Download as pdf or txt
Download as pdf or txt
You are on page 1of 85

3.

1 Measurements and their errors

3.1.1 - Uses of SI units and their prefixes


SI units are the fundamental units, they are made up of:
● Mass (m): kg (kilograms)
● Length (l): m (metres)
● Time (t): s (seconds)
● Amount of substance (n): mol (moles)
● Temperature (t): K (kelvin)
● Electric current (I): A (amperes)

The SI units of quantities can be derived by their equation, e.g. F=ma


For example, to find the SI units of force (F) multiply the units of mass and acceleration kg x m s−2
gives kgm s−2 (Also known as N)

The SI units of voltage can be found by a series of steps:

E
● V= Q where E is energy and Q is charge , E=½ m v 2 so the SI units for energy is kg m2 s−2
(the units for speed (v) are ms−1 so squaring these gives m2 s−2 )

● Q=It (where I is current) so the units for Q are As (ampere seconds)

kgm2 s−2
● So V= As V= k gm2 s−3 A−1

Below are the prefixes which could be added before any of the above SI units:
Name Symbol Multiplier

Tera T 1012

Giga G 109

Mega M 106

Kilo k 103

Centi c 10−2

Milli m 10−3

Micro µ 10−6

Nano n 10−9

Pico p 10−12

Femto f 10−15
Converting mega electron volts to joules:
1eV=1.6x 10−19 J
e.g. convert 76 MeV to joules:
First, convert from MeV to eV by multiplying by 106 76 x 106 eV
Then convert to joules by multiplying by 1.6x 10−19 1.216 x 10−11 J

Converting kWh (kilowatt hours) to Joules:


1 kW = 1000 J/s 1 hour= 3600s
1kWh = 1000 x 3600
= 3.6x 106 J
= 3.6 MJ

3.1.2 - Limitation of Physical Measurements

Random errors affect precision, meaning they cause differences in measurements which causes
a spread about the mean. You cannot get rid of all random errors.
An example of random error is electronic noise in the circuit of an electrical instrument
To reduce random errors:
● Take at least 3 repeats and calculate a mean, this method also allows anomalies to be
identified.
● Use computers/data loggers/cameras to reduce human error and enable smaller
intervals.
● Use appropriate equipment, e.g a micrometer has higher resolution (0.1 mm) than a ruler
(1 mm).

Systematic errors affect accuracy and occur due to the apparatus or faults in the experimental
method. Systematic errors cause all results to be too high or too low by the same amount each
time.
An example is a balance that isn’t zeroed correctly (zero error) or reading a scale at a different
angle (this is a parallax error).
To reduce systematic error:
● Calibrate apparatus by measuring a known value (e.g. weigh 1 kg on a mass balance), if
the reading is inaccurate then the systematic error is easily identified.
● In radiation experiments correct for background radiation by measuring it beforehand and
excluding it from final results.
● Read the meniscus (the central curve on the surface of a liquid) at eye level (to reduce
parallax error) and use controls in experiments.

Precision Precise measurements are consistent, they fluctuate slightly about a mean
value - this doesn’t indicate the value is accurate

Repeatability If the original experimenter can redo the experiment with the same
equipment and method then get the same results it is repeatable
Reproducibility If the experiment is redone by a different person or with different techniques
and equipment and the same results are found, it is reproducible

Resolution The smallest change in the quantity being measured that gives a
recognisable change in reading

Accuracy A measurement close to the true value is accurate

The uncertainty of a measurement is the bounds in which the accurate value can be expected to
lie e.g. 20°C ± 2°C , the true value could be within 18-22°C

Absolute Uncertainty: uncertainty given as a fixed quantity e.g. 7 ± 0.6 V


3
Fractional Uncertainty: uncertainty as a fraction of the measurement e.g. 7 ± 35 V
Percentage Uncertainty: uncertainty as a percentage of the measurement e.g. 7 ± 8.6% V

To reduce percentage and fractional uncertainty, you can measure larger quantities.

Resolution and Uncertainty


Readings are when one value is found e.g. reading a thermometer, measurements are when the
difference between 2 readings is found, e.g. a ruler (as both the starting point and end point are
judged).

The uncertainty in a reading is ± half the smallest division,


e.g. for a thermometer the smallest division is 1°C so the uncertainty is ±0.5°C.
The uncertainty in a measurement is at least ±1 smallest division,
e.g. a ruler, must include both the uncertainty for the start and end value, as each end has
±0.5mm, they are added so the uncertainty in the measurement is ±1mm.

Digital readings and given values will either have the uncertainty quoted or assumed to be ± the
last significant digit e.g. 3.2 ± 0.1 V, the resolution of an instrument affects its uncertainty.

For repeated data the uncertainty is half the range (largest - smallest value), show as
range
mean ± 2 .

You can reduce uncertainty by fixing one end of a ruler as only the uncertainty in one reading is
included. You can also reduce uncertainty by measuring multiple instances,
e.g. to find the time for 1 swing of a pendulum by measuring the time for 10 giving e.g. 6.2 ± 0.1 s,
the time for 1 swing is 0.62 ± 0.01s (the uncertainty is also divided by 10).

Uncertainties should be given to the same number of significant figures as the data.

Combining uncertainties

● Adding / subtracting data - ADD ABSOLUTE UNCERTAINTIES


E.g. A thermometer with an uncertainty of ± 0.5 K shows the temperature of water falling from
298 ± 0.5 K to 273 ± 0.5K, what is the difference in temperature?

298-273 = 25K 0.5 + 0.5 = 1K (add absolute uncertainties) difference = 25 ± 1 K

● Multiplying / dividing data - ADD PERCENTAGE UNCERTAINTIES

E.g. a force of 91 ± 3 N is applied to a mass of 7 ± 0.2 kg, what is the acceleration of the mass?

uncertainty
a=F/m =91/7 = 13m s−2 percentage uncertainty= value
× 100

3 0.2
Work out % uncertainties 91 × 100 + 7 × 100 = 3.3% + 2.9% add % uncertainties
= 6.2%
So a= 13 ± 6.2% m s−2 6.2% of 13 is 0.8
a=13 ± 0.8 m s−2
● Raising to a power - MULTIPLY PERCENTAGE UNCERTAINTY BY POWER

The radius of a circle is 5 ± 0.3 cm, what is the percentage uncertainty in the area of the circle?
Area = π x 25 = 78.5 cm2
Area = π r2
% uncertainty in radius = 0.3
5 × 100 = 6% % uncertainty in area = 6 x 2 (2 is the power from r2 )
= 12%
78.5 ± 12% cm2

Uncertainties and graphs


Uncertainties are shown as error bars on graphs,
e.g. if the uncertainty is 5mm then have 5 squares of error bar on either side of the point
A line of best fit on a graph should go through all error bars (excluding anomalous points).

The uncertainty in a gradient can be found by lines of best and worst fit, this is especially useful
when the gradient represents a value such as the acceleration due to gravity:
● Draw a steepest and shallowest line of worst fit, it must go through all the error bars.
● Calculate the gradient of the line of best and worst fit, the uncertainty is the difference
between the best and worst gradients.

|best gradient−worst gradient|


percentage uncertainty = best gradient × 100%
(modulus lines show it’s always positive)

When the best and worst lines have different y intercepts, you can find the uncertainty in the
y-intercept, which is |best y intercept-worst y intercept|:

|best y intercept−worst y intercept|


percentage uncertainty = best y intercept × 100%
3.1.3 - Estimation of physical quantities

Orders of magnitude - Powers of ten which describe the size of an object, and which can also be
used to compare the sizes of objects.
E.g: The diameter of nuclei have an order of magnitude of around 10−14 m.
100 m is two orders of magnitude greater than 1m.

You may be asked to give a value to the nearest order of magnitude, here you must simply
calculate the value the question is asking you for and give it only as a power of ten
E.g If the diameter of a hydrogen atom is 1.06 × 10−10 m, find the approximate area of the entire
atom (assuming it is perfectly spherical), to the nearest order of magnitude.
Find the area using A= π r2 π × (0.53 × 10−10 )2 = 8.82 × 10−21 m = 1 × 10−20 m (to 1 s.f)
Therefore the area to the nearest order of magnitude is 10−20 m.

Estimation is a skill physicists must use in order to approximate the values of physical quantities,
in order to make comparisons, or to check if a value they’ve calculated is reasonable.
3.2.1 Particles

3.2.1.1 - Constituents of the atom

An atom is formed of 3 constituents: protons, neutrons and electrons. At the centre of an


atom is a nucleus formed of protons and neutrons, therefore they are known as nucleons,
whereas electrons orbit the nucleus in shells.

These particles have properties which can be described in two ways, in SI units
and relative units, as shown below:

Particle Charge (C) Relative Mass (kg) Relative Specific


Charge Mass Charge
(Ckg-1)

Proton + 1.6 ✕ 10−19 +1 1.67 ✕ 10−27 1 9.58 ✕ 107

Neutron 0 0 1.67 ✕ 10−27 1 0

Electron − 1.6 ✕ 10−19 -1 9.11 ✕ 10−31 0.0005 1.76 ✕ 1011

The specific charge of a particle is the charge-mass ratio, and is calculated by dividing a
particle’s charge by its mass.
For example:
A proton has a charge of + 1.6 ✕ 10−19 C , and a mass of 1.67 ✕ 10−27 kg.
1.6×10−19
So its specific charge = 7
−27 = 9.58 × 10 Ckg
−1
1.67 ✕ 10

The proton number is the number of protons in an atom and is denoted by Z , while the
nucleon number is the number of protons and neutrons, denoted by A . These will often be
shown in the form: where ‘X’ is the symbol for the element.
Isotopes are atoms with the same number of protons but different numbers of neutrons. For
example, carbon-14 is a radioactive isotope of carbon, which can be used to find the
approximate age of an object containing organic material. This is done through carbon
dating, which involves calculating the percentage of carbon-14 remaining in the object, and
using the known starting value of carbon-14 (which is the same for all living things) and its
half-life to calculate an approximate age.

3.2.1.2 - Stable and unstable nuclei

The strong nuclear force (SNF) keeps nuclei stable by counteracting the electrostatic force
of repulsion between protons in the nucleus (as they have the same charge). It only acts on
nucleons and has a very short range, where it is attractive up to separations of 3 fm, but
repulsive below separations of 0.5 fm, which is demonstrated in the graph below:

Unstable nuclei are those which have too many of either protons, neutrons or both causing
the SNF to not be enough to keep them stable, therefore these nuclei will decay in order to
become stable. The type of decay the nuclei will experience depends on the amount of each
nucleon in them.

Alpha decay occurs in large nuclei, with too many of both protons and neutrons. A general
equation for alpha decay is:

● The proton number decreases by 2.


● The nucleon number decreases by 4.

Beta-minus decay occurs in nuclei which are neutron-rich (have too many neutrons). A
general equation for beta-minus decay is:
● The proton number increases by 1.
● The nucleon number stays the same.
At first, scientists believed that only an electron (beta-minus particle) was emitted from the
nucleus during beta-minus decay, however observations of the energy levels of the particles
before and after the decay showed that energy was not conserved. This does not follow the
principle of conservation of energy, and therefore neutrinos were hypothesised to account
for this, and later they were observed.

3.2.1.3 - Particles, antiparticles and photons

For every type of particle there is an antiparticle which has the same rest energy and mass
but all its other properties are opposite the particles. For example, the positron is the
antiparticle of the electron, and an electron antineutrino is the antiparticle of a neutrino; this
is how their properties compare:
Particle Mass (kg) Rest energy (Mev) Charge (C)

Electron ( e− ) 9.11 ✕ 10−31 0.511 − 1.6 ✕ 10−19

Positron ( e+ ) 9.11 ✕ 10−31 0.511 + 1.6 ✕ 10−19

Electron neutrino ( ν e ) 0 0

Electron antineutrino ( ν e ) 0 0

Electromagnetic radiation travels in packets called photons, which transfer energy and have
no mass. The energy of photons is directly proportional to the frequency of electromagnetic
radiation, as shown in the equation:
hc
E = hf = λ where h is the planck constant, which is equal to

6.63 × 10−34 Js .

Annihilation is where a particle and its corresponding antiparticle collide, as a result their
masses are converted into energy. This energy, along with the kinetic energy of the two
particles is released in the form of 2 photons moving in opposite directions in order to
conserve momentum.
An important example of an application of annihilation is in a PET scanner, which allows 3D
images of the inside of the body to be taken, therefore making medical diagnoses easier.
This is done by introducing a positron-emitting radioisotope into the patient, as positrons are
released they annihilate with electrons already in the patients system, emitting gamma
photons which can easily be detected.

Pair production is where a photon is converted into an equal amount of matter and
antimatter. This can only occur when the photon has an energy greater than the total rest
energy of both particles, any excess energy is converted into kinetic energy of the particles.

3.2.1.4 - Particle interactions

There are four fundamental forces: gravity, electromagnetic, weak nuclear and strong
nuclear.

Forces between particles are caused by exchange particles. Exchange particles carry
energy and momentum between the particles experiencing the force and each fundamental
force has its own exchange particles. A good example to think about to describe repulsion, is
to imagine an exchange as a heavy ball being thrown from one person to another, as the ball
is thrown it carries momentum to the second person causing them to move back. This same
principle can be applied to describe attraction, but instead of a heavy ball, the exchange
particle is a boomerang.

Interaction Exchange particle Range (m) Acts on

Strong Gluon 3 × 10-15 Hadrons

Weak W boson ( W + or W − ) 10-18 All particles

Electromagnetic Virtual photon ( γ ) Infinite Charged


particles

Gravity Graviton (not on specification) Infinite Particles with


mass
The weak nuclear force is responsible for beta decay, electron capture, and electron-proton
collisions, all of which can be represented as the particle interaction diagrams below.

p + e− → n + ν e
Electron capture: Electron-proton collision:

p + e− → n + ν e

As shown above, the equations for electron capture and an electron-proton collision are the
same, however a different exchange particle is used.

Beta-plus decay: p → n + e+ + ν e Beta-minus decay: n → p + e− + ν e

3.2.1.5 - Classification of particles

All particles are either hadrons or leptons. Their differentiating property is that leptons are
fundamental particles, meaning they cannot be broken down any further, and they do not
experience the strong nuclear force. On the other hand, hadrons are formed of quarks
(quarks are fundamental particles), and hadrons experience the strong nuclear force.

Hadrons can be further separated into baryons, antibaryons and mesons. Baryons are
formed of 3 quarks, antibaryons are formed of 3 antiquarks while mesons are formed from
a quark and antiquark.
The particles you need to know about are summarised in the diagram below:

The baryon number of a particle, shows whether it is a baryon (if 1), antibaryon (if -1) or not
a baryon (if 0). Baryon number is always conserved in particle interactions.

The proton is the only stable baryon, therefore all baryons will eventually decay into a
proton either directly or indirectly.

The lepton number of a particle, shows whether it is a lepton (if 1), antilepton (if -1) or not a
lepton (if 0). There are two types of lepton number you need to know, electron lepton
number and muon lepton number, which represent the named particle. Lepton number is
always conserved in particle interactions.

A muon is sometimes known as a “heavy electron”, and muons decay into electrons.

Strange particles are particles which are produced by the strong nuclear interaction but
decay by the weak interaction. The only strange particles you are expected to know about
are kaons, which decay into pions, through the weak interaction.

Strangeness is a property of particles, which shows that strange particles must be created
in pairs, as strangeness must be conserved in strong interactions. However, in weak
interactions strangeness can change by 0, +1 or -1.

In order to investigate particle physics, particle accelerators may be built however as these
are very expensive to build and run, and also produce huge amounts of data, scientific
investigations rely on collaboration of scientists internationally.
3.2.1.6 - Quarks and antiquarks

As mentioned above, quarks are fundamental particles which make up hadrons. Each type
of quark has different properties, and you need to be aware of 3 types of quark and
antiquark.

For the below table, the properties of the respective antiquarks have the opposite sign:

Type of Charge Baryon Strangeness


quark number

Up (u) + 32 e + 1
3
0

Down (d) − 31 e + 1
3
0

Strange (s) − 31 e + 1
3
-1
You also need to be aware of the following quark combinations for mesons:

Particle Quark combination(s) Charge Strangeness

π⁰ uu or dd 0 0

π⁺ ud +1 0

π⁻ ud -1 0

k⁰ ds 0 +1

k⁺ us +1 +1

k⁻ us -1 -1

As all baryons decay into protons, a neutron will decay into a proton. The equation for this
decay is: n → p + e− + ν e
3.2.1.7 - Applications of conservation laws

These properties must always be conserved in particle interactions:


● Energy and momentum
● Charge
● Baryon number
● Electron lepton number
● Muon lepton number

Strangeness must only be conserved during strong interactions.

To show these conservation laws are obeyed in an interaction, you must find the value of
each property before and after the interaction and make sure they are equal.
For example, beta-minus decay:
(as this is a weak interaction, strangeness does not need to be conserved)
n → p + e− + ν e
Charge Baryon Electron Muon Strangeness
number lepton number lepton
number

Before interaction 0 1 0 0 0

After interaction 1-1+0=0 1+0+0 0+1-1 0+0+0 0+0+0

Change 0 0 0 0 0
All the conservation laws are obeyed therefore this interaction is possible.
Beta-minus and beta-plus decay are both caused by the weak interaction because there is
a change of quark type. This can be shown in the below diagrams:

Beta-minus decay Beta-plus decay

In beta-minus decay, for a neutron to change into a proton, a down quark changes into an up
quark. And vice versa in beta-plus decay.

3.2.2 Electromagnetic radiation and quantum phenomena

3.2.2.1 - The photoelectric effect

The photoelectric effect is where photoelectrons are emitted from the surface of a metal
after light above a certain frequency is shone on it. This certain frequency is different for
different types of metals and is called the threshold frequency.

Image source: Wolfmankurd,CC BY-SA 3.0

The threshold frequency couldn’t be explained by the wave theory, as it suggests that any
frequency of light should be able to cause photoelectric emission as the energy absorbed by
each electron will gradually increase with each incoming wave. However, it could be
explained by the photon model of light which suggested that:
● EM waves travel in discrete packets called photons, which have an energy
which is directly proportional to frequency.
● Each electron can absorb a single photon, therefore a photoelectron is
only emitted if the frequency is above the threshold frequency.
● If the intensity of the light is increased, if the frequency is above the threshold,
more photoelectrons are emitted per second.
The work function of a metal is the minimum energy required for electrons to be emitted
from the surface of a metal, and it is denoted by ϕ .
The stopping potential is the potential difference you would need to apply across the metal
to stop the photoelectrons with the maximum kinetic energy. Measuring stopping potential
allows you to find the maximum kinetic energy of the released photoelectrons, as
E k (max) = eV s . Where V s is the stopping potential and e is the charge of an electron.
This is derived using the fact that energy = charge × voltage .

The photoelectric equation is E = hf = Φ + E k (max) , and it shows the relationship


between the work function, maximum kinetic energy and the frequency of light.

3.2.2.2 - Collisions of electrons with atoms

Electrons in atoms can only exist in discrete energy levels. These electrons can gain energy
from collisions with free electrons, which can cause them to move up in energy level, this is
known as excitation, or they can gain enough energy to be removed from the atom entirely,
this is called ionisation. Ionisation occurs if the energy of the free electron is greater than
the ionisation energy.

If an electron becomes excited, it will quickly return to its original energy level (the ground
state), and therefore release the energy it gained in the form of a photon.

An example of a practical use of excitation is in a fluorescent tube in order to produce light.


Fluorescent tubes are filled with mercury vapour, across which a high voltage is applied, as
shown in the below diagram:
This voltage accelerates free electrons through the tube, which collide with the mercury
atoms causing them to become ionised, releasing more free electrons. The free electrons
collide with the mercury atoms, causing them to become excited. When they de-excite they
release photons, most of which are in the UV range. The (phosphorous) fluorescent
coating on the inside of the tube, absorbs these UV photons and therefore electrons in the
atoms of the coating become excited and de-excite releasing photons of visible light.

When describing the energy difference between energy levels, the values of energy are very
small, therefore the unit, electron volts (eV) is used instead of joules (J). An electron volt is
defined as the energy gained by one electron when passing through a potential difference of
1 volt.

We know that energy = charge × voltage , therefore 1 eV = e × 1 = 1.6 × 10−19 .


And so to convert from eV to joules, multiply your value by 1.6 × 10−19 .
To convert from joules to eV, divide your value by 1.6 × 10−19 .

3.2.2.3 - Energy levels and photon emission

By passing the light from a fluorescent tube through a diffraction grating or prism, you get a
line spectrum. Each line in the spectrum will represent a different wavelength of light
emitted by the tube. As this spectrum is not continuous but rather contains only discrete
values of wavelength, the only photon energies emitted will correspond to these
wavelengths, therefore this is evidence to show that electrons in atoms can only transition
between discrete energy levels.

You can also do this by passing white light through a cooled gas however you would get a
line absorption spectrum, which looks like a continuous spectrum of all possible
wavelengths of light, with black lines at certain wavelengths. These lines represent the
possible differences in energy levels as the atoms in the gas can only absorb photons of an
energy equal to the exact difference between two energy levels.
The difference between two energy levels is equal to a specific photon energy emitted by a
fluorescent tube, or absorbed in a line absorption spectrum. Therefore: ΔE = E 1 − E 2 ,
where E 1 and E 2 , represent energy levels:

As E = hf , hf = E 1 − E 2

3.2.2.4 - Wave-particle duality


Light can be shown as having both wave and particle properties. Examples of light acting as
a wave are diffraction and interference, while an example of light acting as a particle is the
photoelectric effect.

Electrons can also be shown as having both wave and particle properties, as the wave
nature of electrons can be observed through electron diffraction, as only waves can
experience diffraction.

De Broglie hypothesised that if light was shown to have particle properties, then particles
should also have wave-like properties, and he wrote an equation relating the wavelength ( λ )
of an object to its momentum ( mv ):
h
λ= mv where h is the planck constant.
Using the above equation you can see how the amount of diffraction changes as a particle’s
momentum changes. When the momentum is increased, the wavelength will decrease, and
therefore the amount of diffraction decreases, so the concentric rings of the interference
pattern become closer. Whereas, when momentum is decreased, the wavelength increases,
the amount of diffraction increases so the rings move further apart.

Scientists did not always agree that matter had this wave-particle duality, however as time
went on and experimental evidence for this phenomena was gathered (notably electron
diffraction and the photoelectric effect), it was eventually accepted. Knowledge and
understanding of any scientific concept changes over time in accordance to the experimental
evidence gathered by the scientific community. However, these pieces of experimental
evidence must first be published to allow them to be peer-reviewed by the community to
become validated, and eventually accepted.
3.3.1 Progressive and stationary waves

3.3.1.1 - Progressive Waves

A progressive wave transfers energy without transferring material and is made up of particles
of a medium (or field) oscillating e.g. water waves are made of water particles moving up and
down

Amplitude A wave’s maximum displacement from the equilibrium position (units are m)

Frequency, f The number of complete oscillations passing through a point per second,
(units are Hz)

Wavelength, λ The length of one whole oscillation (e.g. the distance between successive
peaks/troughs) (units are m)

Speed, c Distance travelled by the wave per unit time, (units are m/s)

Phase The position of a certain point on a wave cycle, (units are radians, degrees or
fractions of a cycle)

Phase How much a particle/wave lags behind another particle/wave, (units are
difference radians, degrees or fractions of a cycle)

Period, T Time taken for one full oscillation, (units are s)

Two points on a wave are in phase if they are both at the same point of the wave cycle, they will
have the same displacement and velocity and their phase difference will be a multiple of 360°
(2π radians), they do not need the same amplitude, only the same frequency and wavelength.

Two points are completely out of phase when they’re an odd integer of half cycles apart e.g. 5
half cycles apart where one half cycle is 180° (π radians).

The speed of a wave is equal to the wave’s frequency multiplied by its wavelength:
c = fλ
The frequency of a wave is equal to 1 over its period:
1
f= T

3.3.1.2 - Longitudinal and Transverse Waves

Transverse waves - oscillation of particles (or fields) is at right angles to the direction of energy
transfer
● All electromagnetic (EM) waves are transverse and travel at 3 x 108 ms-1 in a vacuum.
● Transverse waves can be demonstrated by shaking a slinky vertically or through the
waves seen on a string, when it's attached to a signal generator.

Longitudinal waves - oscillation of particles is parallel to the direction of energy transfer


● These are made up of compressions and rarefactions and can’t travel in a vacuum.
● Sound is an example of a longitudinal wave, and they can be demonstrated by pushing a
slinky horizontally.

A polarised wave oscillates in only one plane (e.g only up and down), only transverse waves
can be polarised. Polarisation provides evidence for the nature of transverse waves because
polarisation can only occur if a wave’s oscillations are perpendicular to its direction of travel (as
they are in transverse waves).

Polaroid sunglasses are an application of polarisation. They reduce glare by blocking partially
polarised light reflected from water and tarmac, as they only allow oscillations in the plane of the
filter, making it easier to see.
Another application of polarisation is TV and radio signals, which are usually plane-polarised by
the orientation of the rods on the transmitting aerial, so the receiving aerial must be aligned in the
same plane of polarisation to receive the signal at full strength.

3.3.1.3 - Principle of Superposition of waves and formation of stationary waves

Superposition is where the displacements of two waves are combined as they pass each other,
the resultant displacement is the vector sum of each wave’s displacement. There are two types of
interference that can occur during superposition:
● Constructive interference occurs when 2 waves have displacement in the same
direction
● Destructive interference occurs when one wave has positive displacement and the other
has negative displacement, if the waves have equal but opposite displacements, total
destructive interference occurs
Image source: Haade,CC BY-SA 3.0, Image is recoloured

A stationary wave is formed from the superposition of 2 progressive waves, travelling in


opposite directions in the same plane, with the same frequency, wavelength and amplitude.

No energy is transferred by a stationary wave.

● Where the waves meet in phase, constructive interference occurs so antinodes are
formed, which are regions of maximum amplitude.
● Where the waves meet completely out of phase, destructive interference occurs and
nodes are formed, which are regions of no displacement.

A string fixed at one end, and fixed to a driving oscillator at the other gives a good example of the
formation of a stationary wave:
● A wave travelling down the string from the oscillator will be reflected at the fixed end of the
string, and travel back along the string causing superposition of the two
waves, and because the waves have the same wavelength, frequency
and amplitude, a stationary wave is formed. (Labelled combined wave on diagram).

Image source: Wjh31,CC0 1.0, Image is cropped

The lowest frequency at which a stationary wave forms is the first harmonic, which forms a
stationary wave with two nodes and a single antinode. The distance between adjacent nodes (or
antinodes) is half a wavelength (for any harmonic).

You can calculate this frequency by using this formula:


Where L is the length of the vibrating string, T is the tension and μ is the mass per unit length.

You can double the first harmonic frequency to find the second harmonic where there are 2
antinodes, you triple the first harmonic frequency to get the third harmonic where there are 3
antinodes, and so on for the nth harmonic.

Image source: Qef, CC0 1.0, Image is cropped

There are several examples of stationary waves:


● Stationary microwaves can be formed by reflecting a microwave beam at a metal plate, to
find the nodes and antinodes use a microwave probe.
● Stationary sound waves can be formed by placing a speaker at one end of a closed glass
tube, lay powder across the bottom of the tube, it will be shaken at the antinodes and settle
at the nodes. The distance between each node is half a wavelength, and the frequency of
the signal generator to the speaker is known so by c=fλ the speed of sound in air can be
found.
3.3.2 Refraction, diffraction and interference

3.3.2.1 - Interference

Path difference is the difference in the distance travelled by two waves.


A coherent light source has the same frequency and wavelength and a fixed phase difference.
Lasers are an example of light which is coherent and monochromatic, meaning they emit a single
(or small range of) wavelength(s) of light. Lasers are usually used as sources of light in diffraction
experiments as they form clear interference patterns.

Young’s double slit experiment demonstrates interference of light from two-sources. In this
experiment you can use two coherent sources of light or you could use one coherent source and a
double slit in order to form an interference pattern. If you do not have a coherent source of light for
example a light bulb, you could place a single slit before the double slit to make the light have a
fixed path difference, and a filter to make the light monochromatic. Below is a brief procedure to
describe Young’s double slit experiment:
● Shine a coherent light source through 2 slits about the same size as the wavelength of the
laser light so the light diffracts
● Each slit acts as a coherent point source making a pattern of light and dark fringes. Light
fringes are formed where the light meets in phase and interferes constructively, this occurs
where the path difference between waves is a whole number of wavelengths (nλ, where n
is an integer). Dark fringes are formed where the light meets completely out of phase and
interferes destructively, this occurs where the path difference is a whole number and a half
wavelengths ((n+½)λ).

Image source: Stannered, CC BY-SA 3.0

λD
The formula associated with the above experiment is: w= s Where w is the fringe spacing,

λ is the wavelength of light used, D is the distance between the screen and slits, and s is slit separation.

Using white light instead of monochromatic laser light gives wider maxima and a less intense
diffraction pattern with a central white fringe with alternating bright fringes which are spectra,
violet is closest to the central maximum and red furthest.
Lasers can permanently damage your eyesight therefore, when using lasers there are several
safety precautions, which must be followed:
● Wear laser safety goggles
● Don’t shine the laser at reflective surfaces
● Display a warning sign
● Never shine the laser at a person

The type of interference described above can also be demonstrated in sound waves through a very
similar process, however instead of using a double slit, you could use two speakers connected to
the same signal generator. And the intensity of the wave can be measured using a microphone to
find the maxima (equivalent to light fringes), and minima (equivalent to dark fringes).

Evidence for the wave nature of light was provided by Young's double slit experiment because
diffraction and interference are wave properties, and so proved that EM radiation must act as a
wave (at least some of the time). However, this was not was people always thought, there were
theories which suggested light was formed of tiny particles, however this experiment disproved that
theory. Knowledge and understanding of any scientific concept changes over time in
accordance to the experimental evidence gathered by the scientific community.

3.3.2.2 - Diffraction

Diffraction is the spreading out of waves when they pass through or around a gap. The greatest
diffraction occurs when the gap is the same size as the wavelength. When the gap is smaller than
the wavelength most waves are reflected, whereas when it is larger there is less noticeable
diffraction. When a wave meets an obstacle you get diffraction round the edges, the wider the
obstacle compared to the wavelength, the less diffraction.

Monochromatic light can be diffracted through a single slit onto a screen, which forms an
interference pattern of light and dark fringes. The pattern has a bright central fringe, which is
double the width of all other fringes, with alternating dark and bright fringes on either side, the
bright fringes are caused by constructive interference where the waves meet in phase and the
dark fringes are caused by destructive interference where waves arrive completely out of
phase. The intensity of the fringes decreases from the central fringe as shown below:

Image source: Nmurdoch,CC0 1.0


White light could be used instead of monochromatic light and the diffraction pattern you’d see
would be quite different. As white light is made up of all colours, therefore all different
wavelengths of visible light, the different wavelengths of light are all diffracted by different
amounts so you get a spectrum of colour in the diffraction pattern. The diffraction pattern for white
light has a central white maximum with alternating bright fringes which are spectra, violet is
closest to the central maximum and red furthest away. Below are the diffraction patterns for
white light (on the left), and red light (on the right):

Image source: Pieter Kuiper, CC0 1.0

In order to vary the width of the central maximum, you can vary slit width and wavelength:
● Increasing the slit width decreases the amount of diffraction so the central maximum
becomes narrower and its intensity increases.
● Increasing the light wavelength increases the amount diffraction as the slit is closer in
size to the light’s wavelength, therefore the central maximum becomes wider and its
intensity decreases.

A diffraction grating is a slide containing many equally spaced slits very close together. When
monochromatic light is passed through a diffraction grating the interference pattern is much
sharper and brighter than it would be after being passed through a double slit like in Young’s
double slit, this is because there are many more rays of light reinforcing the pattern. This means
measurements of slit widths are much more accurate as they are easier to take.
Below you can see how the intensity of an interference pattern varies as the number of slits
increases: the red graph shows the pattern with a grating with 50 slits, while the green shows the
pattern with a grating with 20 slits.

Image source: Epzcaw,CC BY-SA 3.0


The ray of light passing through the centre of a diffraction grating is called the zero order line,
lines either side of the zero order are the first order lines, then the lines outside the two first order
lines are the second order lines, and so on as showcased in the diagram on the left.

The formula associated with diffraction gratings is d sinθ = nλ


Where d is the distance between the slits, θ is the angle to the normal made by the maximum, n is
the order and λ is the wavelength. As λ increases (by for example changing the laser light from
blue to red), the distance between the orders will increase because θ is larger due to the increase
in diffraction as the slit spacing is closer in size to the wavelength, this means the pattern will
spread out.
The maximum value of sin θ is 1, therefore any values of n, which give sin θ as greater than 1 are
impossible.

You must be able to derive the above formula, the derivation is shown below:

1. Considering the first order maximum, where the path difference between two adjacent
rays of light is one wavelength (as shown in the diagram below), name the angle
between the normal to the grating and the ray of light θ.
2. As you can see a right angle triangle is formed, with side lengths d and λ. And by using the
fact that a right angle is 90°, and angles in a triangle add up to 180°, you can see the upper
angle in the triangle is θ (because the lower angle is 90-θ°).
λ
3. By using trigonometry we can see that for the first maximum sin θ = , (as sin θ =
d
Opp/Hyp) which rearranges to dsin θ = λ , (for the first order).
4. We know that the other maxima occur when the path difference between the two rays of
light is nλ, where n is an integer, therefore we can generalise the equation by replacing λ
with nλ to get: d sinθ = nλ .

Image source: Sjlegg,CC0 1.0, Image is cropped

There are several applications of diffraction gratings:


● You split up light from stars using a diffraction grating to get line absorption spectra
which can be used to show which elements are present in the star.
● X-ray crystallography, which is where x-rays are directed at a thin crystal sheet which acts
as a diffraction grating to form a diffraction pattern, this is because the wavelength of x-rays
is similar in size to the gaps between the atoms. This diffraction pattern can be used to
measure the atomic spacing in certain materials.

3.3.2.3 - Refraction at a plane surface

A refractive index (n) is a property of a material which measures how much it slows down light
passing through it. This is calculated by dividing the speed of light in a vacuum (c) by the speed of
light in that substance (cs).
c
n= cs
A material with a higher refractive index can also be known as being more optically dense.
As light doesn’t slow down significantly when travelling through air (in comparison to travelling
through a vacuum), the refractive index of air is approximately 1.

Refraction occurs when a wave enters a different medium, causing it to change direction, either
towards or away from the normal depending on the material’s refractive index.
Snell’s law is used for calculations involving the refraction of light:
n1 sinθ1 = n2 sinθ2
➔ n1 is the refractive index of material 1,
➔ n2 is the refractive index of material 2,
➔ θ1 is the angle of incidence of the ray in material 1
➔ θ2 is the angle of refraction of the ray in material 2

Image source: Oleg Alexandrov,CC BY-SA 3.0

As the light moves across the boundary of the 2 materials its speed changes, which causes its
direction to change.
In the example above, n2 is more optically dense than n1, therefore the ray of light slows down
and bends towards the normal. However, in the case where n2 is less optically dense than n1
the ray of light will bend away from the normal.

As the angle of incidence is increased, the angle of refraction also increases until it gets closer to
90°. When the angle of refraction is exactly 90° and the light is refracted along the boundary,
the angle of incidence has reached the critical angle (θc). This angle can be found using the
following formula:
n2
sin θc = n where n1 > n2
1

Total internal reflection (TIR) can occur when the angle of incidence is greater than the
critical angle and the incident refractive index (n1) is greater than the refractive index of the
material at the boundary (n2).

Image source: Josell7,CC BY-SA 3.0

A useful application of total internal reflection are optical fibres; these are flexible, thin tubes of
plastic or glass which carry information in the form of light signals. They have an optically dense
core surrounded by cladding with a lower optical density allowing TIR to occur, this cladding also
protects the core from damage and prevents signal degradation through light escaping the
core, which can cause information to be lost.

Image source: User A1,CC BY-SA 3.0


Signal degradation can also be caused by the following:
● Absorption - where part of the signal’s energy is absorbed by the fibre, reducing the
amplitude of the signal, which could lead to a loss of information
● Dispersion - this causes pulse broadening, which is where the received signal is broader
than the original transmitted signal. Broadened signals can overlap causing loss of
information. There are two types of dispersion:
➔ Modal - caused by light rays entering the fibre at different angles, therefore they
take different paths along the fibre, (for example some may travel down the middle
of the fibre, while others are reflected repeatedly,) this leads to the rays taking a
different amount of time to travel along the fibre, causing pulse broadening.
➢ This can be reduced by making the core very narrow, therefore making the
possible difference in path lengths smaller.
➔ Material - caused by using light consisting of different wavelengths, meaning light
rays will travel at different speeds along the fibre, which leads to pulse broadening
➢ This can be prevented by using monochromatic light.

Both absorption and dispersion can also be reduced by using an optical fibre repeater, which
regenerates the signal during its travel to its destination.
3.4.1 Force, energy and momentum

3.4.1.1 - Scalars and vectors


Scalars and vectors are physical quantities, scalars describe only a magnitude while vectors
describe magnitude and direction. Below are some examples:

Scalars Vectors

Distance, speed, mass, Displacement, velocity, force/weight,


temperature acceleration

There are two methods you can use to add vectors:

Calculation - This should be used when the two vectors are perpendicular.

For example, two forces are acting perpendicular to each other and have magnitudes of 5 N and
12 N. Find the resultant force, and its direction from the horizontal.

To find the resultant vector (R) you can use pythagoras:


122 + 52 = 169 = R2 R = 13 N

In order to find the direction, you can use trigonometry:


5
tan θ = 12
θ = 22.6°

Direction = 22.6° from the horizontal

It is very important to state how the angle you find signifies the direction.

Scale drawing - This should be used when vectors are at angles other than 90°.

For example, a ship travels 30 m at a bearing of 060°, then 20 m east. Find the magnitude and
direction of its displacement from its starting position.

You will need to draw a scale diagram, using a ruler and a protractor as shown on the right. Make
sure to show the scale you are using.
Finally, measure the missing side and convert it to the magnitude using your scale and measure
the missing angle θ, to find the bearing of the displacement.

Magnitude = 4.9 cm = 49 m to scale Direction = 072°

The opposite of adding two vectors is called resolving vectors, and is done using trigonometry. It
is extremely helpful to do this in certain situations because vectors which are perpendicular don’t
affect each other, so can be evaluated separately.

There are formulas to show how to resolve the vector V, into its components x and y.
x = V cos θ y= V sin θ

However, if you struggle with remembering formulas a good hint to remember is:

If you are moving from the original vector through the angle θ to get to your component, use cos.
If you are moving away from the angle θ to get to your component, use sin.

For example, a ball has been fired at a velocity of 10 m/s, at an angle of 30° from the horizontal,
find the vertical and horizontal components of velocity.

x= 10 cos 30° y= 10 sin 30°


= 8.7 m/s = 5 m/s

This next example is much trickier as it involves a plane, however the method is exactly the same:
A block of weight 50N is resting on a plane inclined at 15°. Find the vertical and horizontal
components of weight acting on the block.

Using the fact that the angles in a triangle add up to 180°, you can see that the angle between the
component of weight perpendicular to the plane and weight is also 15°. Using this fact and the
useful hint in red above, you can easily find the components of weight.

Parallel component = 50 sin 15°


(because you are moving away from the angle) = 12.9 N

Perpendicular component = 50 cos 15°


(because you are moving through the angle) = 48.3 N

For an object to be in equilibrium, the sum of all of the forces acting on it must be zero. If an
object is in equilibrium it has no resultant force and therefore it is either at rest or moving at a
constant velocity as according to Newton’s first law.

You can show an object is in equilibrium by either:


● Adding the horizontal and vertical components of the forces acting on it, showing they equal
zero.
● Or if there are 3 forces acting on the object you can draw a scale diagram, if the scale
diagram forms a closed triangle, then the object is in equilibrium.

3.4.1.2 - Moments

The moment of a force about a point is the force multiplied by the perpendicular distance from the
line of action of the force to the point.
Moment = Force X Perpendicular distance to line of action of force from the point

A couple is a pair of coplanar forces (meaning they are forces within the same plane), where the
two forces are equal in magnitude but act in opposite directions.

To find the moment of a couple, you multiply the one of the forces by the perpendicular distance
between the lines of action of the forces.
Moment of a couple = Force X Perpendicular distance between the lines of action of forces
The principle of moments states that for an object in equilibrium, the sum of anticlockwise
moments about a pivot is equal to the sum of clockwise moments.

You can use this fact to answer certain questions, for example:
Find the value of F from the diagram on the right.

Σ clockwise moments = Σ anticlockwise moments

Taking moments around A:


(2 × 50) + (3 × 35) = (3.5 × F )
205 = 3.5F F = 58.6 N

Note, in the example moments are taken about A, as the distance from A to A is 0, the moment
caused by the 12 N force is also 0, therefore it can be ignored.

The centre of mass of an object is the point at which an object’s mass acts.
If an object is described as uniform, its centre of mass will exactly at its centre.

3.4.1.3 - Motion along a straight line

Speed - This is a scalar quantity which describes how quickly an object is travelling.
Displacement (s) - The overall distance travelled from the starting position (includes a direction as
it is a vector quantity).
Δs
Velocity (v) - rate of change of displacement -
Δt
Δv
Acceleration (a) - rate of change of velocity -
Δt
Instantaneous velocity is the velocity of an object at a specific point in time. It can be found from
a displacement-time graph by drawing a tangent to the graph at the specific time and calculating
the gradient.
Average velocity is the velocity of an object over a specified time frame. It can be found by
dividing the final displacement by the time taken.
Uniform acceleration is where the acceleration of an object is constant.

Acceleration-time graphs represent the change in velocity over time. Therefore the area under
the graph is change in velocity.

Velocity-time graphs represent the change in velocity over time. Therefore the gradient of a
velocity time graph is acceleration, and the area under the graph is displacement.
Displacement-time graphs show change in displacement over time, and so their gradient
represents velocity.

When an
object is moving at uniform acceleration, you can use the following formulas:

v = u + at s = ( u+v at2 v 2 = u2 + 2as


2 )t s = ut + 2
Where s = displacement, u = initial velocity, v = final velocity, a = acceleration, t = time

When approaching questions which require the use of these formulas, it is useful to write out the
values you know, and the ones you want to find out in order to more easily choose the correct
formula to use.

For example:
A stone is dropped from a bridge 50 m above the water below. What will be its final velocity (v)
and for how long does it fall (t)?

Note, in this example the stone is dropped therefore we can assume, initial velocity is zero. Also
because the stone is dropped we know its acceleration will be g (9.81 m/s2), which is the
acceleration due to gravity.
s = 50 m u = 0 m/s v=? a = 9.81 m/s2 t=?

2 2
Using v = u + 2as , you can find v.
v 2 = 02 + 2 × 9.81 × 50 v 2 = 981 v = 31.3 m/s

at2
Using s = ut + 2
, you can find t.
2 2
50 = 4.905t t = 10.19 t = 3.19 s

3.4.1.4 - Projectile motion

The vertical and horizontal components of a projectile’s motion are independent, therefore the
projectile’s horizontal and vertical motion can be evaluated separately using the uniform
acceleration formula, where acceleration is constant.

For example:
A ball is projected from the ground at 20 m/s, at an angle of 60° to the horizontal. Find the time
taken for the ball to reach the ground and its maximum height. Ignore the effect of air resistance.

Firstly, you must resolve the initial speed into its components:

Vertical component = 20 sin 60° Horizontal component = 20 cos 60°


= 17.3 m/s = 10 m/s

(To answer this particular question, you only need to consider the vertical component but this is not
always the case).

Maximum vertical height occurs when the vertical component of velocity first becomes 0, therefore:
s=? u = 17.3 m/s v = 0 m/s a = -g (-9.81 m/s2 ) t = ?

2 2
Using v = u + 2as , you can find s.
0 = 17.32 + 2 ×− 9.81 × s 19.62s = 300 s = 15.3 m

Maximum height = 15.3 m


Using v = u + at , you can find t.
0 = 17.3 − 9.81t 9.81t = 17.3 t = 1.76 s

Using the fact that the time for the journey will be double the time to reach the maximum height:
Time to reach ground = 3.5 s

Free fall is where an object experiences an acceleration of g.

Friction is a force which opposes the motion of an object, and it is also known as drag or air
resistance when considering friction experienced in a fluid. Frictional forces convert kinetic energy
into other forms such as heat and sound.

The magnitude of air resistance increases as the speed of the object


increases.

Lift is an upward force which acts on objects travelling in a fluid, it is caused


by the object creating a change in direction of fluid flow, and it acts
perpendicular to the direction of fluid flow.

Terminal speed occurs where the frictional forces acting on an object and
driving forces are equal, therefore there is no resultant force and so no
acceleration so the object travels at a constant speed. A good example of an
object reaching terminal speed, or terminal velocity as it is also known, is a skydiver:
● As they leave the plane they accelerate because their weight is greater than the air
resistance.
● As the skydiver’s speed increases, the magnitude of air resistance also increases. This
continues until the force of weight and air resistance become equal, at which point terminal
velocity is reached.

Air resistance will affect both the vertical and horizontal components of a projectile’s motion as
shown in the diagram below:

As you can see, with air resistance the maximum height is reached earlier, and the vertical and
horizontal distance travelled decreases.
3.4.1.5 - Newton’s laws of motion

➔ Newton’s 1st law - An object will remain at rest or travelling at a constant velocity, until it
experiences a resultant force.
➔ Newton’s 2nd law - The acceleration of an object is proportional to the resultant force
experienced by the object: F = ma where F is the resultant force, m is the object’s mass
and a is its acceleration.
➔ Newton’s 3rd law - For each force experienced by an object, the object exerts an equal
and opposite force.

A free-body diagram, is a diagram which shows all the forces that act on an object, below is an
example:

A free-body diagram will show you how each of the


forces acting on the object compare with each other. In
this example, all the arrows look equal therefore we
know that the car is travelling at a constant velocity.

Here is an example where you would have to use Newton’s 2nd law and a free-body diagram:

Find the acceleration of the ball in the diagram below:

Firstly, you must find the mass (m) of the ball as you are only given the weight.

100
As Weight = mass x g, the mass = 9.81 = 10.2 kg

Next, you must find the resultant force (F).


75 − 10 = 65 N to the right

Finally, you can use F = ma, to find acceleration:


65
65 = 10.2 × a a = 10.2 a = 6.4 m/s2
3.4.1.6 - Momentum

Momentum is the product of mass and velocity of an object.

M omentum = mass × v elocity

Momentum is always conserved in any interaction where no external forces act, which means the
momentum before an event (e.g a collision) is equal to the momentum after. This fact is used to
find the velocity of objects after collisions, for example:

A car with a mass of 500 kg, and a velocity of 4 m/s, collides with a stationary truck with a mass of
1500 kg. The two vehicles join together and move on with a velocity V. Find the value of V.

First find the momentum before the collision.


Total momentum before = (500 × 4) + (1500 × 0) = 2000 kgm/s

Total momentum before = Total momentum after

Therefore, 2000 = 2000 × V V = 1 m/s

Δ(mv) Δv
Newton’s 2nd law states F = ma , therefore, F = Δt as a= Δt . From this you can see that
force is the rate of change of momentum.

Rearranging the above equation leads to F Δt = Δ(mv) (where F is constant, and t is an impact
time).

F Δt is known as impulse, and impulse is the change in momentum as demonstrated in the


equation above.

The area of a force-time graph is F × Δt , therefore it is also equal to change in momentum:


Here is an example question about impulse.

A ball is hit with a baseball bat with a force of 100 N, with an impact time of 0.5 s. What is the
change in momentum of the ball?

To find the impulse we must use the equation F Δt = Δ(mv) .


Change in momentum = 100 × 0.5 = 50 kgm/s

An important application of calculating impulse is during the design of car safety features. For
example, cars have crumple zones, which crumple upon impact, seat belts which stretch upon an
impact, and air bags all of which increases the impact time of the car or the passenger. This
causes the force exerted on passengers to decrease, meaning they are less likely to be seriously
injured.

There are two types of collisions:


● Elastic - where both momentum and kinetic energy are conserved
● Inelastic - where only momentum is conserved, while some of the kinetic energy is
converted into other forms (e.g heat, sound, gravitational potential) and may be larger or
smaller after a collision

If the objects in a collision stick together after the collision, then it is an inelastic collision.

An explosion is another example of an inelastic collision as the kinetic energy after the collision is
greater than before the collision.

3.4.1.6 - Work, energy and power

Work done (W) is defined as the force causing a motion multiplied by the distance travelled in the
direction of the force.
W = F s cos θ where s is the distance travelled and θ is the angle between the direction of the
force and the direction of motion.

As work is a measure of energy transfer, the rate of doing work = the rate of energy transfer.
ΔW F ×Δs
Power (P) is the rate of energy transfer therefore, P = Δt = Δt = Fv as v = Δs
Δt .
For a variable force you can’t use the formula to find work done, however the area under a
force-displacement graph is equal to the work done.

Efficiency is a measure of how efficiently a system transfers energy. It is calculated by dividing the
useful power output by total energy input.

usef ul output power usef ul output power


E f f iciency = input power E f f iciency (percentage) = input power × 100

If you multiply the value of efficiency by 100, you receive the value of efficiency as a percentage.

3.4.1.7 - Conservation of energy

The principle of conservation of energy states that energy cannot be created or destroyed, but
can be transferred from one form to another. Therefore, the total energy is a closed system stays
constant.
T otal energy in = T otal energy out
You may need to use this when answering questions involving gravitational potential energy and
kinetic energy.

Change in gravitational potential energy =


ΔE p = mgΔh

Kinetic energy =
E k = 12 mv 2
As an example, think about a ball being thrown up into the air. The thrower gives the ball kinetic
energy therefore it moves upwards, however as it does, the ball slows down because kinetic
energy is transferred to gravitational potential energy. Eventually, all of the kinetic energy is
transferred to gravitational potential energy and the ball stops momentarily, after which the ball’s
gravitational potential energy is converted back into kinetic energy and the ball falls to the ground.

It is very important to note that work is being done by the ball to work against resistive forces,
therefore the initial kinetic energy given to the ball is not equal to the maximum gravitational
potential when the ball has stopped in mid-air. This is because the kinetic energy of the ball is
being transferred to the environment in the form of heat due to air resistance.

Here is an example of a question where you could use the principle of conservation of energy,
(noting that the effect of air resistance is ignored, therefore ΔE p = ΔE k ):

As a simple pendulum of mass 500g swings, it rises up by a height of 10cm at its maximum
amplitude from its equilibrium position. What is the maximum speed the pendulum can reach
during its oscillation? (Ignore the effect of air resistance)

Firstly, find the maximum gravitational potential energy (which will be at the amplitude).
ΔE P = 0.5 × 9.81 × 0.1 = 0.4905 J

Then, equate this to the kinetic energy formula (subbing in known values), and rearrange to find v.
1 2
2 × 0.5 × v = 0.4905 v 2 = 1.962 v = 1.4 m/s
3.4.2 Materials

3.4.2.1 - Bulk properties of solids

The density of a material is its mass per unit volume, and it’s a measure of how compact a
substance is.

Hooke’s law states that extension is directly proportional to the force applied, given that the
environmental conditions (e.g temperature) are kept constant. This can be shown by the straight
part of the force-extension graph shown to the right; a straight line graph through the origin,
which shows the force and extension are directly proportional.

The limit of proportionality (P) is the point after which Hooke’s law is no longer obeyed. The
elastic limit (E) is just after the limit of proportionality and if you increase the force applied beyond
this, the material will deform plastically (be permanently stretched).

Hooke’s law can be described as the equation F = kΔL , where k is the spring constant, which
is a measure of the stiffness of the spring, and ΔL is the extension.

S tress = F
Tensile stress - Force applied per unit cross-sectional area. A
S train = ΔL
Tensile strain - This is caused by tensile stress, and is L
defined as the extension over the original length.

When work is done on a material to stretch or compress it, this energy is stored as elastic strain
energy. This value cannot be calculated using the formula W = F s cos θ because the force is
variable, however you can find it by calculating the area under a force-extension graph. Therefore,
elastic strain energy = 12 F ΔL .
Breaking stress is the value of stress at which the material will break apart, this value will depend
on the conditions of the material e.g its temperature.
Force-extension graphs can show the properties of a specific object. There are two main
behaviours that a material can exhibit on a force-extension graph:
● Plastic - This is where a material will experience a large amount of extension as the load is
increased, especially beyond the elastic limit
● Brittle - This is where a material will extend very little, and therefore is likely to fracture
(break apart) at a low extension.

Once a material is stretched beyond its elastic limit, a force-extension graph showing loading and
unloading, will not return to the origin, however the loading and unloading lines will be parallel
because the material’s stiffness is constant, as shown on the left. The area between the loading
and unloading line is the work done to permanently deform the material.

When a stretch is elastic (material returns to original shape once force is removed), all the work
done is stored as elastic strain energy, however when a stretch is plastic work is done to move
atoms apart, so energy is not stored as elastic strain energy but is dissipated as heat.
This fact is used when designing safety features for cars, for example work is done to deform
crumple zones plastically to decrease the car’s kinetic energy, and seat belts stretch in order to
convert some of the passenger’s kinetic energy into elastic strain energy.

When a spring is hung vertically and stretched, kinetic energy is converted into elastic strain
energy, and if it the force is removed, the elastic strain energy will be transferred back to kinetic
energy (and in a similar fashion to the ball example in 3.4.1.7 - Conservation of energy), this
kinetic energy is then transferred to gravitational potential energy as it rises.

Stress-strain graphs are similar to force-extension graphs, however they describe the behaviour
of a material rather than the behaviour of a specific object. They show a material’s ultimate tensile
stress (UTS), which is the highest point on the graph as it shows the maximum stress the material
can withstand. Their shape can also show whether a material is ductile (can undergo a large
amount of plastic deformation before fracturing), brittle, or plastic.
3.4.2.2 - Young modulus

The Young modulus is a value which describes the stiffness of a material.


It is known that up to the limit of proportionality, for a material which obeys Hooke’s law, stress is
proportional to strain, therefore the value of stress over strain is constant, this value is the Young
modulus.

T ensile Stress
Y oung M odulus (E) = T ensile strain

Using the formulas from the previous section, this can be rewritten as:

FL
E= ΔLA

You can find the Young modulus of a material from a stress-strain graph by finding the gradient of
the straight part of the graph.
3.5.1 Current electricity

3.5.1.1 - Basics of electricity

Electric current (I) - the flow of charge per unit time, or the rate of flow of charge.
ΔQ
I= Δt (where Q is charge)

Potential difference (V) - the energy transferred per unit charge between two points in a circuit.
W
V = Q (where W is energy transferred)

Resistance (R) - this is a measure of how difficult it is for charge carriers to pass through a
component, and is measured by dividing the potential difference across a component by the
current flowing through it.
V
R= I

3.5.1.2 - Current-voltage characteristics

Ohm’s law states that for an ohmic conductor, current is directly proportional to the potential
difference across it, given that physical conditions (e.g temperature) are kept constant.

You must be able to recognise and understand the properties of certain components as
demonstrated by current-voltage graph (where either current or voltage is on the y-axis):

● Ohmic conductor - this component follows Ohm’s law therefore its current-voltage graph
will look like a straight line through the origin. (This is provided physical conditions are kept
constant).

● Semiconductor diode - when looking at the current-voltage graph of this component you
must consider its forward and reverse bias. The forward bias of a diode is the direction in
which it will allow current to flow easily past the threshold voltage, which is the smallest
voltage needed to allow current to flow. In the direction of the reverse bias, the resistance of
the diode is extremely high meaning that only a very small current can flow.
● Filament lamp - This component contains a length of metal wire, which heats up as
current increases, therefore the resistance of this component increases as current
increases. At low currents the metal wire will not heat up significantly therefore for very low
currents, Ohm’s law is obeyed. However, as the current increases (in either direction), the
graph begins to curve due to the increasing resistance.

Unless a question states otherwise, ammeters can be assumed to have zero resistance, meaning
they will not affect the measurement of current in a circuit at all, and voltmeters can be assumed to
have infinite resistance, meaning no current can flow through them, meaning their measurement of
potential difference across a component is exact.

3.5.1.3 - Resistivity

Resistivity ( ρ ) is a measure of how easily a material conducts electricity, it is defined as the


product of resistance and cross-sectional area, divided by the length of the material. Resistivity will
give the value of resistance through a material of length 1 m and cross-sectional area 1 m2 which is
useful when you need to compare materials even though they may not be the same size, however
resistivity is also dependent on environmental factors, such as temperature.

RA
ρ= L

When the temperature of a metal conductor increases, its resistance will increase. This is
because the atoms of the metal gain kinetic energy and move more, which causes the charge
carriers (electrons) to collide with the atoms more frequently causing them to slow down, therefore
current decreases and so resistance increases (as R = VI ).
However, the opposite is true for thermistors: as the temperature of a thermistor increases, its
resistance decreases. This is because increasing the temperature of a thermistor causes
electrons to be emitted from atoms, therefore the number of charge carriers increases and so
current increases causing resistance to decrease. On the left, is the temperature-resistance graph
of thermistor.

One application of a thermistor in circuits is a temperature sensor, which can trigger an event to
occur once the temperature drops or reaches a certain value. For example, it could be used to turn
on the heating once room temperature drops below a specific value.

A superconductor is a material which, below a certain


temperature, known as the critical temperature, has zero
resistivity. The critical temperature of a superconductor
depends on the material it is made out of, and most known
superconductors have an extremely low critical temperature
which lie close to 0 K (-273°C).

With a resistivity of zero, resistance also drops to zero


therefore applications of superconductors include:
● Power cables, which would reduce energy loss through heating to zero during
transmission.
● Strong magnetic fields, which would not require a constant power source. These could be
used in maglev trains, where there would be no friction between the train and rail, and in
certain medical applications.

3.5.1.4 - Circuits

There are two rules for adding the resistances of resistors in circuits, depending on whether the
resistors are in series or in parallel:

● In series - RT = R1 + R2 + R3 + ... (Where RT is total resistance and Rn is the resistance


of resistor n).
1 1 1 1
● In parallel - R = R + R + R + ...
T 1 2 3

You may need to use both of these rules when calculating the resistance of one circuit, for
example: Find the resistance of the circuit in the diagram to the right.

Firstly, find the resistance of the parallel combinations of resistors:


1 1 1 5 RT = 3.2 Ω
RT = 4 + 16 = 16

Then, use the series rule to add the remaining two resistors to the value calculated for the parallel
combination.
RT = 10 Ω + 3 Ω + 3.2 Ω = 16.2 Ω So the total resistance is 16.2 Ω.
Power (P) is the energy transferred over time (rate of transfer of energy). P = Et where E is
energy transferred and t is time. Another formula for power is P = V I , which can be combined with
the formula V = I R , to form two variations:
2
P =VI = V
R = I 2R
As power is the energy transferred over time, the product of power and time is the energy
transferred therefore E = V It .

Below is an example in which you need to use the above formulas:

A lamp has a power of 60 W, and is connected to a power source of 240 V. Find the energy
transferred by the lamp in 2 minutes and the current in the lamp.

To find the energy transferred you can use the (rearranged) formula E = P t , making sure
time is converted into seconds.

E = 60 × 120 = 7200 J

P
To find the current, you can use the (rearranged) formula I = V .

60
I= 240 = 0.25 A

In a series circuit,
● The current is the same everywhere in the circuit.
● The battery p.d is shared across all elements in the circuit, therefore the total sum of the
voltages across all elements is equal to the supply p.d.
In a parallel circuit,
● The sum of the currents in each parallel set of branches is equal to the total current.
● The potential difference across each branch is the same.

When joining together battery cells, you can use either a series or parallel configuration.
When joined in series, the total voltage across the cells is equal to the sum of the individual
voltages of the cells:
V T = V 1 + V 2 + V 3 + ...

When identical cells are joined in parallel, the total voltage is equal to the voltage of one cell.
This is because the current is split equally between branches, therefore the overall potential
difference is the same as if the total current was flowing through a single cell:
V T = V 1 = V 2 = V 3 = ...

In DC circuits, charge and energy are always conserved. Kirchoff’s two laws describe how this is
achieved:
Kirchoff’s first law - the total current flowing into a junction is equal to the current
flowing out of that junction. This shows that no charge is lost at any point in the circuit.
Kirchoff’s second law - the sum of all the voltages in a series circuit is equal to the
battery voltage. This shows that no energy is lost at any point in a circuit.

3.5.1.5 - Potential divider

A potential divider is a circuit with several resistors in series connected across a voltage source,
used to produce a required fraction of the source potential difference, which remains constant.
You can also make a potential divider supply a variable potential difference by using a variable
resistor as one of the resistors in series, therefore by varying the resistance across it, you can
vary the potential difference output. For example, if the resistance across R1 increases, the output
p.d will decrease as circuit current has decreased and V=IR.

You could replace the variable resistor in the


circuit to the right to a thermistor or light
dependent resistor (LDR), in order to form a
temperature or light sensor. A light dependent resistor’s resistance decreases as light intensity
increases.

These types of sensors can be used to trigger certain events, for example in the circuit above, a
light dependent resistor is used. If the light intensity falls, resistance across R1 will increase so the
circuit current decreases and resistance across R2 decreases, so the p.d out decreases. If you
want this effect to be reversed, switch the position of the LDR and resistor, therefore the p.d out
would increase as light intensity decreased and the circuit could go on to cause a light bulb to be
switched on because a threshold voltage has been met.

3.5.1.6 - Electromotive force and internal resistance

Batteries have an internal resistance (r) which is caused by electrons colliding with atoms inside
the battery, therefore some energy is lost before electrons even leave the battery. It is represented
as a small resistor inside the battery.
Electromotive force (emf / ε ) is the energy transferred by a cell per coulomb of charge that
passes through it: ε = QE
As you can see in the circuit below, the sum of the internal resistance (r), and load resistance (R) is
the total resistance (RT) in the circuit.
RT = R + r

And so emf is the product of the total resistance and the current of the circuit, because V = I R .
ε = IR + Ir ε = I (R + r)
The p.d across the resistor R, is known as the terminal p.d (V), whereas the p.d across the
resistor r, is known as lost volts (v) because this value is equal to the energy wasted by the cell
per coulomb of charge.
V = IR v = Ir

Therefore, emf is the sum of the terminal p.d and lost volts: ε = V + v .

The emf of a battery can be measured by measuring the voltage across a cell using a voltmeter
when there is no current running through the cell, which means it is in an open circuit.

Here are two example questions where you need to use the above formulas:

A cell has an emf of 5 V, the value of lost volts is 2 V, and the resistance of R is 10 Ω. Find the
current in the circuit.

Firstly, write down what you know.


Lost volts = Ir = 2 V Emf = 5 V

Next, find the p.d across R using ε = I R + I r .


5 = IR + 2 IR = 3 V

Using V = I R , find the current.


3 = I × 10 I = 0.3 A

A cell has an emf of 10 V, the current flowing through the circuit is 2 A, the resistance of R1 is 3.5 Ω
and the resistance of R2 is 0.5 Ω. Find the value of lost volts and the internal resistance of the cell.

Firstly, find the terminal p.d (sum of the potential differences across R1 and R2/V).
V = IR V 1 = 2 × 3.5 = 7 Ω V 2 = 2 × 0.5 = 1 Ω V = 7 + 1 = 8V

Find the lost volts using ε = V + v .


10 = 8 + v v = 2V

Finally find internal resistance using v = I r .


2=2×r r = 1Ω
3.6.1 Periodic motion

3.6.1.1 - Circular motion

An object moving in a circular path at constant speed has a constantly changing velocity as velocity
has both magnitude and direction, therefore the object must be accelerating (this is known as
centripetal acceleration). We know from Newton’s first law that to accelerate, an object must
experience a resultant force, therefore an object moving in a circle must experience a force, this is
known as the centripetal force, and it always acts towards the centre of the circle.

Angular speed (ω) is the angle an object moves through per unit time. It can be found by dividing
the object's linear speed (v) by the radius of the circular path it is travelling in (r), or by dividing the
angle in a circle (in radians) by the object’s time period (T).
v 2π 1
ω= r = T = 2πf as f= T

Angles can be measured in units called radians. One radian is defined as the angle in the sector
of a circle when the arc length of that sector is equal to the radius of the circle, as shown in the
diagram below.

Considering a complete circle, its arc length is 2πr , dividing this by r, you get 2π which is the angle
in radians of a full circle. From this you can convert any angle from degrees to radians by
π
multiplying by 180 , and from radians to degrees by multiplying by 180
π .

Image source: User:Stannered. Original image by en:User:Ixphin , CC BY-SA 3.0


Centripetal acceleration (a) can be found using the formula below:
v2
a= r = ω2 r
Using Newton’s second law, F = ma , we can derive the formula for centripetal force (F) from the
formula above.
mv 2
F = r = mω 2 r

3.6.1.2 - Simple harmonic motion (SHM)

An object is experiencing simple harmonic motion when its acceleration is directly proportional
to displacement and is in the opposite direction. These conditions can be shown through the
equation:
a = − ω2 x
Where a is acceleration, ω is angular speed, x is displacement from the equilibrium position

An example of a simple harmonic oscillator is the simple pendulum, as shown in the diagram
above. The pendulum oscillates around a central midpoint known as the equilibrium position.
Marked on the diagram by an x is the measure of displacement, and by an A is the amplitude of
the oscillations, this is the maximum displacement. You could also measure the time period (T) of
the oscillations by measuring the time taken by the pendulum to move from the equilibrium
position, to the maximum displacement to the left, then to the maximum displacement to the right
and back to the equilibrium position.

Using the measurements above, you could use the following formulas:

x = Acos ωt v = ±ω √ (A2 − x2 ) where v is velocity

You can represent the displacement (x), velocity (v) and acceleration (a) of a simple harmonic
oscillator on a graph. From the first equation above you can see that the displacement-time
graph will follow a cosine or sine curve, with a maximum A, and minimum -A because A and ω are
constants:

As we know that velocity is the derivative of displacement, we can draw a velocity-time graph
by drawing the gradient function of the above graph, noting that the maximum (ωA) and
minimum velocity (-ωA)) occurs when x is 0, as expected from the above formula:

Finally, we know acceleration is the derivative of velocity, so we can draw an


acceleration-time graph by drawing the gradient function of the above graph.

As seen above maximum speed = ω A and maximum acceleration = ω 2 A .

3.6.1.3 - Simple harmonic systems

Simple harmonic systems are those which oscillate with simple harmonic motion, examples
include:
● Simple pendulum - A small, dense bob of mass m hangs from a string of length l , which is
attached to a fixed point. When the bob is displaced by a small angle (less than 10°), and
let go it will oscillate with SHM. For this type of system, you can use the following formula:


l
T = 2π g
where T is time period, l is the length of the string, g is acceleration due to gravity

The reason the angle by which the pendulum is displaced must be less than 10°, is
because during the derivation of the above formula a small angle approximation is used,
and so for larger initial angles this approximation is no longer valid, and would not be a
good model.
During the oscillations of a simple pendulum, its gravitational potential energy is transferred
to kinetic energy and then back to gravitational potential energy and so on.

● Mass-spring system - There are two types of mass-spring system, where the spring is
either vertical or horizontal, the only difference between these two is the type of energy
which is transferred during oscillations. For the vertical system, kinetic energy is converted
to both elastic and gravitational potential energy, whereas for the horizontal system, kinetic
energy is converted only to elastic potential energy. For either of these types of system you
can use the following formula:
T = 2π mk where T is time period, m is the mass, k is the spring constant

For any simple harmonic motion system, kinetic energy is transferred to potential energy and
back as the system oscillates, the type of potential energy depends on the system.
At the amplitude of its oscillations the system will have the maximum amount of potential energy,
as it moves towards the equilibrium position, this potential energy is converted to kinetic energy so
that at the centre of its oscillations the kinetic energy is at a maximum, then as the system
moves away from the equilibrium again, the kinetic energy is transferred to potential energy until it
is at a maximum again and this process repeats for one full oscillation. The total energy of the
system remains constant (when air resistance is negligible, otherwise energy is lost as heat).
Image source: Saksun Young,CC BY-SA 4.0

The diagram above shows the variation of energy with displacement, while the diagram below
shows the variation of energy with time, for a simple harmonic system staring at its amplitude.

Damping is where the energy in an oscillating is lost to the environment, leading to reduced
amplitude of oscillations. There are 3 main types of damping:
● Light damping - This is also known as under-damping and this is where the amplitude
gradually decreases by a small amount each oscillation.
● Critical damping - This reduces the amplitude to zero in the shortest possible time (without
oscillating).
● Heavy damping - This is also known as over-damping, and this is where the amplitude
reduces slower than with critical damping, but also without any additional oscillations.

3.6.1.4 - Forced vibrations and resonance

Free vibrations occur when no external force is continuously acting on the system, therefore the
system will oscillate at its natural frequency.
Forced vibrations are where a system experiences an external driving force which causes it to
oscillate, the frequency of this driving force, known as driving frequency, is significant. If the
driving frequency is equal to the natural frequency of a system (also known as the resonant
frequency), then resonance occurs.

Resonance is where the amplitude of oscillations of a system drastically increase due to gaining
an increased amount of energy from the driving force. Resonance has many applications for
example:
● Instruments - An instrument such as a flute has a long tube in which air resonates, causing
a stationary sound wave to be formed.
● Radio - These are tuned so that their electric circuit resonates at the same frequency as
the desired broadcast frequency.
● Swing - If someone pushes you on a swing they are providing a driving frequency, which
can cause resonance if it’s equal to the resonant frequency and cause you to swing higher.

However, resonance can also have negative consequences, such as causing damage to a
structure, for example a bridge when the people crossing it are providing a driving frequency close
to the natural frequency, it will begin to oscillate violently which could be very dangerous and
damage the bridge. Therefore damping can be used to decrease the effect of resonance,
different types of damping will have different effects, as the degree of damping increases, the
resonant frequency decreases (shifts to left on a graph), the maximum amplitude decreases and
the peak of maximum amplitude becomes wider, these effects are shown in the graph below,
where ζ is the damping ratio, ζ = 1 represents critical damping.

Image source: Geek3,CC BY 3.0


3.6.2 Thermal physics

3.6.2.1 - Thermal energy transfer


The internal energy of a body is equal to the sum of all of the kinetic energies and potential
energies of all its particles. The kinetic and potential energies of a body are randomly distributed.

The internal energy of a system can be increased in two ways:


● Do work on the system to transfer energy to it, (e.g moving its particles/changing its
shape).
● Increase the temperature of the system.

When the state of a substance is changed, its internal energy also changes, this is because the
potential energy of the system changes, while the kinetic energy of the system is kept
constant. This can be demonstrated by measuring the temperature of water as it boils:
● The temperature increases up until 100°C, after which the energy gained through heating
the water is no longer used to increase the temperature (and therefore kinetic energy), but
instead is used to break bonds between water molecules so it can change state to water
vapour, and so the potential energy is increased.

Below is a graph showing how the internal energy of a substance varies with temperature:

You can measure the amount of energy required to change the temperature of a substance using
the following formula: Q = mcΔθ Where Q is energy required, m is the mass, c is the specific
heat capacity, and Δθ is the change in temperature.

The specific heat capacity of a substance is the amount of energy required to increase the
temperature of 1 kg of a substance by 1 °C/1 K, without changing its state.

You can measure the amount of energy required to change the state of a substance using the
following formula: Q = ml
Where Q is energy required and l is the specific latent heat.
The specific latent heat of a substance is the amount of energy required to change the state of 1
kg of material, without changing its temperature. There are two types of specific latent heat:
the specific latent heat of fusion (when solid changes to liquid) and specific latent heat of
vaporisation (when liquid changes to gas).

You will need to be able to do calculations using the above formulas and understand
continuous-flow questions, below are some examples:

A kettle has a power of 1200 W, and contains 0.5 kg of water at 22°C, how long will it take for the
water in the kettle to reach 100°C? (specific heat capacity of water = 4200 J/kg°C )
Firstly, you must find the energy required to increase the temperature of the water to 100°C
using Q = mcΔθ .
Q = 0.5 × 4200 × (100 − 22) = 2100 × 78 = 163800 J
Power is the energy transferred over time, therefore to find the value of time taken we must
divide the energy required by the power.
Q Q 163800
P = t →t= P t= 1200 = 136.5 s

An ice cube of mass 0.01 kg at a temperature of 0°C is dropped into a glass of water of mass 0.2
kg, at a temperature of 19°C. What is the final temperature of the water once the ice cube has fully
melted? (specific heat capacity of water = 4200 J/kg°C, specific latent heat of fusion of ice = 334 × 103 J/kg)
Firstly, find the energy required to change the state of the ice.
Q = ml Q = 0.01 × 334 × 103 = 3340 J
Next, you must set up a pair of simultaneous equations to show the energy transfer in the
water and in the ice separately. As we know that the energy transfer is the same in both as
the system is closed, we can equate these values to find the final temperature (T).
For ice: Q = ml + mcΔθ (because the ice changes state and temperature)
Q = 3340 + 0.01 × 4200 × (T − 0) Q = 3340 + 42T
For water: Q = mcΔθ
Q = 0.2 × 4200 × (19 − T ) Q = 15960 − 840T
Set them equal: 3340 + 42T = 15960 − 840T
882T = 12620
T = 14.3 °C

Water flows past an electric heater with a power of 9000W at a rate of 0.5 kg/s. What is the
increase in temperature of the water per second that it flows past the heater? (specific heat capacity
of water = 4200 J/kg°C)
The power of the heater is 9000W, therefore we know 9000 J of energy are transferred
every second, we also know 0.5 kg of water flows past the heater in that second, using
these pieces of information we can find the increase in temperature:
Q 9000
Q = mcΔθ Δθ = mc Δθ = 0.5×4200 = 4.3°C
3.6.2.2 - Ideal gases

The gas laws describe the experimental relationship between pressure (p), volume (V), and
temperature (T) for a fixed mass of gas. They are empirical in nature, meaning they are not
based on theory but arose from observation and experimental evidence. The 3 gas laws you need
to be aware of are:

1. Boyle’s Law -When temperature is constant, pressure and


volume are inversely proportional

2. Charles’ Law -When pressure is constant, volume is directly


proportional to absolute temperature

3. The Pressure Law -When volume is constant, pressure is


directly proportional to absolute temperature

The absolute scale of temperature is the kelvin scale. All equations in thermal physics will use
temperature measured in kelvin (K). A change of 1 K is equal to a change of 1°C , and to convert
between the two you can use the formula:
K = C + 273
Where K is the temperature in kelvin and C is the temperature in Celsius.

Absolute zero (- 273°C ), also known as 0 K, is the lowest possible temperature, and is the
temperature at which particles have no kinetic energy and the volume and pressure of a gas are
zero.
pV
You can combine all the experimental gas laws into one to get T = k where the constant k is
dependent on the amount of gas used measured in moles, therefore you can rewrite the above
pV
equation to get T = nR , where n is the number of moles of gas, and R is the molar gas
constant (8.31 J mol-1 K-1). You can rearrange this further to get:
pV = nRT , which is the ideal gas equation.
1 mole of a substance is equal to 6.02 × 1023 atoms/molecules, so you can convert between the
number of moles (n) and the number of molecules (N) by multiplying the number of moles by
6.02 × 1023 , which is defined as the Avogadro constant (NA).
N
N = n × NA ⇒ n = NA
You can substitute in the above equation into the ideal gas equation to get it in terms of molecules
N RT
rather than moles: pV = N A . You can simplify this further by using the Boltzmann constant (k),
which is equivalent to NR , leading to the equation:
A

pV = N kT
Molar mass is the mass (in grams) of one mole of a substance and can be found by finding the
relative molecular mass, which is (approximately) equal to the sum of the nucleons in a molecule
of the substance.

Work is done on a gas to change its volume when it is at constant pressure, (this is usually done
through the transfer of thermal energy) the value of work done can be calculated using the formula:
W ork done = pΔV
Where p is the pressure and ΔV is the change in volume.

This is especially useful when working with a graph of pressure against volume, as work done is
simply the area under the graph.
3.6.2.3 - Molecular kinetic theory model

Brownian motion is the random motion of larger particles in a fluid caused by collisions with
surrounding particles, and can be observed through looking at smoke particles under a
microscope. Brownian motion contributed to the evidence for the existence of atoms and
molecules.

You can use a simple molecular model to explain each of the gas laws:
● Boyle’s law - Pressure is inversely proportional to volume at constant temperature
E.g If you increase the volume of a fixed mass of gas, its molecules will move further apart
so collisions will be less frequent therefore pressure decreases.
● Charles’s law - Volume is directly proportional to temperature at constant pressure
When the temperature of a gas is increased, its molecules gain kinetic energy meaning
they will move more quickly and because pressure is kept constant (therefore frequency of
collisions is constant) the molecules move further apart and volume is increased.
● Pressure Law - Pressure is directly proportional to temperature at constant volume
When the temperature of a gas is increased, its molecules gain kinetic energy meaning
they will move more quickly, as volume is constant the frequency of collisions between
molecules and their container increases and they collide at higher speeds therefore
pressure is increased.

The above laws are empirical in nature, meaning they are not based on theory but arose from
observation and experimental evidence. The kinetic theory model however is the opposite and
arose from only theory.

The kinetic theory model equation relates several features of a fixed mass of gas, including its
pressure, volume and mean kinetic energy. There are several underlying assumptions, which lead
to the derivation of this equation, these assumptions and the derivation are outlined below.
Assumptions -
● No intermolecular forces act on the molecules
● The duration of collisions is negligible in comparison to time between collisions
● The motion of molecules in random, and they experience perfectly elastic collisions
● The motion of the molecules follows Newton’s laws
● The molecules move in straight lines between collisions

Derivation -
1. First, you must consider a cube with side lengths l, full of gas molecules. One of these
molecules, has a mass m and is travelling towards the right-most wall of the container, with a
velocity u. Assuming it collides with this wall elastically, its change in momentum is
mu − (− mu) = 2mu .

2. Before this molecule can collide with this wall again it must travel a distance of 2l. Therefore the
2l
time between collisions is t, where t= u .
3. Using these two bits of information we can find the impulse, which is the rate of change of
momentum of the molecule. As impulse is equal to the force exerted, we can find pressure by
2
dividing our value of impulse by the area of one wall: l .
2mu mu2 mu2 mu2 mu2
F = 2l = l
P = l = l3
= V
u l2
As shown, the above equation can be further simplified because l3 is equal to the cube’s volume (V).

4. The molecule we have considered is one of many in the cube, the total pressure of the gas will be
the sum of all the individual pressures caused by each molecule.
m((u1 )2 +(u2 )2 +...+(un )2 )
P = V
5. Instead of considering all these speeds separately, we can define a quantity known as mean
square speed, which is exactly what it sounds like, the mean of the square speeds of the gas
molecules. This quantity is known as u2 , and we multiply it by N, the number of particles in the gas, to
get an estimate of the sum of the molecules’ speeds.

N mu2
P = V
6. The last step is to consider all the directions the molecules will be moving in. Currently we have only
considered one dimension, however the particles will be moving in all 3 dimensions. Using pythagoras’
theorem we can work out the speed the molecules will be travelling at:

c 2 = u2 + v 2 + w 2
Where u, v, and w are the components of the molecule’s velocity in the x, y and z directions.
As the motion of the particles is random we can assume the mean square speed in each direction is the
same.

u2 = v 2 = w 2 ∴ c 2 = 3u 2
The last thing to do now it put this into our equation and rearrange:

pV = 31 N mc2 or pV = 31 N m(crms )2
As c2 and (crms )2 are equivalent

An ideal gas follows the gas laws perfectly, meaning that there is no other interaction other than
perfectly elastic collisions between the gas molecules, which shows that no intermolecular
forces act between molecules. As potential energy is associated with intermolecular forces, an
ideal gas has no potential energy, therefore its internal energy is equal to the sum of the kinetic
energies of all of its particles.

There are several equations which allow you to find the kinetic energy of a single gas molecule,
therefore these can be used to find the internal energy of an ideal gas:
1 2
2
m(c rms ) = 32 kT = 3RT
2N A
As you can see from the middle equation, the kinetic energy of a gas molecule is directly
proportional to temperature (in Kelvin).

Below is an example question using the above equations, as well as knowledge of molar mass and
molecular mass from the previous section:

A bottle contains 128 g of oxygen at a temperature of 330 K. Find the sum of the kinetic energies
of all the oxygen molecules. (Molecular mass of oxygen gas = 32 g)
Firstly, find the number of moles of gas, then multiply this by the avogadro constant to find
the number of molecules.
mass 128
Number of moles = molar mass = 32 = 4
Number of molecules = 4 × 6.02 × 1023 = 2.408 × 1024
3
Then, use 2 kT to find the kinetic energy of one molecule and then multiply this by the
number of molecules:
Kinetic energy of a single molecule = 3
2 × 1.38 × 10−23 × 330 = 6.831 × 10−21
Sum of kinetic energies = 6.831 × 10 −21
× 2.408 × 1024 = 16450 J

Knowledge and understanding of gases has changed greatly over time; the gas laws were
discovered by a number of scientists and later explained by the development of the kinetic theory
model, however this model wasn’t accepted at first. Knowledge and understanding of any scientific
concept changes over time in accordance to the experimental evidence gathered by the scientific
community.
3.7.1 Fields

A force field is an area in which an object experiences a non-contact force. Force fields can be
represented as vectors, which describe the direction of the force that would be exerted on the
object, from this knowledge you can deduce the direction of the field. They can also be
represented as diagrams containing field lines, the distance between field lines represents the
strength of the force exerted by the field in that region.

Force fields are formed during the interaction of masses, static charge or moving charges.
Different types of fields are formed depending on which interaction takes place:
● Gravitational fields - formed during the interaction of masses
● Electric fields - formed during the interaction of charges

The forces these fields exert have a number of similarities and differences:

Similarities Differences

Forces both follow an inverse-square law In gravitational fields, the force exerted is
always attractive, while in electric fields the
Use field lines to be represented force can be either repulsive or attractive.

Both have equipotential surfaces Electric force acts on charge, while


gravitational force acts on mass.

3.7.2 Gravitational fields

3.7.2.1 - Newton’s law


Gravity acts on any objects which have mass and is always attractive.

Newton’s law of gravitation shows that the magnitude of the gravitational force between two
masses is directly proportional to the product of the masses, and is inversely proportional to
the square of the distance between them, (where the distance is measured between the two
centres of the masses).
Gm1 m2
F = r2
Where G is the gravitational constant, m1/m2 are masses and r is the distance between the centre
of the masses.

Image source:
I, Dennis Nilsson,CC BY 3.0
3.7.2.2 - Gravitational field strength
There are two types of gravitational field; a uniform field or radial field. These can be represented
as the following field lines:

The arrows on the field lines show the direction that a force acts on a mass. A uniform field exerts
the same gravitational force on a mass everywhere in the field, as shown by the parallel and
equally spaced field lines. In a radial field the force exerted depends on the position of the
object in the field, e.g in the diagram above, as an object moves further away from the centre, the
magnitude of the force would decrease because the distance between the field lines increases.
The Earth’s gravitational field is radial, however very close to the surface it is almost completely
uniform.

Gravitational field strength (g) is the force per unit mass exerted by a gravitational field on an
object. This value is constant in a uniform field, but varies in a radial field. There are two formulas
you can use to calculate this; the first is general, while the second is used only for radial fields:

F GM
g= m g= r2 (For radial fields only)

3.7.2.3 - Gravitational potential

Gravitational potential (V) at a point is the work done per unit mass when moving an object
from infinity to that point. Gravitational potential at infinity is zero, and as an object moves from
infinity to a point, energy is released as the gravitational potential energy is reduced, therefore
gravitational potential is always negative.
V = − GM r
(For a radial field)
Where M is the mass of the object causing the field, r is the distance between the centres of the objects.

The gravitational potential difference( ΔV ) is the energy needed to move a unit mass
between two points and therefore can be used to find the work done when moving an object in a
gravitational field.

W ork done = mΔV Where m is the mass of the object moved.


Equipotential surfaces are surfaces which are created through joining points of equal potential
together, therefore the potential on an equipotential surface is constant everywhere. As these
points all have equal potential, the gravitational potential difference is zero when moving along the
surface, so no work is done when moving along an equipotential surface. The red lines on the
diagram to the right represent equipotential surfaces.

As shown by the equation for gravitational potential above, gravitational potential (V) is
inversely proportional to the distance between the centres of the two objects (r). This can be
represented on a graph of potential (V) against distance r:

Gravitational field strength (g) at a certain distance can be measured by drawing a tangent to
the curve at that distance and calculating its gradient, then multiplying by -1:

−ΔV
g= Δr

If you plot a graph of gravitational field strength (g)


against distance (r), you can find the gravitational
potential difference by finding the area under the
curve.
3.7.2.4 - Orbits of planets and satellites

Kepler’s third law is that the square of the orbital period (T) is directly proportional to the cube
of the radius (r): T 2 ∝ r3 . This can be derived through the following process:

1. When an object orbits a mass, it experiences a gravitational force towards the centre of the
mass, and as the object is moving in a circle, this gravitational force acts as the centripetal
force. Therefore we can equate the equations of centripetal force and gravitational force:
2 GM m
C entripetal f orce = mv Gravitational f orce =
r r2
mv 2 GM m
r = r2

2. Rearrange the equation to make it v 2 the subject.


GM
v2 = r

3. Velocity is the rate of change of displacement, therefore you can find v in terms of radius (r) and
orbital period (T):
2πr 4π 2 r2
v= T ⇒ v2 =
T2
Because the circumference
diameter of a circle is 2πr, and the object will travel this distance in one orbital period.

4. Substitute the equation for v² in terms of r and T, into the original equation (from step 2).
2 2
4π r GM
T2
= r

5. Rearrange to make T² the subject.


2 4π 2 3
T = GM × r
2
4π 2
As GM is a constant, this shows that T ∝ r3 (T
2
is directly proportional to r3 ).

The total energy of an orbiting satellite is made up of its kinetic and potential energy, and is
constant. For example, if the height of a satellite is decreased, its gravitational potential energy will
decrease, however it will travel at a higher speed meaning kinetic energy increases, therefore total
energy is always kept constant.

T otal energy of a satellite = kinetic energy + potential energy


The escape velocity of an object is the minimum velocity it must travel at, in order to escape
the gravitational field at the surface of a mass. This is the velocity at which the object’s kinetic
energy is equal to the magnitude of its gravitational potential energy.

Ek = Ep
1 2 GM m
2 mv = As gravitational potential energy = mΔV
r

⇒v =
√ 2GM
r
From the above equation, you can see that the escape velocity is independent of the mass of the
object.

A synchronous orbit is one where the orbital period of the satellite is equal to the rotational
period of the object that it is orbiting, for example a synchronous satellite orbiting Earth would
have an orbital period of 24 hours.

Geostationary satellites follow a specific geosynchronous orbit, meaning their orbital period is 24
hours and they always stay above the same point on the Earth, because they orbit directly
above the equator. These types of satellites are very useful for sending TV and telephone
signals because it is always above the same point on the Earth so you don’t have to alter the
plane of an aerial or transmitter.

In order to find the orbital radius of a geostationary satellite you can use the relationship that was
derived above:

GM T 2
T2 = 4π 2
GM
× r3 ⇒ r3 = 4π 2

3 6.67×10−11 ×5.97×1024 ×(24×60×60)2


r = 4π 2
= 7.53 × 1022 m

r = 4.2 × 107 m
Which is around 36 000 km above the Earth’s surface.

Low-orbit satellites have significantly lower orbits in comparison to geostationary satellites,


therefore they travel much faster meaning their orbital periods are much smaller. Because of
this, these satellites require less powerful transmitters and can potentially orbit across the entire
Earth’s surface, this makes them useful for monitoring the weather, making scientific
observations about places which are unreachable and military applications. They can also be
used for communications but because they travel so quickly, many satellites must work together to
allow constant coverage for a certain region.
3.7.3 Electric fields

3.7.3.1 - Coulomb’s law

Coulomb’s law states that the magnitude of the force between two point charges in a vacuum is
directly proportional to the product of their charges, and inversely proportional to the
square of the distance between the charges.
1 Q1 Q2
F = 4πε0 r2
Where ε0 is the permittivity of free space, Q1/Q2 are charges, r is the distance between charges.
Air can be treated as a vacuum when using the above formula, and for a charged sphere,
charge may be assumed to act at the centre of the sphere.

Image source: Dna-Dennis,CC BY 3.0, Image is cropped

If charges have the same sign the force will be repulsive, and if the charges have different signs
the force will be attractive.

The magnitude of electrostatic forces between subatomic particles is magnitudes greater


than the magnitude of gravitational forces, this is because the masses of subatomic particles
are incredibly small whereas their charges are much larger.

For example, consider the gravitational and electrostatic force between two protons, with centres 2
pm ( 2 × 10−12 m) apart.

GM m
● Gravitational force, F = r2 :
−11 −27 −27
6.67×10 ×1.67×10 ×1.67×10
F = (2×10−12 )2
= 4.65 × 10−41 N
1 Q 1 Q2
● Electrostatic force, F = 4πε0 r2
:
−19 −19
1 1.6×10 ×1.6×10
F = 4π×8.85×10−12
× (2×10−12 )2
= 5.75 × 10−5 N
F
● Ratio of electric force over gravitational force, F electric
gravitational
−5
5.75×10
4.65×10−41
= 1.24 × 1036

36
Therefore the electrostatic force between the two protons is 1.24 × 10 times greater
than the gravitational force.
3.7.3.2 - Electric field strength

Electric field strength (E) is the force per unit charge experienced by an object in an electric field.
This value is constant in a uniform field, but varies in a radial field. There are three formulas you
can use to calculate this value; the first is general, the second is used to find the magnitude of E in
a uniform field formed by two parallel plates, while the third is used only for radial fields:

E= F
E= V 1 Q
Q d (for uniform fields formed by parallel plates) E= 4πε0 r2 (for radial fields)

Where V is the potential difference across the plates, d is the distance between the plates, ε0 is the
permittivity of free space, Q is the magnitude of charge, r is the distance between charges.

Like gravitational fields, electric fields can be uniform or radial and can also be represented by the
following field lines:

The field lines show the direction of the force acting on a positive charge. A uniform field exerts
that same electric force everywhere in the field, as shown by the parallel and equally spaced field
lines, whereas in a radial field the magnitude of electric force depends on the distance between
the two charges.

You can derive an equation to calculate the work done by moving a charged particle between
the parallel plates of a uniform field using the equations for electric field strength defined above:

Work done = F × d
Rearrange E = FQ to get F as the subject. F = EQ
ΔV ΔV
Rearrange E = d to get d as the subject. d= E
Substitute the above values into the general formula for work done:
Work done = E Q × ΔV
E Work done = QΔV

Uniform electric fields made by two parallel plates can sometimes be used to find out whether a
particle is charged, and whether its charge is negative or positive. This is done by firing the particle
at right angles to the field and observing its path: a charged particle will experience a constant
electric force either in or opposite to the direction of the field (depending on its charge), this
causes the particle to accelerate and so it follows a parabolic shape. If the charge on the particle
is positive it will follow the direction of the field, if the charge is negative it will move opposite to the
direction of the field.
For example, in the diagram to the left, the
particle must have a positive charge as it follows
a parabolic shape in the direction of the field.

3.7.3.3 - Electric potential

Absolute electric potential (V) at a point is the potential energy per unit charge of a positive
point charge at that point in the field. The absolute magnitude of electric potential is greatest at
the surface of a charge, and as the distance from the charge increases, the potential decreases
so electric potential at infinity is zero. To find the value of potential in a radial field you can use
1 Q
the formula: V = 4πε0 r
Where ε0 is the permittivity of free space, Q is the charge, r is the distance from the charge.

Whether the value of potential is negative or positive depends on the sign of the charge (Q),
when the charge is positive, potential is positive and the charge is repulsive, when the charge
is negative, potential is negative and the force is attractive (similarly to gravitational potential
which is always attractive).

The gradient of a tangent to a potential (V) against distance (r) graph, gives the value of
electric field strength (E) at that point:

ΔV
E= Δr
Electric potential difference ( ΔV ) is the energy needed to move a unit charge between two
points. Therefore, the work done ( ΔW ) in moving a charge across a potential difference is equal
to the product of potential difference and charge.

ΔW = QΔV

Electric fields have equipotential surfaces, just like gravitational fields. The potential on an
equipotential surface is the same everywhere, therefore when a charge moves along an
equipotential surface, no work is done. Between two parallel plates the equipotential surfaces
are planes which are equally spaced and parallel to the plates, whereas equipotential surfaces
around a point charge form concentric circles.

Image source: Balajijagadesh,CC BY-SA 4.0

If you plot a graph of electric field strength (E) against distance (r), you can find the electric
potential difference by finding the area under the graph.
● Equipment:
○ Signal generator
○ Vibration generator
○ Stand
○ Pulley
○ Wooden bridge
○ 100g masses with holder
○ Metre ruler
○ 1.5m long string
○ Balance
● Method:
○ Set up the apparatus as shown in the diagram.
○ Adjust the length l so that it is 1.000m, measured using the metre ruler.
○ Increase the frequency f until the string oscillates at the first harmonic. Read and
record f.
○ Reduce l by 0.100m and adjust f again until it oscillates at the first harmonic.
Measure and record f and repeat this, reducing l by 0.100m each time down to
0.500m.
○ Repeat the experiment twice more and find and record the mean f for each l.
○ Measure the mass of the string on the balance and record it.
● Graphs and calculations:
○ Plot a graph of the mean value of f against 1/l and draw a line of best fit. The wave
speed will be two times the gradient.
2f
○ λ = 2l ⇒ v = 2f l = 1 = 2G where G is the gradient.
l
○ The tension of the string is equal to the weight of the hanging mass (if 100g,
0.981N) and its mass per unit length can be found by dividing the mass of the string
by its length (1.5m).


T which can be compared to the
○ The speed of the wave is also given by v= μ
value obtained by the graph.
● Safety:
○ The stand could topple over and cause injury so a counterweight can be used if it is
deemed unstable.
● Improvements and notes:
○ The experiment can be repeated with different masses to change the tension and
different thicknesses of string to change the mass per unit length in order to
investigate the effect of changing these parameters.
○ An oscilloscope can be used to verify the signal generator’s readings.
○ The signal generator should be left for about 20 minutes to stabilise.
Double slit experiment:
● Equipment:
○ Laser
○ Double slit
○ White screen
○ Metre ruler
● Method:
○ Ensuring that the
laboratory is
partially darkened, set up the apparatus as shown in the diagram. The single slit
might not be required if the laser beam is wide enough to illuminate both the slits of
the double slit.
○ Adjust the apparatus so that D is 0.500m, measured using the metre ruler.
○ Measure across a large number of fringes using the ruler and divide by the number
of fringe widths to find the fringe width w.
○ Increase D by 0.100m and repeat this, increasing it by 0.100m each time up to
1.500m.
○ Repeat the experiment twice more and find and record the mean w for each D.
○ If it is not quoted on the double slit, the slit separation must be measured using a
vernier calliper or a travelling microscope.
● Graphs and calculations:
○ Plot a graph of w against D and draw a line of best fit. The wavelength of the laser
light will be the gradient multiplied by the slit separation.
λD ws
○ w= s ⇒λ= D
= Gs where G is the gradient.
● Safety:
○ Shining a laser into someone’s eyes can be dangerous, so ensure that the lasers
are pointed away from them.
● Improvements and notes:
○ Make sure that the screen and double slit are aligned perfectly normal to the laser to
avoid parallax error in the measurement of the fringe width. A set square can be
used in conjunction with the ruler.
Diffraction grating experiment
● Equipment:
○ Laser
○ Diffraction grating
○ White screen
○ Metre ruler
● Method:
○ Ensuring that the
laboratory is partially
darkened, set up the
apparatus as shown in
the diagram.
○ Adjust the apparatus so that D is 1.000m, measured using the metre ruler.
○ Measure the distances h1 and h2 on either side of the central maximum using the
ruler and find the mean of each of them.
● Graphs and calculations:
○ Work out the slit separation by finding the reciprocal of the number of slits per metre
quoted on the diffraction grating.
○ Find θ1 and θ2 by using θn = arctan( hDn ) .
○ Find λ from both of these angles using λ=dsin(θ)/n and then calculate the mean.
● Safety:
○ Shining a laser into someone’s eyes can be dangerous, so ensure that the lasers
are pointed away from them.
● Improvements and notes:
○ Make sure that the screen and double slit are aligned perfectly normal to the laser to
avoid parallax error in the measurement of the fringe width. A set square can be
used in conjunction with the ruler.
● Equipment:
○ Stand and clamp
○ Electromagnet
○ Steel ball bearing
○ Light gate
○ Stopwatch
● Method:
○ Set up the apparatus as
shown in the diagram.
○ The position of the lower
light gate should be adjusted
such that the height h is
0.500m, measured using the
metre ruler.
○ Turn on the electromagnet and attach the ball bearing. Reset the stopwatch to zero
and switch off the electromagnet. Read and record the time t on the stopwatch.
○ Reduce h by 0.050m by moving the lower light gate upwards and repeat this,
reducing h by 0.050m each time down to 0.250m.
○ Repeat the experiment twice more and find and record the mean t for each h.
● Graphs and calculations:
○ Plot a graph of 2h/t against t and draw a line of best fit. The gradient will be g.
○ h = ut + 12 gt2 ⇒ 2h
t = 2u + g t
○ The y intercept is 2u, where u is the speed of the ball bearing when it reaches the
top light gate (which we don’t need).
● Safety:
○ Use a counterweight or clamp the stand to the table to avoid it toppling over and
causing injury.
● Improvements and notes:
○ The distance between the upper light gate and the starting position of the ball
bearing must be kept constant so that it reaches the upper light gate with the same
speed each time.
○ The ball bearing can be cushioned by a pad which is inside a tray at the bottom of
the clamp so it does not fall off dangerously.
○ The ball bearing should be dense to help mitigate the effects of air resistance.
○ To reduce parallax error in measuring the height, the ruler can be clamped directly
next to the light gates.
● Equipment:
○ 2 1.5m long steel wires
○ Main scale and vernier scale
○ 1kg masses and 2 1kg holders
○ Micrometer
○ Metre ruler
● Method:
○ Set up the apparatus as shown in the diagram.
○ Measure the initial length l of the test wire with the
metre ruler.
○ Add a 1kg mass holder to both wires so they are taut
and record the initial scale reading.
○ Add an additional 1kg mass to the test wire and
record the new scale reading. Find its extension e by
subtracting the initial scale reading from this and
record it.
○ Add another 1kg mass and repeat this, adding 1kg
each time up to around 8kg.
○ Repeat the experiment twice more and find and
record the mean e for each m, where m is the mass
of the 1kg masses on the test wire’s holder.
○ Measure the diameter d of the test wire at various points along it using the
micrometer and find and record the mean diameter.
● Graphs and calculations:
2
πd
○ Calculate the cross-sectional area A of the wire by A= 4
○ Find the force F on the test wire for each m by calculating mg and tabulate this.
○ Plot a graph of F against e and draw a line of best fit. The young modulus E will be l
multiplied by the gradient divided by A.
stress F /A Fl lG
○ E= strain = e/l = Ae = A where G is the gradient.
● Safety:
○ The wire will be stretched very tightly and could break and injure eyes, so safety
goggles must be worn.
○ If the wire breaks, the masses could fall and cause injuries, so a sand tray should be
placed beneath them to catch them.
● Improvements and notes:
○ The comparison wire compensates for sagging of the beam and thermal expansion
effects and provides a reference point against which to measure the extension.
○ The original length l of the test wire should be as long as possible to reduce
uncertainty in its measurement.
● Equipment:
○ 1m long constantan wire
○ Voltmeter
○ Ammeter
○ Low voltage power supply
○ Micrometer
○ Metre ruler
● Method:
○ Measure the diameter d of
the constantan wire at
various points along it
using the micrometer and
find and record the mean
diameter.
○ Set up the apparatus as shown in the diagram.
○ Adjust the length l to 0.100m using the crocodile clips, measured using the metre
ruler.
○ Read and record the current I on the ammeter and the voltage V on the voltmeter.
Calculate the resistance R by using R=V/I and record this.
○ Increase l by 0.100m and repeat this, increasing it by 0.100m each time up to
0.800m.
○ Repeat the experiment twice more and find and record the mean R for each l.
● Graphs and calculations:
2
πd
○ Calculate the cross-sectional area A of the wire by A= 4
○ Plot a graph of the mean value of R against l and draw a line of best fit. The
resistivity will be the gradient multiplied by the cross-sectional area of the wire.
RA
○ ρ= l = GA where G is the gradient.
● Safety:
○ Disconnect the crocodile clips in between measurements to avoid the wire heating
up and causing burns if touched. If the current rises too high, reduce the voltage
using the variable power supply.
○ If the wire is tight, safety goggles should be worn in case it snaps and injures eyes.
● Improvements and notes:
○ The wire heating up might additionally cause the resistance of the wire to change,
affecting measurements. To reduce this, disconnect it in between measurements or
reduce the voltage of the supply so the current is lower.
○ The wire should be free from kinks and held straight so the measurement of the
length is as accurate as possible.
● Equipment:
○ Battery or cell
○ Voltmeter
○ Ammeter
○ Variable resistor
○ Switch
● Method:
○ Set up the apparatus as shown in the diagram.
○ With the switch open, record the reading V on the
voltmeter.
○ Set the variable resistor to its maximum value,
close the switch and record V and the reading I on the ammeter. Open the switch
between readings.
○ Decrease the resistance of the variable resistor and repeat this, obtaining pairs of
readings of V and I over the widest possible range.
● Graphs and calculations:
○ Plot a graph of V against I and draw a line of best fit. The y-intercept will be the emf
and the gradient will be the negative internal resistance.
○ ε = I (R + r) = V + I r ⇒ V =− rI + ε
● Safety:
○ Another resistor can be included in series with the other to avoid high currents which
could be dangerous and make the wires get hot.
● Improvements and notes:
○ Only close the switch for as long as it takes to take each pair of readings. This will
prevent the internal resistance of the battery or cell from changing during the
experiment.
○ Use fairly new batteries/cells because the emf and internal resistance of run down
batteries can vary during the experiment.
Mass-spring system
● Equipment:
○ Spring
○ 50g masses with 50g holder
○ Stand and clamp
○ Pin and blu-tack
○ Metre ruler
○ Stopwatch
● Method:
○ Set up the apparatus as shown
in the diagram, with no masses slotted on the 50g holder.
○ Pull the mass hanger vertically downwards a few centimetres and release it. Start
the stopwatch when it passes the fiducial marker (pin and blu-tack at the centre).
○ Stop the stopwatch after 10 complete oscillations and record this time T 10 . Divide
T 10 by 10 to find the time period T of the mass-spring system and record this.
○ Add a 50g mass to the 50g holder and repeat this, adding 50g each time up to 500g,
recording the total hanging mass m and corresponding time period T for each.
○ Repeat the experiment twice more and find and record the mean T for each m.
● Graphs and calculations:
○ Plot a graph of T^2 against m and draw a line of best fit. The gradient will be 4π^2
divided by the spring constant.

⇒ T2 = 4π 2

m
○ T = 2π k k m
● Safety:
○ Suspended masses could be dangerous if the masses fall off and injure someone.
To avoid this, only pull down the spring by a few centimetres and don’t attach too
heavy masses to the spring.
● Improvements and notes:
○ If the spring starts to move horizontally during its oscillation, stop the oscillation and
start it again making sure it is pulled vertically downwards.
○ Timing more oscillations for each mass reduces the percentage uncertainty in the
time period.
○ The fiducial marker should be at the centre of the oscillation (equilibrium position) so
the mass is moving past it at the fastest speed and there is the least uncertainty in
starting and stopping the stopwatch.
○ A motion tracker and data logger can be used to find a more accurate value for the
time period and eliminating random error in starting and stopping the stopwatch.
Simple pendulum
● Equipment:
○ Pendulum bob on 2m long string
○ Stand and clamp
○ Pin and blu-tack
○ Metre ruler
○ Stopwatch
○ Two wooden blocks
● Method:
○ Set up the apparatus as shown in the diagram.
○ Adjust the point of suspension so that the distance L
from that point to the centre of mass of the pendulum
bob is 1.500m.
○ Pull the pendulum to the side and release it so that it has a small amplitude and
travels in a straight line. Start the stopwatch when it passes the fiducial marker (pin
and blu tack at the centre).
○ Stop the stopwatch after 10 complete oscillations and record this time T 10 . Divide
T 10 by 10 to find the time period T of the pendulum and record this.
○ Decrease the length L by 0.100m and repeat this, reducing L by 0.100m each time
down to 0.500m.
○ Repeat the experiment twice more and find and record the mean T for each L.
● Graphs and calculations:
○ Plot a graph of T^2 against L and draw a line of best fit. The gradient will be 4π^2
divided by g.

⇒ T2 = 4π 2

L
○ T = 2π g g L
● Safety:
○ No notable risks.
● Improvements and notes:
○ It is recommended to use a small pendulum bob so that it is easier to measure the
length to its centre of mass. Additionally, you could measure first the length of the
string and then add the radius of the bob onto this to find L.
○ The angle between the string and the downwards vertical at the maximum amplitude
should be no more than about 15° for the equations to work.
○ Like the spring experiment, a motion tracker and data logger can be used.
Boyle’s law
● Equipment:
○ Stand and clamp
○ Syringe
○ Rubber tubing
○ Pinch clip
○ String
○ 100g masses with 100g holder
● Method:
○ With the plunger removed from the
syringe, measure the inside diameter d of
the syringe using a vernier caliper.
○ Replace the plunger and draw in about
4.0ml of air and record this. Fit the rubber
tubing over the nozzle and clamp it with
the pinch clip as close to the nozzle as
possible.
○ Set up the apparatus as shown in the diagram, with only the 100g holder and one
100g mass suspended.
○ Gently move the plunger up and down to ensure it is not sticking and release it.
Record the new volume V on the syringe scale.
○ Add two 100g masses to the holder and repeat this, adding two 100g masses each
time until the total mass is 1000g.
○ Repeat the experiment twice more and find and record the mean V for each m.
● Graphs and calculations:
2
πd
○ Calculate the cross sectional area A of the syringe using A= 4
○ Calculate and tabulate the force F exerted by each mass m using F=mg.
○ Find the pressure exerted by this force using F/A. Subtract this from standard
atmospheric pressure, 101kPa, to obtain the pressure P of the air sample at each V.
○ Plot a graph of 1/V against P and draw a line of best fit. A straight line through the
origin should be obtained, showing that the pressure is inversely proportional to the
volume.
● Safety:
○ The stand could topple over and cause injury so a counterweight can be used if it is
deemed unstable.
● Improvements and notes:
○ The clamp should be high enough that it does not distort the syringe barrel and
make it more difficult for the plunger to move freely.
○ The syringe can be lubricated to prevent the plunger from sticking.
Charles’ law
● Equipment:
○ Capillary tube
○ Sulfuric acid
○ 2 litre beaker
○ 2 elastic bands
○ 30cm ruler
○ Thermometer
○ Kettle
● Method:
○ Set up the apparatus as shown in the diagram with
the open end of the capillary tube at the top and
add hot water from the kettle. The hot water
should cover the air sample.
○ Stir the water well using the thermometer
and read and record the value of its
temperature, θ, and the length of the air
sample, l, on the 30cm ruler.
○ Allow the water to cool by 5°C and repeat
this, taking measurements every 5°C down
to room temperature.
● Graphs and calculations:
○ Plot a graph of l against θ, draw a line of best fit, and find the gradient m. The value
l1
of absolute zero can be found by θ0 = θ1 − l
m , where θ 1 and 1 are pair of values
on the line of best fit.
○ l1 = mθ1 + c ⇒ c = l1 − mθ1
−c −(l1 −mθ1 ) l1
At absolute zero, l=0, so 0 = mθ0 + c ⇒ θ0 = m = m = θ1 − m
● Safety:
○ Safety goggles must be worn because concentrated sulfuric acid can cause damage
to the eyes.
○ Boiling water is being used which could cause burns, so care must be taken it does
not spill.
● Improvements and notes:
○ The tube needs to be perfectly clean with no traces of other chemicals to prevent
the thread of sulfuric acid from splitting.

You might also like