A = A0 e−λt

Note that we do not measure the true activity as that would mean detecting
all of the radiation given off by the sample. However, we assume that the
measured activity is proportional to the true activity and therefore all our
measurements behave in the same way.

(c) show that the half-life


ln 2
t1 =
2 λ

The half-life is defined as the time take for half of the nuclei to decay. Therefore
I can substitute N = N20 into equation 17.1 to give:

1 −λt 1
=e 2
2
1
ln = −λt 1
2 2

ln 2
t1 =
2 λ
125

(d) use the equations in (a), (b) and (c) to solve problems

(e) recognise and use the equation


I = I0 e−µx (17.4)
as applied to attenuation losses

When a wave or ionising radiation travels through a medium its amplitude


will reduce due to attenuation. This is due to scattering and/or absorption by
the medium. Note that is this different to a reduction in intensity due to the
radiation spreading out. It is assumed that if a given fraction of the intensity
is absorbed in a unit length then eqaution 17.1 can be used, replacing N with
intensity, t with distance, x and the constant of proportionality with µ. This
gives
dI
= −µx (17.5)
dx
Using a similar logic to that for equation 17.2 we can show that equation 17.4
is a solution to this equation.

(f) recall that radiation emitted from a point source and travelling through
a non-absorbing material obeys an inverse square law and use this to solve
problems

Figure 17.1: Radial radiation

Radiation emitted from a point source spreads out symmetrically in all direc-
tions as shown in figure 17.1. The intensity is defined as the power per unit
area of radiation. A distance r from the centre of the source the radiation is
spread over the surface of a sphere and is therefore calculated using
P
I= (17.6)
4πr2

Since I ∝ r−2 this is known as an inverse-square law.


126 CHAPTER 17. NUCLEAR PHYSICS

(g) estimate the size of a nucleus from the distance of closest approach of a
charged particle

When particles approach a nucleus head-on they begin with kinetic energy EK .
All of this kinetic energy is converted to electrical potential energy, EP E.
If the energy of the alpha particles is known then a distance of minimum
separation can be calculated.

Example Question
Calculate the minimum distance of separation of a alpha particle of
kinetic energy 4.0 MeV travelling directly towards a gold nucleus (Z =
79).
Answer
Equating the initial kinetic energy to the electrostatic potential gives
Q1 Q2
Ek =
4π0 d
therefore
Q1 Q2
d=
4π0 Ek
2 × 79 × (1.6 × 10−19 )2
=
4π0 × 4.0 × 106 × 1.6 × 10−19
= 5.7 × 10−14 m

It turns out that if high energy alpha particles are fired at the nucleus then
Rutherford’s deflection formulae break down and the distribution of alpha
particles no longer fits the expectation. Under these circumstances it can be
assumed the alpha particle is interacting with the nucleus and therefore has
been able to approach to within the nuclear radius.
127

(h) understand the concept of nuclear binding energy, and recognise and use
the equation ∆E = c2 ∆m (binding energy will be taken to be positive)

It turns out that the mass of a nucleus is always smaller than the total mass
of the constituent protons and neutrons. This difference is called the mass
deficit and can be converted to an energy using ∆E = c2 ∆m. The reduction
in mass corresponds to the energy released by combining the nucleons.

Example Question
Calculate the binding energy of a helium-4 nucleus of mass 4.0015 u
Answer
The helium-4 nucleus is composed of two protons and two neutrons.
Their total mass is

2 × 1.007 28 u + 2 × 1.008 67 u = 4.0319 u

The mass deficit, ∆m, is given by

∆m = 4.0319 u − 4.002 15 u = 0.0304 u

Therefore the binding energy is calculated as

0.0304 × 1.66 × 10−27 × c2 = 4.54 × 10−12 J = 28.4 MeV


128 CHAPTER 17. NUCLEAR PHYSICS

(i) recall, understand and explain the curve of binding energy per nucleon
against nucleon number

A useful measure of the stability of the nucleus is given by its binding energy
per nucleon. This is the binding energy divided by the nucleon number. A plot
of the binding energy per nucleon against nucleon number is show in figure
17.2. The nucleus with the largest binding energy is iron-56 and this therefore
is the most stable nucleus. Nuclei with nucleon numbers below iron-56 will
release energy when they undergo nuclear fusion and those above iron-56 will
release energy when they undergo fission.

9
Binding Energy per Nucleon / MeV

8
7
6
5
4
3
2
1
0
0 40 80 120 160 200 240
Nucleon Number

Figure 17.2: Variation of binding energy per nucleon with nucleon number

(j) recall that antiparticles have the same mass but opposite charge and spin
to their corresponding particles

All normal particles have an antiparticle partner with the same mass, but
some properties which are opposite including electrical charge.

(k) relate the equation ∆E = c2 ∆m to the creation or annihilation of particle-


antiparticle pairs

A particle-antiparticle pair can be created whenever there is enough energy


present to do so. For example, if a photon of light is near an atomic nucleus
it can spontaneously convert into an electron-positron pair.
129

Example Question
Calculate the minimum energy a photon must have in order to create
an electron-positron pair.
Answer

E = c2 ∆m = c2 × 2 × 9.11 × 10−31 = 1.02 MeV

(l) recall the quark model of the proton (uud) and the neutron (udd)

The theory of quarks was developed to explain the large number of particles
discovered in the early particle colliders. Particles made of quarks are called
hadrons. Normal matter is made up of two types of quark, the up quark
(charge + 23 e) and the down quark (charge − 31 e). From the charges it is possible
to see that the proton must be uud and the neutron udd.

(m) understand how the conservation laws for energy, momentum and charge
in beta-minus decay were used to predict the existence and properties of the
antineutrino

The existence of the antineutrino was first predicted from the energy spectrum
of beta decay. Beta particles are produced when a neutron in the nucleus is
converted into a proton and a high energy electron. These electrons leave the
nucleus at high speed and are detected as beta particles. In such a scenario
(a two body process) the electrons should have a fixed amount of energy due
to the conservation of momentum. However, the electron was found to have
a range of energies. The explanation provided was that whenever an electron
was emitted with little energy a third particle has carried away a lot of energy
(and vice-versa). The particle had to be neutral (to conserve charge) and of
very small mass and was named the neutrino.

The full beta decay equation now becomes

n → p + e− + ν e

Further evidence for the existence of this third particle comes from bubble
chamber tracks left by beta decay which show the nucleus and electron both
recoiling away from a particle which does not leave a trace in the bubble
chamber - the neutrino. A photo of this decay can be seen at the science
photo library here: http://www.sciencephoto.com/media/1210/view

(n) balance nuclear transformation equations for alpha, beta-minus and beta-
plus emissions

When these nuclear decays occur total nucleon number remains the same and
charge is conserved. In order to make life easier we give beta-minus particles
130 CHAPTER 17. NUCLEAR PHYSICS

a proton number of −1 and a nucleon number of zero. Beta-plus particles


(positrons) are given a proton number of +1. Alpha particles are helium-4
nucleii (42 He).

As an example, if we know that magnesium-23 decays by beta-plus decay we


can write
12 Mg →Z X +1 e
23 A 0 +

We can deduce the identity of X by conserving nucleon number (A + 0 = 23)


and proton number (Z + 1 = 12). Therefore:

12 Mg 11 Na +1 e
23
→23 0 +

(o) recall that the standard model classifies matter into three families: quarks
(including up and down),leptons (including electrons and neutrinos) and force
carriers (including photons and gluons)

An important feature to note is that quarks and gluons are never observed
on their own. There are further ‘generations’ of quarks and leptons but these
only exist at higher energies.

(p) recall that matter is classified as baryons and leptons and that baryon
numbers and lepton numbers are conserved in nuclear transformations.

Baryons are made up of three quarks and have a baryon number of +1. Simi-
larly, leptons have a lepton number of +1. Anti-baryons and anti-leptons have
respective numbers of -1.

For example, conservation of baryon number prohibits the following:

p + n → p + e+ + e−
B = 1 + 1 6= 1 + 0 + 0

Conservation of lepton number also shows why the anti-electron neutrino is


required in beta decay:

n → p + e− + ν e
L = 0 = 0 + 1 + (−1)
B =1=1+0+0
18 The Quantum Atom

Content

• linespectra

• energy levels in the hydrogen atom

Candidates should be able to:

(a) explain atomic line spectra in terms of photon emission and transitions
between discrete energy levels

When a gas of atoms is given energy (e.g. by an electric field) that energy
is emitted at electromagnetic waves at a few, specific frequencies. These fre-
quencies are the same for every atom of a specific element; however they differ
between elements. At typical line spectrum is shown in Figure

The explanation for this behaviour is the quantisation of energy levels within
the atom. Electrons are only able have occupy certain (discrete) energies
and when they move between these energies they emit (or absorb) photons of
electromagnetic radiation. These photons have different frequencies depending
on the amount of energy they carry away.

(b) apply E = hf to radiation emitted in a transition between energy levels

When an electron in an atom falls from one energy level to a lower one the
excess energy is emitted as a photon. The energy of this photon is equal to
the energy lost.

131
132 CHAPTER 18. THE QUANTUM ATOM

Example Question
An electron in an atom falls from the n = 3 state to the n = 2 state
as shown below. Calculate the wavelength of the photon emitted.

E/eV
−2.61 eV n=3 γ
−5.88 eV n=2

−23.5 eV n=1

Answer
The difference in energy between the two levels is given by

E = (−2.61 eV) − (−5.88 eV) = 3.27 eV

The wavelength can be calculated using


hc
E = hf =
λ
hc hc
λ= = = 379 nm
E 5.24 × 10−19 J

(c) show an understanding of the hydrogen line spectrum, photons and energy
levels as represented by the Lyman, Balmer and Paschen series

The example above shows a single transition between energy levels. In reality
when an atom is excited it will emit photons corresponding to many transitions
at once. The example of the Hydrogen atom is shown in Figure 18.1. These
transitions can be grouped into series based on which energy level the electron
ends up in. Here there are three series shown which correspond to the Lyman
(falling to n = 1), Balmer (falling to n = 2) and Paschen (falling to n = 3)
Series. For simplicity only six energy levels are shown but in reality each series
has potentially infinitely many possible starting states, although most of them
will have very similar energies (as the starting energy approaches zero).
133

E/eV
0 n=∞
n=6
n=5
n=4
Paschen Series (IR)
n=3
Balmer Series (Visible)

n=2

Lyman Series (UV)

n=1

Figure 18.1: Transitions corresponding to the Hydrogen spectrum

(d) recognise and use the energy levels of the hydrogen atom as described by
the empirical equation
−13.6 eV
En = (18.1)
n2

By studying the line spectra of hydrogen equation 18.1 can be determined


from experiment. The equation can be used to determine the wavelengths of
the emission lines of the Hydrogen spectrum:
1 1
 
Eγ = −13.6 eV −
n22 n21

(e) *explain energy levels using the model of standing waves in a rectangular
one-dimensional potential well

The orbital model of electrons in an atom allows electrons to have any energy
we require. However, if one considers the electron to be acting as a standing
wave then the idea of discrete energy levels comes naturally.

The simplest model of an electron as a standing wave is to consider the stand-


ing wave as a linear wave bound at each end. With an electron we define a
‘potential well’, i.e. a region of space in which the electron has zero potential
energy and the rest of space the electron would have infinite potential energy.
The electron is therefore bound within this space.
134 CHAPTER 18. THE QUANTUM ATOM

n=3

n=2

n=1

Figure 18.2: Three standing waves in a potential well

For each of the standing waves described in Figure 18.2 the wavelength can
be calculated using
2L
λn = (18.2)
n

If the wavelength is related to the de Broglie wavelength of the electron then


each of these standing waves can be given an momentum and hence energy.

p2 h2 h2 n2
En = = = (18.3)
2m 2mλ2 8mL2

Equation 18.3 clearly can not match the empirical equation (18.1), but it does
show a dependence on n and discrete energy levels.

(f) *derive the hydrogen atom energy level equation En = −13.6


n2
eV
algebraically
using the model of electron standing waves, the de Broglie relation and the
quantisation of angular momentum.

In order to adapt the above model to fit an atom, the idea of the potential
well was adapted to say that instead of fitting inside a potential well, a whole
number of wavelengths should fit around the circumference of the atom. This
gives a new criterion:
2πr = nλ (18.4)
135

The de Broglie wavelength equation can be substituted into equation 18.4 to


give
nh
mvr = (18.5)

The quantity on the left is the angular momentum. It turns out that our
quantisation rule based on wavelength is equivalent to stating that the angular
momentum is quantised. The value 2π h
is so common in quantum theory that
it has its own symbol, ~.

In fact, equation 18.5 was Bohr’s starting point for his model of the atom.

Now we have a rule for the quantisation we can apply it to the classical model
of the hydrogen atom. The electron in the classical model has electrostatic po-
tential energy due to its attraction to the nucleus and kinetic energy due to its
orbit around the nucleus. In order to calculate the kinetic energy we calculate
the v 2 by equating the centripetal force to the electrostatic attraction.

mv 2 e2
=
r 4π0 r2
e2
v2 = (18.6)
4mπ0 r

Note that we use this expression for v 2 twice in our derivation.

Energy can now be calculated:

E = KE + PE
1 e2
= mv 2 + − (18.7)
2 4π0 r
e2 e2
= −
8π0 r 4π0 r
e2
=− (18.8)
8π0 r

Note that in equation 18.7 the PE is negative due to the opposite signs of the
electron and the nucleus and that the total energy (18.8) is negative due to
the bound state of the electron.

We can now introduce our quantisation criteria (18.5) by calculating the al-
lowed values of r. This makes use of the v 2 term from equation 18.6. The
difficult point is to remember that equation 18.5 should be rearranged to give
r squared.
136 CHAPTER 18. THE QUANTUM ATOM

n2 h2
r2 = (18.9)
4π 2 m2 v 2
n2 h2 4mπ0 r
= 2 2 (18.10)
4π m e2
2 2
n h 0
r= (18.11)
πme2

Finally, equation (18.11) is substituted into the equation for energy (18.8)

e2
E=−
8π0 r
e2 πme2
=−
8π0 n2 h2 0
me4
=− 2 2 2
80 n h
E1
= 2
n

me4
where E1 = − = −2.17 × 10−18 J = −13.6 eV
820 h2
which matches the empirical formula (18.1)!

Note that many calculators give a value of zero if you type this equation in as
one calculation. This is because me4 = 5.97 × 10−106 which casio calculators
cannot cope with. A way around this is to calculate directly in electron-volts
by dividing through by e thus requring only me3 to be calculated.

This powerful piece of reasoning also gives a value for the radius of the hydro-
gen atom which matches that measured by experiment.

This reasoning can be extended to nuclei with different charges; however it


only works with a single electron as further inter-electron interactions are not
taken into account.
19 Interpreting Quantum
Theory

Background Information

(This is not part of the Pre U syllabus but it will help with your understand-
ing.)

Wave Equation

We have already seen how oscillations can be expressed as second order dif-
ferential equations in Simple Harmonic Motion.

To describe waves we need to go a step further and consider how they vary
with time and distance.

Let’s start by imagining some ocean waves. If we stand on the end of a jetty,
i.e. in a fixed position, we can see the water moving up and down as time goes
by.

Equally we could take a photograph, i.e. a fixed time, and see how the wave
changes with distance.

So for a general wave equation we need a second order differential equation


for the oscillations with respect to time, with position kept constant, and with
respect to position with time kept constant. To do this we use partial differ-
entiation which is written with a curly ∂ and means keep the other variables
constant.

Here is the equation which describes wave motion where u is a function of


time and position and c is the speed of the wave.

∂2u 2
2∂ u
= c
∂t2 ∂x2

137
138 CHAPTER 19. INTERPRETING QUANTUM THEORY

The derivation of this is slightly beyond the level of the Pre U although it was
first done in the 18th Century.

Schrödinger’s Wave Equation

In the 1920’s, Austrian physicist Erwin Schrödinger used de Broglie’s idea


of particles having wavelengths to try and formulate a wave equation which
could explain some of the quantum phenomena which had been observed.

He came up with the following equation originally to describe the energy levels
in Hydrogen but it has been incredibly successful (matching experimental
data to a high degree of accuracy) in explaining a wide number of quantum
mechanical phenomena.

∂Ψ ~2 ∂ 2 Ψ
i~ =−
∂t 2m ∂x2

Note the partial differentials and Planck’s constant as well as the complex
numbers.

You will not need to know this but you need to know that it exists. Also
note the wave function Ψ (Psi). It is a function of space and time but when
Schrödinger first came up with the equation, he didn’t know what it repre-
sented.

(a) *interpret the double-slit experiment using the Copenhagen interpretation


(and collapse of the wavefunction), Feynman’s sum-over-histories and Ev-
erett’s many-worlds theory

The double-slit experiment involved shining a beam of light, electrons or other


particles through two small slits and observing the interference pattern cast
on a screen. This can be explained using wave theory, path difference and
superposition. (see earlier chapter.)

If the beam of light is reduced in intensity until only one photon passes through
the slits at a time, the individual photon will hit the screen at a unique place
although it is impossible to predict exactly where. If we continue to send
individual photons through we end up with the same diffraction pattern as
before.

The problem here is that the diffraction pattern needs a wave interpretation
whereas we are sending individual particles through. A particle can’t split up
between two slits and interfere with itself? So when the photon is released it
is a particle, as it passes through the slits it is a wave and when it hits the
screen it becomes a particle again.
139

What causes it to change from one thing to another?

Where is the photon just before it hits the screen?

What is it a wave of?

There hasn’t yet been a satisfactory explanation to these questions but here
are a few attempts.

• Copenhagen Interpretation
The Danish physicist Niels Bohr set up a conference in Copenhagen with
some of the best scientific minds of the time and came up with the first
explanation.
Schrodinger had already developed his famous wave equation which had
great success at calculating quantum interactions although he didn’t
establish what the wave function represented.
It was Max Born who later suggested that the wave function squared
(similar to intensity) represented the probability of finding a particle at
that point.
Just before the photon hits the screen it could be anywhere where the
wave function isn’t zero and yet once it hits the screen and a measure-
ment is made, it is in one precise point on the screen. We call this change
from having a probability of being in a number of places to being in one
definite location the wave function collapsing.
This is the Copenhagen interpretation, until a measurement is made,
the particle is simultaneously in all the possible different states and it is
only the act of making a measurement which forces it into one unique
outcome.

• Everett’s Many World’s Theory


A weakness with the Copenhagen interpretation is the question of what
is needed to make the wave function collapse? We have slightly skirted
around the issue by saying that it is when a measurement is made. But
what constitutes a measurement? Do we need a human observer? Par-
ticles will undergo many interactions, do they really exist in all possible
states until an observation is made?
Everett proposed a solution to this by saying that whenever multiple
opportunities occur, for example where our photon hits the screen, the
universe splits into multiple versions of itself. Each universe is identical
apart from this one difference so one universe will have the photon hitting
the centre of the screen whilst in another it will hit the first maximum.
When you make a reading the result will depend on which branch of
140 CHAPTER 19. INTERPRETING QUANTUM THEORY

reality you are in. There will be multiple universes with different versions
of you in them but you will never meet up. The only reality you will
know is what happens in the branch you are in.
This is great fun to think about, there will be a parallel universe where
you have already taken your pre U Physics a year early, scored full
marks and are now sipping cocktails on a beach. It also does away with
all the arbitrary reasons for the wave function collapsing. The problem
with it is the number of universes which are continually being generated.
Every time a subatomic particle interacts with another one, the universe
splits. Thinking about how many interactions happen every second leads
to a mind blowing number of universes and trying to visualise how they
can exist alongside each other, perhaps using extra dimensions, is near
impossible.

• Feynman’s sum-over-histories
The last interpretation was proposed by Feynman as a mathematical
way of dealing with quantum phenomena. In the case of the double slit
experiment we can only know that a photon leaves the laser and where
it arrives on the screen after it has been detected. We have no way of
knowing which slit it passed through or what route it took to get from
the laser to the screen. What Feynman did was to assume that the
photon takes every possible path. Not just straight lines, not just direct
routes. A photon could go to the ends of the universe and back as it
makes its way to the screen. If we add all the possible paths together
some of them will cancel out if they arrive out of phase, or reinforce if
they are in phase. What we end up with is a probability of the photon
being at any given point on the screen. This probability corresponds
exactly with the observed diffraction pattern.

(b) *describe and explain Schrödinger’s cat paradox and appreciate the use of
a thought experiment to illustrate and argue about fundamental principles

Unhappy with the Copenhagen interpretation for the quantum world, Schrödinger
set up a now famous thought experiment. The new quantum physics was very
counter intuitive in terms of observed physical phenomena and yet gave an
incredibly accurate model of the sub atomic world. But the world we see
around us is made up of particles so, by extension, the same physics should
hold true for both. What Schrödinger did was to take a purely random quan-
tum phenomena and use it to control a macroscopic event.

Schrödinger’s Cat Paradox A cat is put inside a box with a vial of poison
and a radioactive material. If there is a radioactive decay, it will break the
vial and release the poison and kill the cat. There is a 50/50 chance of the
141

substance decaying in a given time so there is a 50/50 chance of the cat


being dead or alive. According to the Copenhagen interpretation, until a
measurement is made, both states exist. So before the box is opened, the
particle has and hasn’t decayed and consequently the cat is simultaneously
dead and alive. Schrödinger’s use of a dead cat was inspired but he put
it forward to show how ridiculous the whole situation was. Later, after it
gathered an inordinate amount of attention he regretted ever coming up with
it.

In the many world’s interpretation, the universe splits into two. One has a
box with a decayed particle and a dead cat whilst in the other the cat is alive.
When we open the box we see the outcome according to whichever branch of
the universe we are in.

(c) *recognise and use ∆p∆x ≥ h


2π as a form of the Heisenberg uncertainty
principle and interpret it

Imagine two waves, one is a sine wave and the other is a short pulse.

The sine wave has a definite frequency and wavelength but stretches out in-
definitely. The pulse is made up of many sine waves of different frequencies
(see Fourier transforms for more on this) but has a fixed size.

So with waves we can see that there is a trade-off between a clearly defined
position and wavelength.

We also know that the deBroglie wavelength of a particle is related to the


momentum.
h
λ=
p

Mathematically the combination of position and wavelength which gives the


lowest combined uncertainty is with a Gaussian (normal) wave and from this
we can place a lower bound on the uncertainty in momentum and position.

Heisenberg’s Uncertainty Principle

h
∆p∆x ≥

so the more accurately we know the momentum of a particle, the less we can
know about its position.

note: this is not the effect of taking the measurement which introduces the
uncertainty but is an intrinsic limitation due to the wave nature of particles.
142 CHAPTER 19. INTERPRETING QUANTUM THEORY

(d) *recognise that the Heisenberg uncertainty principle places limits on our
ability to know the state of a system and hence to predict its future

If we roll a dice we consider it a random process but theoretically if we could


measure the exact velocity, spin, gravitational force, air resistance, frictional
forces of the table etc. we would know what side the dice would land on.

In the case of quantum particles we can’t know the initial conditions precisely
because of the limits imposed by Heisenberg’s principle. The more accurately
we know one thing, the less we know about the other and so we cannot accu-
rately predict what will happen in the future.

(e) *recall that Newtonian physics is deterministic, but quantum theory is


indeterministic

With Newtonian physics, the physics we observe in the everyday world, the
initial conditions will set into motion a chain of events which determine what
happens in the future. If you watch cricket on television you may have seen a
computer making leg before wicket decisions by continuing the trajectory of
the ball and seeing if it would have hit the wicket had the batsman’s leg not
been in the way. By knowing the flight of the ball the subsequent path could
be determined.

In quantum physics we cannot know all the initial conditions and so the future
cannot be determined.

We say that Newtonian physics is deterministic whilst quantum physics is


indeterministic.

(f) *understand why Einstein thought that quantum theory undermined the
nature of reality by being:

(i) indeterministic (initial conditions do not uniquely determine the future)

(ii) non-local (for example, wave-function collapse)

(iii) incomplete (unable to predict precise values for properties of particles).

Einstein was not happy with quantum theory and couldn’t accept that nature
could be governed by probability. His famous quote ”God does not play dice
with the universe” sums up his frustration.

Einstein wanted the universe to be deterministic but we have already seen


that this is not what quantum theory predicts.
143

Spukhaft Fernwirkung

Spooky action at a distance. Einstein was concerned that quantum physics


seemed to allow faster than light travel. Going back to the double slit exper-
iment, the instant before the particle hits the screen it could be in a whole
range of different places and yet the moment that the measurement is made,
the wave function collapses and it is in one unique place. How did it get there
so quickly?

He also proposed another famous thought experiment with Podoslki and Rosen
known as the EPR paradox.

If we start with two particles fired with equal velocities from a common origin
their subsequent motion can be described using a single, two-particle wave
function. After some time a measurement of the position of one of the particles
will instantaneously fix the other particle even though it could be a great
distance away and was not being measured directly.

The final thing which upset Einstein was not being able to know everything
about the state of a particle i.e. incomplete knowledge. We cannot know the
position and momentum simultaneously.

Quantum mechanics has been enormously successful with predictions match-


ing experimental results to an incredibly high level of accuracy. Yet the inter-
pretation of what it actually means is completely counter intuitive and there
is no currently accepted version of what is going on. As a result many physi-
cists, after struggling with the underlying philosophy for a few years, end up
with the Mantra;

don’t try and understand what is going on, just do the maths.
20 Astronomy and cosmology

Content

• standard candles

• stellar radii

• Hubble’s law

• the Big Bang theory

• the age of the Universe

(a) understand the terms luminosity and luminous flux

(b) recall and use the inverse square law for flux

L
F = (20.1)
4πd2

Stars are described as LUMINOUS because they emit electromagnetic waves.

The LUMINOSITY of a luminous (“hot”) object is defined as the amount


of electromagnetic wave energy emitted by the object per second i.e. it is
the emission Power of the object and is thus measured in Watts (W). It is a
measure of the absolute “brightness” of the object.

FLUX (F) is Power per unit area (an Intensity). For a spherical emitter of
radius R (and thus surface area of 4πR2 ), this means that F = 4πd
L
2.

145
146 CHAPTER 20. ASTRONOMY AND COSMOLOGY

(c) understand the need to use standard candles to help determine distances
to galaxies

Standard Candles are types of stars or galaxies for which the Luminosity
(the absolute brightness) can be determined directly from observations. By
measuring the observed brightness and comparing it to the absolute brightness
of the object, it is possible to determine how far away that object is.

The best-known and most widely used Standard Candles are Cepheid Vari-
ables and Type 1a Supernovae.

Cepheid Variables (named after the star Delta Cephei) pulsate and vary in
brightness with a frequency that is related to the star’s Luminosity. The peri-
ods of some Cepheids have been measured as a few days, others a few months.
Once the absolute brightness has been found from the observed period of the
pulsation, it can be compared with the brightness the star appears to have as
observed from the Earth and hence the distance can be determined.

Type 1a supernovae occur when one of the stars (a white dwarf) in a binary
system gains mass, becomes unstable and catastrophically explodes emitting
vast quantities of light and other electromagnetic energy in the process. The
maximum absolute brightness achieved is related to the rate at which the
emission fades (the so-called Light Curve). Thus, again, once the absolute
brightness is known, a comparison with how bright the object appears to be
from the Earth will yield its distance. Because they are so bright, these objects
can be easily seen in distant galaxies, so that distances well beyond the limits
of the Milky Way can be established.

(d) recognise and use Wien’s displacement law


1
λmax ∝ (20.2)
T
to estimate the peak surface temperature of a star either graphically or alge-
braically

Black bodies of a particular temperature T will emit radiation over a range


of wavelengths (a radiation distribution known as a Planck Curve):

The distribution peaks at a wavelength λmax which is determined by the re-


ciprocal of the absolute temperature T of the object. This is known as Wien’s
Displacement Law. The constant of proportionality has a value of 2.90 x 10-3
m K:

2.90 × 10−3
λmax =
T
147

figs/chapt-20/media/image1.png

Figure 20.1: Planck curves

The greater the temperature, the larger the area under the curve, suggesting
that more radiation is being emitted altogether, consistent with the Stefan-
Boltzmann relation. Refer to Figure 20.1 and consider the T = 8000 K and the
T = 6000 K graphs. The area beneath the T = 8000 K graph is (8000/6000)4
times that beneath the T = 6000 K graph, i.e. 3.16 times greater.

The Sun’s photosphere is at a temperature of a little under 6000 K; notice


that the Planck Curve for this temperature peaks at a wavelength lying in the
visible part of the electromagnetic spectrum.

Wien’s Displacement Law enables astronomers to determine the temperature


of a star from observations of the light (and other radiation) emitted by that
star.

(e) recognise and use Stefan’s law for a spherical body

L = 4πσr2 T 4 (20.3)

The Stefan-Boltzmann Law (sometimes simply called Stefan’s Law) states that
the Flux from a hot object is proportional to the fourth power of the abso-
lute temperature, T. Strictly speaking, this applies only to idealised radiation
emitters (referred to as “black bodies”); the constant of proportionality is the
Stefan-Boltzmann constant, σ, which has a value of 5.67 x 10-8 W m-2 K-4 :

F = σT 4

Therefore, for an emitted with surface area A, the Luminosity will be given
by

L = σAT 4

and in the case of a spherical emitter (such as a star) of radius R, this becomes
148 CHAPTER 20. ASTRONOMY AND COSMOLOGY

L = 4πR2 σT 4

(f) use Wien’s displacement law and Stefan’s law to estimate the radius of a
star

Having determined the temperature, T, one can then use the Stefan-Boltzmann
relation to determine the Luminosity, L. This is a measure of the actual
amount of radiation being emitted by the star. By measuring the actual
amount of radiation received per second per unit area (i.e. the flux, F), one
can then calculate how big the star must be (the surface area from which the
radiation is being emitted) in order to produce that amount of Flux.

Example Question
Worked example: Proxima Centauri, the nearest star to the Sun
Data:
Parallax angle = 0.8 seconds of arc
Peak wavelength, λmax = 967 nm
Flux measured at Earth = 3.56 x 1011 W m-2

1. Calculate the distance to Proxima Centauri.

2. Calculate its surface temperature.

3. Calculate the Flux at the star’s surface

4. Hence calculate the radius of the star.

Answer

1. D (in parsecs) = 1 / angle of parallax in seconds of arc = 1.25


pc = 3.85 x 1016 m

2. From Wien’s Displacement Law, T = 2.90 x 10-3 m K / 967 x


10-9 m = 3000 K

3. From Stefan’s Law, F = σ T4 = 5.67 x 10-8 x 8.10 x 1013 = 4.6


x 106 W m-2

4. Flux measured at a distance of 3.85 x 1016 m is 3.56 x 10-11 W


m-2 .
So, Fat Earth / Fstar’s surface =. (radius of star)2 / (distance to
star)2

• Radius of star = 1.07 x 108 m


149

(g) understand that the successful application of Newtonian mechanics and


gravitation to the Solar System and beyond indicated that the laws of physics
apply universally and not just on Earth

Newton’s law of Universal Gravitation states that two masses, M and m, whose
centres are separated by a distance r, will mutually attract with a gravitational
force given by

GM m
F =
r2

where G, the Universal Gravitational Constant, = 6.67 x 10-11 N m2 kg-2 .

One of the great triumphs of the law was to demonstrate consistency with
Kepler’s Laws of Planetary Motion, formulated empirically some 70 years
earlier. Kepler’s Laws state that the planets move about the Sun in elliptical
orbits and whilst Newton’s Law can be applied to such orbits, for simplicity we
consider a planet moving in a circular orbit. If the radius of the orbit is r, then
from rotational mechanics, the planet will experience a constant centripetal
force of mv 2 /r. This origin of this force is the gravitational attraction given
by Newton’s equation and by equating the two formulae it is possible to show
that

4π 2 r3
T2 = (20.4)
GM

i.e. that T2 α r3 as stated by Kepler’s 3rd Law. Newton’s Law, when com-
bined with his Laws of Motion, was applied to other objects observed to be
in gravitational orbits with great success. The movements of planetary satel-
lites, binary star systems, galactic spiral arms and even clusters of galaxies
themselves have all been shown to be consistent with the relationships. Fa-
mously, the relationships demonstrated orbital irregularities in the motion of
the planet Uranus, high led to the discovery of the planet Neptune beyond it.
Similar anomalous behaviour in the rotations of spiral galaxies observed by
Vera Rubin has led to the speculation of the existence of Dark Matter.
150 CHAPTER 20. ASTRONOMY AND COSMOLOGY

Example Question
From the orbital data for the Earth, calculate the mass of the Sun
(assuming a circular orbit).
Answer
Radius of Earth’s orbit = 150 x 106 km
Orbital period of earth = 1 year = 3.2 x 107 s

• From Kepler’s 3rd Law (above), M = 1.96 x 1030 kg

(h) recognise and use


∆λ ∆f v
≈ ≈ (20.5)
λ f c
for a source of electromagnetic radiation moving relative to an observer

When a source of waves moves away from an observer (in a stationary medium),
the observer will receive waves of a longer wavelength (and thus lower fre-
quency) than those emitted by the source. If the source moves towards the
observer then the observed wavelength is shorter (frequency is higher). This
is called Doppler Shift. For example, with sound waves this can be perceived
as a change in the pitch of the emitted sound.

The greater the speed of the source, v, the greater the change in the wavelength
∆λ (or frequency, ∆f , whichever is being measured) of the waves received by
the observer in which c is the velocity of the emitted waves (a relationship
strictly only true for if c is much greater than v).

The dark lines observed in the visible light spectra of stars and galaxies are
caused by the absorption of specific frequencies (colours) by elements present
in those objects, enabling astronomers to determine their composition. In the
1920s, Edwin Hubble discovered that the absorption lines for distant galaxies
were shifted towards the red end of the colour (emission) spectrum. This Red
Shift indicated that the galaxies were moving away (receding) from the Earth.
But there were two particular features of his discovery that made a special
impact:

a) Galaxies were receding in all directions.

b) The more distant the galaxy, the higher its recessional velocity.

(The distances to the galaxies were established using Standard Candles, espe-
cially Cepheid Variables.)
151

(i) state Hubble’s law and explain why galactic redshift leads to the idea that
the Universe is expanding and to the Big Bang theory

These observations have led to the conclusion that the universe began with
a Big Bang and what Hubble observed was the expansion of space itself. A
helpful picture is that of the infinite scaffolding by M C Escher (Figure 20.2):

figs/chapt-20/media/image2.jpeg

Figure 20.2: M C Escher drawing

No matter which junction you view from, if each scaffold pole is expanding,
junctions in all directions will appear to recede. Note that, in effect, the
junctions themselves do not move: the space in between them does. Also,
more distant junctions (with more expanding poles between them and the
observer) will appear to recede with greater speeds.

In such a model, the speed of “recession” (expansion) is proportional to the


distance of the observed galaxy and observations are by and large consistent
with this:

V = Ho d

in which the constant of proportionality, Ho, is known as the Hubble Constant,


the value of the gradient of the graph of v plotted against d (Figure 20.3):

Recent measurements (see graph) suggest a value of Ho of a little over 67 km


s-1 Mpc-1

Given that, in the observable universe, the greatest distance from which radi-
ation could be received is given by the speed of light x the age of the universe,
it follows that the age of the universe will be given by 1/Ho . For the value of
Ho quoted, this equates to an age of

When measuring the red shifts of distant galaxies, the calculated velocities
thus more properly indicate the rate at which the intervening space is stretch-
ing. The value of Δλ/λ (or Δf/f) is called the Cosmological Red Shift, z,
indicating that the space has expanded by a factor of 1 + z in order to pro-
duce the observed Doppler Shift.
152 CHAPTER 20. ASTRONOMY AND COSMOLOGY

figs/chapt-20/media/image3.jpeg

Figure 20.3: Hubble’s Law

The quasar 3C273 was the first object of its kind to be identified. So called
because they were star-like but much more luminous (“quasi-stellar objects”)
they are now known to consist of supermassive black holes that draw in a huge
disc of orbiting gas, causing large emissions of radiation across a wide range
of wavelengths. Their spectra exhibit large red shifts.
153

Example Question

1. One emission line in the spectrum of 3C273 appears at a wave-


length of 475.0 nm; in the laboratory the same line is measured
at 410.2 nm. Calculate the recessional velocity of the quasar.

2. Using the value of H0 quoted above, hence calculate the distance


of 3C273.

Answer

1. Δλ = 475.0 – 410.2 nm = 64.8 nm.

From Doppler equation, Δλ/λ = v/c, this gives v = 4.74 x 107 ms-1
or 4.74 x 104 km s-1 .

1. Assume a value of H0 of 67 km s-1 Mpc-1 .

• d = 4.74 x 104 km s-1 / 67 km s-1 Mpc-1 = 707 Mpc.

(j) explain how microwave background radiation provides empirical support for
the Big Bang theory

(k) understand that the theory of the expanding Universe involves the expan-
sion of space-time and does not imply a pre-existing empty space into which
this expansion takes place or a time prior to the Big Bang

(l) recall and use the equation

v ≈ H0 d (20.6)

for objects at cosmological distances

(m) derive an estimate for the age of the Universe by recalling and using the
Hubble time
1
t= (20.7)
H0
In the 1940s, George Gamow suggested that the observed ratio of hydrogen to
helium in the universe could be explained by assuming the universe was much
hotter and denser a long time ago, consistent with the idea of a Big Bang
origin stemming from Hubble’s work. His theory predicted the existence of a
“leftover” radiation which was eventually discovered by chance in 1965. This
is today known as the Cosmic Microwave Background Radiation and it has
the biggest cosmological red shift. It was produced when the early universe
had cooled down to about 3000K, low enough for electrons to combine with
protons to produce atoms, a process resulting in the emission of photons of
154 CHAPTER 20. ASTRONOMY AND COSMOLOGY

wavelengths around 1 mm. Because of cosmological expansion, the wavelength


is now a thousand times bigger (millimetres) and the temperature a thousand
times smaller, about 3 K.

Refinements to the Big Bang model have included a period of Inflation very
early on which gave rise to the eventual clumping of matter, which accounts for
the later existence of stars and galaxies (and planets and humans). The Cos-
mic Background Explorer satellite revealed such clumping (measured as tiny
temperature fluctuations in the background radiation) and further evidence is
still being sought to confirm Inflation as a part of the model.

figs/chapt-20/media/image4.jpeg

Figure 20.4: COBE satellite image


Appendices

155
A Equations

Equations to be learnt

Equations you need to remember for pre-U

velocity v= ∆x
∆t

acceleration a= ∆v
∆t

resultant force F = ma

momentum p = mv

resultant force F = ∆p
∆t

impulse Impulse = change in momentum

density Density = mass/volume

pressure Pressure = force/area

pressure in a liquid P = ρgh

weight W = mg

power P = Fv

GPE ∆E = mg∆h

change in gravitational potential = g∆h

energy in a spring E = 12 F x

157
158 APPENDIX A. EQUATIONS

efficiency % efficiency = useful energy or power


total energy or power in × 100

current I= ∆Q
∆t

potential difference V = W
Q

resistance R= V
I

electrical power P =VI

electrical work done W = V It

ρl
resistance R= A

resistors in series RT = R1 + R2 + . . .

resistors in parallel 1
RT = 1
R1 + 1
R2 + ...

frequency f = 1/T

wave speed v = fλ

Malus’ law Intensity ∝ cos2 θ

photoelectric equation hf = φ + 12 mvmax


2

angular velocity v = rω

period T = 2π/ω

circular motion F = mv 2 /r

electric field strength E = F/Q

capacitance C = Q/V

field strength/potential E = − dX
dV

the ideal gas law pV = nRT


E
Boltzmann factor e− kT

activity A = − dN
dt
EQUATIONS TO BE LEARNT 159

luminous flux F = L
4πd2

Hubble’s law v ≈ H0 d

Hubble time t = 1/H0

Equations you need to remember for paper 3, part B, but not


for any other part of the examination.

moment of inertia I = Σmr2

d2 x
shm dt2
= −ω 2 x

x = A cos ωt
160 APPENDIX A. EQUATIONS

Derivations

Equations you need to derive and remember for pre-U

energy in a spring E = 12 kx2

kinetic energy E = 12 mv 2

emf E = I(R + r)

emf E = V + Ir

electrical power P = I 2R

critical angle sin c = 1/n

centripetal acceleration a = v 2 /r

centripetal acceleration a = rω 2

uniform electric field F d = QV

uniform electric field E = V /d

energy in a capacitor W = 21 CV 2

1 Q2
energy in a capacitor W = 2 C

Q
electric field due to a point charge E= 4πε0 r2

Kepler’s third law r3 ∝ T 2

gravitational field strength g= Gm


r2

radius of curvature of particle in B-field r= mv


BQ

Hall effect V = Bvd

kinetic theory of gases pV = 31 N mhc2 i

activity of a source A = λN
DERIVATIONS 161

activity at time t A = A0 e−λt

half-life t1 = ln2
λ
2

Equations you need to derive for paper 3, part B, but not for
any other part of the examination.

moment of inertia of a ring I = mr2

moment of inertia of a disc I = mr2 /2

moment of inertia of a rod about one end I = ml2 /3

moment of inertia of a rod about its centre I = ml2 /12

velocity in shm v = −Aω sin ωt

acceleration in shm a = −Aω 2 cos ωt

total energy in shm E = 21 mA2 ω 2

Q1 Q2
electric potential (from electric force) W = 4πεo r

hydrogen atom energy levels En = −13.6eV


n2

You might also like