Download as pdf or txt
Download as pdf or txt
You are on page 1of 42

�����������������������

��������������������������������������������������
Carlo Margio 2017-05-18

Lecture 1

����������
At the start of the 20th century (1900’s) some physicists thought that all the important physics prob-
lems had been solved. The fields of mechanics, gravitation, optics, and electromagnetism seemed
complete. The future of physics, they thought, would be the applying the existing theories to any
problems that came up or doing ever more accurate calculations. Physics up to this point is now
known as Classical Physics.
They had some justification for believing that physics was complete. Most of what we experience
relies on a few important effect. These include:
◼ dynamics or the motion of all matter — by applying Newton’s laws
◼ The interaction of electrons with each other — this accounts for the push between bodies in
contact such as your elbow on the desk.
◼ The passage of light and other electromagnetic (EM) waves through space and its interaction
with matter — this allows us to see things and to receive warmth and energy from the Sun.
◼ The force of gravity that provides the attraction of all matter to other matter — this accounts for
the motion of the planets and the force that keeps us on Earth.
◼ The behaviour gases and materials when heated — statistical mechanics.
It was a shock when experiments showed that Classical Physics described none of these effects
correctly!
Modern physics was born.
2 ��� Q.nb

���������������������
������������������

�������������������������������������������������������
Wolfson 34.2

���������������������������
In the 1800’s it was found that if you heat a body it emits electromagnetic radiation or electromag-
netic waves. Gustav Kirchhoff (1824 – 1887) among others studied these emissions and found that
for a black object the colour distribution of the light emitted depended only on the temperature of the
body.

An electric stove element for instance approximates a black body. It glows dull red at lower settings
and bright orange-red at the hottest setting. If the element were to get even hotter it would glow
white or blue, but at these temperatures the stove element and pots would melt.

Figure 2.1 Electric stove element glowing red at low temperature and orange-red at a higher
temperature.
Humans at 37 °C or 310 K emit about 60 watts of electromagnetic radiation, mainly in the infrared
band.
An ideal blackbody absorbs all the EM radiation that falls on its surface. That is why it appears black
to us because our eyes perceive the absence of light to be black and see a body as black when it
reflects little light.
Kirchhoff found experimentally that at a given temperature all blackbodies had a well defined radi-
ance curve. That is to say that for a particular temperature, say 3000 K a black body emits a definite
intensity at a given wavelength.
Q.nb ��� 3

Figure 2.2 blackbody radiance — energy per unit wavelength interval versus wavelength in μm
Radiance vs λ
radiance

5780 K Sun

4000 K

3000 K

1000 K

λ(μm)
1 2 3

Figure 2.3 Mathematica plot of same graph with correct colours. Also temperature of the Sun used
in this graph is closer to actual value.

���������������
To do this experiment we use any material to make a cavity with a small hole in it. The hole approxi-
mates a blackbody because light that enters it reflected many times internally and is completely
absorbed or mostly absorbed before it has a chance to reflect back out of the small hole.
4 ��� Q.nb

Figure 2.4 A cavity with a small hole. EM radiation incident on the hole is usually complete
absorbed so the hole approximates well a perfect blackbody.

���������������������
In 1896 Wilhelm Wien found experimentally that the wavelength of the peak of the blackbody radi-
ance curve is inversely proportional to temperature in Kelvins.

λpeak T = b Wien displacemnt law


(1)
b = 2.898 × 10-3 m·K

In 1911, Wien was awarded the Nobel Prize in Physics “for his discoveries regarding the laws
governing the radiation of heat.”

�����������������
If we have a blackbody with a temperature of 3000 K, what is the wavelength of the peak of the
radiance distribution?
Solution
b
λpeak = T


λ���� = /� ����[{� → ������}� �������]

���� × ��-�

λpeak = 1 μm or 1000 nm
The wavelength of the peak of radiance curve is about 1 μm or 1,000 nm. This wavelength is longer
than the wavelength of red light and so it is outside the range of our vision. We call the part of the
EM spectrum the near infrared.

������������������λmedian���������������������������������������������������
Q.nb ��� 5

The peak wavelength of the blackbody radiance curve while informative does not tell us a great deal
about the rest of the power in the distribution. A more interesting value is the median wavelength.
That is for a given temperature the wavelength that divides the power spectrum in half so that half
the power is emitted in EMR with a wavelength longer than λmedian and half the power is emitted in
EMR with a wavelength shorter than λmedian .
The experimental law that relates λmedian to temperature in Kelvins is

λmedian T = 4.11 × 10-3 m·K (2)

��������������������
The total power radiated over the entire spectrum by a blackbody is given by the Stefan-Boltzmann
law

P = ϵ σ A T4 Stefan - Boltzmann law (3)

where:
ϵ — emissivity (0 → perfect reflector, 1 → black body)
σ - Stefan - Boltzmann constant (W · m-2 · K-4 )
A - surface area of the body (m2 )
T - temperature (K)

�����������������
What is the total power output of the Sun?
T = 5780 K
R⊙ = 6.955 × 108 m
Solution
P = σ A T4
= σ 4 π R2⊙ T 4

� π �� σ �� /� ����� → ����� � → ����� × ��� � ��������


������� × ����

P = 3.85 × 1026 Watts


About 1026 Joules, an enourmous quantity of energy is output by the Sun every second!

����������������������������
In the late 1800’s Rayleigh and Jeans did an analysis of a cavity blackbody experiment using classi-
cal physics. They treated electrons in the walls of the cavity as harmonic oscillators and used the
equipartition of energy theorem to find the power in the modes of vibration of the electrons. The
classically derived a formula, known as the Rayleigh-Jeans formula, disagreed with experiment. The
energy became unbounded at short wavelengths.
6 ��� Q.nb

Radiance vs λ
radiance

5000 K — experiment

5000 K — Rayleigh-Jeans

λ(μm)
1 2 3 4

Figure 2.5 Plot of Rayleigh-Jeans formula versus the experimental values at 5000 K. The marked
disagreement at short wavelengths was the ultraviolet catastrophe of Classical Physics.

Figure 2.6 Another view showing the visible, infrared, and UV EM regions.
Because at the time the shortest wavelength known was ultraviolet radiation, this problem with
Classical Physics was called the ultraviolet catastrophe.

����������������������������
Max Planck (1858 — 1947) resolved the ultraviolet catastrophe by proposing that vibration states of
electrons could only be excited in discrete chunks of energy, and that this energy was proportional
to their frequency. To excite a high frequency electron vibration requires a larger minimum energy
than that required to excite a slower vibrating electron. So at lower temperatures there is not enough
energy to excite the high frequency states and the electrons do not vibrate at these higher frequen-
Q.nb ��� 7

cies. Planck’s theoretical curve, fell away at shorter wavelengths and agreed perfectly with
experiments
2 πhc 2
R (λ, T ) = hc
— Planck blackbody radiation law (4)
λ 5 e λkT - 1

Planck introduced his all important constant, and found that to match experiment results the value of
the constant is:

h = 6.63 × 10-34 J · s Planck' s constant (5)

The new constant quantifies the fundamental “graininess” of physical reality at small scales.
Planck’s work implies that the energy of the vibrating molecules that emit the blackbody radiation
must come in multiples of a fundamental unit h f.

E=nhf n = 0,1,2,...

where:
f - the oscillation frequency of the electron
n — a counting number related to the energy of the electron
h — Planck’s constant

�����������������
An electron is oscillating with frequency f = 4.3 × 1014 Hz. What is the minimum non-zero energy of
the electron in electronvolts.
Solution
1 eV
E = nhf · e J

� ��
/� ����� → ���� � → ��� × ���� � ��������

�������

E = 1.8 eV
This is a typical value for an electron in a cavity that is emitting red visible light.

�������������������������������������
Below is a new type of diagram called an energy level diagram. It is a simple but most useful way of
representing allowed energy states and the gaps between them. In this case we see that states of
an electron vibrating at a single frequency f. We will make use of energy level diagrams throughout
this topic.
8 ��� Q.nb

�����������������������������������
By taking the derivative of Planck's blackbody equation and setting the derivative to zero we find
that the radiation peaks at a wavelength that is inversely proportional to the temperature This result
was found experimentally by Wien several years earlier and expressed as Wien’s displacement law.
λpeak T = 2.898 × 10-3 m · K (6)
Planck’s blackbody equation also agrees with the experimental median wavelength of the blackbody
radiance distribution — the wavelength that divides the power distribution in half, with half the power
emitted by wavelengths shorter than this and half the power emitted by wavelengths longer than it.
λmedian T = 4.11 × 10-3 m · K (7)

Lecture 2

�����������
Wolfson 34.3

������������������������
In 1887 Heinrich Hertz observed that metals eject electrons when struck by light or other EM
radiation.
Q.nb ��� 9

Figure 2.7 Classical model of photoelectric effect.


Doing an analysis using Classical Physics we would conclude the following:
◼ the electron oscillates in the incoming EM waves.
◼ the electron absorbs more energy and the amplitude of the oscillations grow.
◼ eventually the electron has enough energy to escape the surface of the metal.

Figure 2.8 Photoelectric apparatus and graph showing potential between metal surface and
electrode.
Above is the apparatus used to study the photoelectric effect. Taylors College owns several of these
apparatuses and you can do the experiment in the Physics B labs. As you can see light strikes a
metal surface inside a vacuum tube. The vacuum is required so that the electron will not be stopped
by air molecules. An electron leaves the metal surface with a certain amount of kinetic energy. An
10 ��� Q.nb

electrode is held at a negative potential relative to the metal so that the electrode repels the elec-
tron. That’s why this voltage is called the stopping potential. Only if the electron has enough kinetic
energy to climb the potential barrier will it hit the electrode and be recorded as current in the
ammeter.

���������������������������������������������������������
Experimental result Incorrect! Classical Prediction
Current begins immediately, even in extremely Dim light → small amplitude waves.
dim light Should take a long time for electron
to get enough amplitude to escape.
Maximum kinetic energy of electron Brighter light → large amplitude.
is independent of light intensity. Should give large electron kinetic
energy.
Below cut - off frequency, no electrons at all If the light (EMR) has sufficient intensity
are emitted. then electrons are emitted. There is no
cut - off frequency.
Above the cut - off frequency, the Cut - off frequency is a mystery in
maximum K.E.of the electrons increase Classical Physics. So no prediction
linearly with the frequency. about this

Table 2.9 Failure of classical physics to explain the photoelectric effect.


In 1905 Albert Einstein during his Annus mirabilis (extraordinary or miraculous year) suggested that
light comes in quanta or packets of energy each with energy given by the formula

hc
E= hf = energy of a photon (8)
λ

where:
h — Planck’s constant
f — frequency of the photon
λ — wavelength of the photon (c = f λ)
To quote Einstein
In fact, it seems to me that the observations on “black-body radiation”, photoluminescence, the
production of cathode rays by ultraviolet light and other phenomena involving the emission or
conversion of light can be better understood on the assumption that the energy of light is
distributed discontinuously in space. According to the assumption considered here, when a light
ray starting from a point is propagated, the energy is not continuously distributed over an ever
increasing volume, but it consists of a finite number of energy quanta, localised in space, which
move without being divided and which can be absorbed or emitted only as a whole.
Albert Einstein, 1905
These light quanta as Einstein called them later became known as photons. Einstein receive the
Nobel Prize in 1921 “for his services to Theoretical Physics, and especially for his discovery of the
law of the photoelectric effect”.
So we can now give the correct quantum definition of the photoelectric effect.

The photoelectric effect is the ejection of an electron


(9)
from the surface of a metal when the electron is struck by a photon.
Q.nb ��� 11

At a given fixed frequency, brighter light or more intense light means more photons, but each of
these photons still has the same energy, h·f.
Each material has a minimum energy that is required to eject an electron from its surface — this is
call the work function ϕ of the material.

For example:

Table 1 Work functions of some elements (based on Wolfson Table 34.1)


So when a photon with energy Ephoton strikes a metal, part of the photon’s energy is used to over-
come the work function ϕ, the remaining photon energy goes in to kinetic energy of the electron:
Kmax = Ephoton - ϕ
We say Kmax because sometimes an electron that escapes is not directed at the electrode, or
perhaps the electron is not on the surface of the metal and it gives up more energy to the metal than
the surface work function.
Using equation (8) above we get

Kmax = h f - ϕ (10)

where:
Kmax — the maximum kinetic energy of an ejected electron
h — Planck’s constant
f — frequency of the photon
ϕ — the work function of the metal.

�����������������
Find Kmax for an electron ejected from Cs using blue light λblue = 450 nm
Solution
c
c=fλ → f = λ
—1
hc
Ephoton = h f = λblue
—2 using 1
hc 1
Kmax = λ e
- ϕCs
12 ��� Q.nb

�� ���
���� = - ϕ /� ϕ → ����� λ →  /� �����
�λ ���
������

Kmax = 0.62 eV

�����������������
What is the stopping potential of the electron in Example problem 1 above.
Solution
The electron has energy 0.62 eV. This by definition is the energy required to move an electron
through a potential difference of 0.62 volts. So an electron with 0.62 eV of kinetic energy will have
enough energy to move up a potential hill of 0.62 volts.

Stopping potential = 0.62 V

���������������������
So electromagnetic (EM) waves are quantised particles or photons, and yet they interfere as waves
in, say, Young’s double-slit experiment.

This is called wave-particle duality and is the central mystery of quantum mechanics.

It has been joked that, “light behave like a particle on Mondays, Wednesdays, Fridays, and Satur-
days, and like a wave on Tuesday, Thursdays, and Sundays”.

������������������
Sometimes when a photon scatters off an electron in a crystal, none of the photon’s energy is
transferred to the crystal. We then get Compton scattering and the photon scatters elastically. This
scattering shows perhaps better than any other experiment the particle properties of a photon.

The Compton effect is the elastic collision of a photon with a stationary electron. (11)

When EM radiation falls on an electron, Classically the electron should move in the electromagnetic
field and re-radiate the absorbed energy.
Analysing this classically it should radiate circular waves at the same frequency as the incoming
waves
Q.nb ��� 13

Figure 2.10 Classical analysis shows electron re radiating incoming EM wave as a circular wave of
the same frequency.
But experiment shows the following:

Figure 2.11 Experiment shows photo scatters elastically from stationary electron and wavelength of
the photon becomes longer after scattering to conserve energy and momentum.

◼ the scattered photon has a lower frequency and longer wavelength than the incident photon.
◼ the photon and electron scatter elastically like billiard balls - energy and moment are conserved.
The change in wavelength of the incident photon is given y the formula

Δλ = λ - λ0 = λC (1 - cos θ) Compton shift (12)

where λC is the Compton wavelength of an electron

h
λC = = 2.43 x 10-12 m = 2.43 pm electron Compton wavelength (13)
me c

For a photon we have already seen that the energy is given by


hc
E = hf =
λ
We will now give you an equation for the momentum of the photon.
14 ��� Q.nb

h
p= momentum of a photon (14)
λ

By conservation of energy, any energy that the photon loses is given to the electron. Also initially
the electron is not moving and has zero momentum, so the photon has all the initial momentum. By
conservation of momentum the momentum of the photon after the collision plus the momentum of
the electron after the collision must equal the initial momentum of the photon.

�����������������
After a Compton scattering event the electron has the maximum kinetic energy when the photon
scatters through what angle?
Solution
The electron will have maximum KE when it has maximum momentum. Seeing momentum is
conserved the electron will have greatest momentum when the change in momentum of the photon
is greatest — that is when the photon is scattered at 180°.

����������������������
The photon is a particle of electromagnetic radiation or light when visible. Like all particles including
electrons, protons, and neutrons, the photon has a wavelength.

h
λ= wavelength of a photon (15)
p

Wave particle duality means that at sometimes the photon’s particle properties are more evident as
in Compton scattering, and at other times its wave properties are more evident such as in Young’s
double slit experiment.
Photons in free space travel at the speed of light (because light is just photons). So by the wave
equation

c = fλ (16)

The energy of a photon is given by

hc
E= hf = energy of a photon (17)
λ

Where the second form of the energy follows from the first by using the relationship c = f λ.

h hc
Finally looking at λ = p
and E = λ
we see that for a photon

E= pc energy and momentum of a photon (18)

����������������
Q.nb ��� 15

An incoming photon has a wavelength of λ0 = 50.00 pm. It scatters elastically off a stationary elec-
tron at an angle of θ = π; that is it travels back in the direction it came from.
Find the following:
A) p0 — the initial momentum of the photon.
B) λ — the wavelength of the scattered photon.
C) p — the momentum of the scattered photon.
D) pe — the momentum of the scattered electron.

Figure 2.12 a 50.00 pm photon scatters off a stationary electron at an angle of θ = π.


Solution
◼ A)
Initial momentum of photon
h
p0 = λ0

� �����
/� λ� →  /� �����
λ� ����
����� × ��-��

p0 = 1.326 × 10-23 kg· m·s-1


◼ B)
Wavelength of scattered photon
λ = λ0 + Δλ
Δλ = λC (1 - cosθ) = λC [1 - (-1)] = 2 λC
so
λ = λ0 + Δλ
= 50.00 + 2·2.43 pm
�����
�� + �� × ���� /� λ� →  /� �����
����
�����

λ = 54.86 pm
◼ C)
Momentum of scattered photon
h
p= λ
16 ��� Q.nb

� �����
/� λ →  /� �����
λ ����
������� × ��-��

p = -1.209 × 10-23 kg· m·s-1 (in the negative x-direction)


◼ D)
Momentum of scattered electron
The incoming photon has all its momentum in the x-direction, that is zero momentum in the y-
direction. So by conservation of momentum the sum of the momentum of the scattered photon and
scattered electron must have zero momentum in the y-direction. We can treat this as a 1D problem.

By conservation of momentum we get the initial momentum is equal to the final momentum.
p0 = p e + p
We can add the x-components of momentum as scalars.
pe = p 0 - p
= [1.326 - (-1.209)] ×10-23
����� + �����
�����

pe = 2.535 × 10-23 kg· m·s-1

�������������������������������
The experiments that led to the development of quantum mechanics related to electromagnetic
radiation emitted by objects and gasses under certain special conditions.

�������������
Only a very small band of the electromagnetic spectrum is visible to the human eye. We call this
section of the EM spectrum, light.
Visible light:
colour wavelength frequency photon energy
red 700 nm 4.3 × 1014 Hz 1.8 eV
violet 390 nm 7.7 × 1014 Hz 3.2 eV
Table 2.13 The two limiting colours of visible light, red and violet, showing the wavelength, fre-
quency, and photon energy of each.
Note that colour is completely a property of the human eye and has nothing to do with any property
of the band of the electromagnetic spectrum that we call light.

�����������
Invisible wavelengths slightly shorter than violet light are called ultraviolet, and invisible wavelengths
slightly longer than red light are called infrared.

The diagram below shows a portion of the EM spectrum and the energy of a photon of a given
wavelength and frequency. The longer wavelengths are on the left and the shorter wavelengths are
on the right.
Q.nb ��� 17

Figure 2.14 The electromagnetic spectrum showing visible light, names of bands of the spectrum,
and the energy of a photon of a given wavelength and frequency.

Lecture 3

����������������������������������������������������
����
Wolfson 34.4

���������������������������������
In 1804 William Wollasten used a glass prism to diffract sunlight. Shorter wavelengths have a
slightly higher index of refraction in glass and are deflected more by the glass. Wollasten noticed
dark lines in the Sun’s spectrum. This type of spectrum, with dark lines on a continuous background,
is known as an absorption spectrum.

Ten year later Joseph Fraunhofer a German master glassmaker made an extremely high quality
prism. Fraunhofer made the first careful measurements of the amount of deflection of each dark
lines and the Sun’s absorption spectrum, and these dark lines now bear his name.

Included among these lines is the hydrogen absorption spectrum.

Figure 2.15 Dark Fraunhofer lines in the absorption spectrum of the Sun.
18 ��� Q.nb

Figure 2.16 Blackbody spectrum of the Sun showing Fraunhofer absorption lines.

�����������������
An emission spectrum is bright lines on a black background. The emission spectrum of many ele-
ments can be studied in the laboratory. A low pressure gas sample of the element is placed in a
vacuum tube and exited using electric discharge. The hydrogen spectrum is shown below.

Figure 2.17 The four visible hydrogen emission spectrum lines in the Balmer series. (Wikipedia)
In 1884 Johann Balmer realised the for hydrogen the visible lines of the emission spectrum obeyed
the relationship

1
λ = RH  212 - n12 

where:
RH = 1.097 × 107 m-1 — the Rydberg constant for hydrogen
n = 3, 4, 5, 6,...
Two of Balmer’s colleagues quickly confirmed the existence of other non-visible lines of the series in
the spectrum of hydrogen in white stars.
Q.nb ��� 19

Figure 2.18 Emission and absorption spectrum for Hydrogen. (Line on far left is ultraviolet and not
visible to naked eye.)
Soon after, other series of lines in the hydrogen spectrum were discovered and the Swedish physi-
cist Johannes Rydberg in 1888 devised the Rydberg formula.

1 1 1
= RH - — Rydberg formula (19)
λ nf 2 ni 2

where:
RH = 1.097 × 107 m-1 — the Rydberg constant for hydrogen
n f — final state
n f = 1 — Lyman series (ultraviolet)
n f = 2 — Balmer series (first 4 lines are visible)
n f = 3 — Paschen series (infrared)
ni — initial state
ni = n f + 1, n f + 2, n f + 3,...

��������������������������������
In 1911 the New Zealand physicist Ernest Rutherford showed that atoms consist of a small, dense,
positively charge nucleus surrounded by electrons.
20 ��� Q.nb

Figure 2.19 Helium atom model. A nucleus with two protons and two neutrons. There are two
electrons circling.

�����������������������������������������
Classically an orbiting electron is accelerating so it should emit EM radiation. It would lose energy to
the radiation and spiral in to the nucleus. All matter would very quickly disintegrate as the electron
fall into the nucleus and attach themselves to the protons. For solids the electrons form the bonds
that give the matter its crystalline or metallic structure. The electrons hold the nuclei apart.

We observe, however, that matter is stable over very long periods of time so a Classical analysis of
the atom does not agree with observed behaviour of matter and must be wrong.

Figure 2.20 Classically, an orbiting electron is accelerating and would emit EM radiation causing it
to give up energy and spiral in to the nucleus.

���������������������
In 1913 Danish physicist Niels Bohr proposed an atomic theory that accounted for the spectral lines
of hydrogen given by the Rydberg formula. The model also gave a value for the Rydberg constant in
terms of fundamental physical constants such as Planck’s constant h, the speed of light c, and the
elementary charge e. Despite its notable success for the hydrogen atom it is only a first approxima-
tion to the much more accurate Schrödinger equation which we will look at in the next section,
Bohr conjectured that the angular momentum of the electron in a hydrogen atom is quantised in
units of Planck’s constant divided by 2π. Remembering that
  
L = r ×p
Q.nb ��� 21

L = r me v
So according to Bohr’s model angular momentum is quantised as follows:

L = me vr = n ℏ n = 1, 2, 3, …

where:
me vr - is the angular momentum of the electron (  )
r ×p
h
ℏ= 2π
(pronounced h-bar) Planck’s constant divided by 2π
also called the reduced Planck constant
n = 1, 2, 3, ...
This implies the electron can only have discrete energy values given by

Ry
En = - electron energy (20)
n2

where:
Ry = 13.606 eV — the Rydberg energy
n = 1, 2, 3,...
Because the electron in the Bohr model is orbiting the nucleus like a planet orbiting the Sun, if the
angular momentum is quantised, the electron can only be in discrete orbitals. The electron jumps
from one orbital to another by emitting or absorbing a photon with energy equal to the energy
difference between the two orbitals.

Figure 2.21 The Bohr model of the atom showing a electron jumping to a lower energy level and
emitting a photon with energy equal to the difference in energy between the two levels.
The n = 1 orbital is called the ground state.
The n = 2 orbital is called the first excited state.
The Bohr model predicts that the radius of the nth orbital is proportional to n2 .

rn = n2 a0

where
a0 = 0.0529 nm is called the Bohr radius
n — the orbital number n = 1, 2, 3,...
Note that En is negative and approaches zero as n → ∞. That is because the positive nucleus
makes a Coulomb potential well for the electron. We call an hydrogen atom ionized when its elec-
tron has been completely separated from the nucleus by a large distance — that is when n = ∞. If
we use the convention that the ionized atom (n = ∞) has zero energy, then as the electron enters
22 ��� Q.nb

the potential well of the nucleus it will have less energy or a negative value because it would take
positive work on the system to pull the electron back out of the potential well.

Figure 2.22 Energy of the electron in the Bohr model. The curve shows the Coulomb potential, but
the energy level is the sum of the potential and kinetic energy of the orbiting electron.
We can now get a value for the energy of the photon that is emitted when an electron falls from a
higher initial energy level ni (say ni = 3) to a lower final energy level n f (say n f = 2). It is just the
difference in the energies of the two states as given by equation 20.
Ry Ry
E ni → n = - - -
f nf 2 ni 2

1 1
Eni →n f = Ry - — photon energy with electron jump (21)
nf 2 ni 2

where
Eni → n f — the energy of the photon emitted after the electron does the quantum jump
ni — the number of the higher energy orbital
nf — the number of the lower energy orbital
Ry = 13.6 eV — the Rydberg energy

��������������������������������������
It is a bit easier to visualise the energy of the photons that result from the Bohr atom using an
energy level diagram. Below is the diagram showing energies for the n = 1 to n = 4 orbitals of the
hydrogen atom and their energy in electronvolts. The energy of higher numbered orbitals get closer
and closer together and are hard to show on an energy level diagram. Also shown is the wavelength
of the photon emitted for the first three transitions from a higher state to the ground state.
Q.nb ��� 23

Figure 2.23 Energy level diagram for the Bohr atom showing the first three emission lines and
electron jumps for transitions from the first three excited states back to the ground state.
And here is an more detailed energy level diagram showing the first few emission lines of the Lyman
series n f = 1, the Balmer series n f = 2, and the Paschen series n f = 3.
24 ��� Q.nb

Figure 2.24 An energy level diagram for the Bohr atom showing three series of emission lines and
the corresponding electron jumps.

�����������������
A) What is the energy in electronvolts (eV) of the emitted photon when the electron in a
hydrogen atom falls from the n = 4 state to the n = 2 state?
B) Find the wavelength of the photon from part A.
C) In what part of the electromagnetic spectrum does this photon belong?

Solution
◼ A) What is the energy in electronvolts (eV) of the emitted photon when the electron in a hydrogen
atom falls from the n = 4 state to the n = 2 state?
The higher initial state n = 4 is the ni state has energy
R 13.606
En=4 = - 42y = - 16
= -0.85 eV
and the final state n = 2 has energy
R 13.606
En=2 = - 22y = - 4
= -3.4 eV
The change in the electrons energy is therefore
Δ E4→ 2 = -2.55 eV
So the photon carries away an energy of
Eγ = 2.55 eV
������ = ����[{�� → �� � � → �}� �������]�
��
�� = /� ������
�� �
��
�� = /� ������
���
��→� = �� - ��
��������

������

-�������

◼ B) Find the wavelength of the photon from part A.


hc
E = hf = λ
hc h c 1 eV
λ= E
= E e J

�� �
- /� ������
��→� �
������� × ��-�

λ = 480 nm
OR
An alternative way to find the wavelength is as follows:
The initial state is ni = 4 and the final state is nf = 2
1 1 1
λ
= RH n 2 - n 2
f i
Q.nb ��� 25

1 1
= R H  22 - 42

1 1
= RH  4 - 16

3
= RH 16
= 2.057 × 106
1 16
λ= RH 3
1 16
= = 4.86 × 10-7
1.097 × 107 3
= 486 nm
◼ C) In what part of the electromagnetic spectrum does this photon belong?
The photon belongs to the visible part of the EM spectrum that we call light. It is toward the violet
end and is in-fact the colour cyan.

Lecture 4

����������������
Wolfson 34.5

���������������������
Electromagnetic radiation (EMR) behaves as a particle in some experiments (photoelectric effect,
Compton effect), and as a wave in other experiments (Young’s double slit). In 1924 Louis de Broglie
(1892 – 1987) suggested the same should be true of all particles including protons, neutrons, and
electrons. The wavelength of any particle including a photon is

h
λ = any particle (22)
p

Wave phenomena can only be detected for small momenta, when the matter wave wavelength is
long enough to be detectable. Because momentum p = m v, we can get a small momentum from an
object with small mass moving slowly.

De Broglie's hypothesis was confirmed in 1927 when two American physicists, Davisson and
Germer, observed an interference pattern for slow electrons fired at a crystal lattice — the rows of
the crystal acting as very small slits for the electrons to pass through. Two years later, in 1929, De
Broglie received the Nobel Prize in Physics for his matter waves postulate.

Figure 2.25 Single-slit interference pattern made by electrons as observed by Davisson and Ger-
26 ��� Q.nb

mer in 1927.

�����������������
An electron an a proton are moving at the same speed. Which has the longer wavelength?
Solution
The wavelength of the electron λe divided by the wavelength of the proton λp is
λe h pp pp
λp
= pe h
= pe
by the De Broglie hypothesis
Now substituting p = m v
λe mp v mp
λp
= me v
= me
≃ 1836
The electron has a smaller mass so when traveling at the same speed as a proton it has less
momentum p = mv. The wavelength λ goes as the inverse of momentum so the electron has the
longer wavelength.

������������������������������
Matter waves explain the quantisation of angular momentum in the Bohr atom. This is not surprising
as it was this very phenomenon that led De Broglie to the concept of matter waves. The electron is
circulating in the attractive Coulomb force of the nucleus, so like a planet orbiting the Sun, the
electrons must have the given velocity to maintain a circular orbit against of a particular radius. This
velocity is found by setting the required inward radial acceleration, equal to the Coulomb attractive
force.
m v2 e2
r
=k r2
Solving for v we get,
k e2
v=
me r
So for a given radius the electron has a particular velocity, and therefore a particular momentum.
h h
This momentum means that for a given radius the electron has a given wavelength λ = p
= me v
. If
this given wavelength forms a standing wave when wrapped around the orbit path, the wave inter-
feres constructively with itself and we have an allowed orbit. Bohr showed that an allowed orbit has
radius r = n2 a0 .

Figure 2.26 An electron with just the right momentum to maintain a circular orbit and to have stand-
Q.nb ��� 27

ing waves for its wavelength.


Increase r by a small amount from an allowed value will lead to a slightly smaller velocity (longer
wavlength) and a larger orbit circumference. This results in a wavelength that does not form a
standing wave when wrapped around the circular path of the orbit, so the wave will interfere destruc-
tively with itself and the orbit is disallowed.

Figure 2.27 if the radius is increase slightly from an allowed orbital radius, we then get destructive
interference and a disallowed orbital.

�����������������
An electron has energy E = 1.0 × 10-30 J.
A) Find the electron’s wavelength.
B) Find its velocity.
Solution
◼ A)
1 (me v)2 p2
E = 2
me v2 = 2 me
= 2m
so

p= 2 me E — (1)
h h
Now λ = p=
2 me E


������ = ���������� �� → �
����

λ= /� ������
� �� ��
�����������

λ = 4.9118 × 10-4 m
So the wavelength to 2 s.f. is about half a millimetre
λ = 0.49 mm
◼ B)
28 ��� Q.nb

p 2 me E 2E
v= me
= me
= me

� ��
�= /� ������
��
�������

v = 1.5 m· s-1 (2 s.f.)

�����������������������������
Wolfson 34.6
One of the founders of quantum mechanics, German physicist Werner Heisenberg, buiding on de
Broglie’s hypothesis, discovered in 1927 the uncertainty principle. The uncertainty principle states
that due to a particle’s wave nature it is not possible to know the particle’s position and momentum
at the same time.

Δp Δx ≥ ℏ — Heisenburg uncertainty principle (23)

where:
Δp — the uncertainty in the particle’s momentum
Δx — the uncertainty in the particle’s position
h
ℏ= 2π
— hbar, Planck’s constant divided by 2π.

The more accurately we know the position of the particle, the less accurately we know the
momentum.

������������������
Heisenberg did many thought experiments that have since been carried out in reality. Let’s do our
own thought experiment. Picture electrons directed horizontally through a large gap. These elec-
trons have large vertical position uncertainty Δx. By the uncertainty principle, there will be a small
vertical momentum uncertainty Δp, or there will be little deviation in the vertical direction.

Figure 2.28 A horizontal beam of electrons directed through a large vertical gap have a large
vertical Δx, and so a small vertical Δp.
Q.nb ��� 29

This is what we would expect from light going through a large slit — little diffraction and a sharp
image on the screen.

Figure 2.29 A horizontal beam of electrons directed through a small vertical gap have a small
vertical Δx, or a well defined horizontal position, and so vertical Δp is large.
Now imaging that we make the slit narrow. The same horizontal beam of electrons when it passes
through the slit now has a well defined vertical position Δx, this means that by the uncertainty
principle it will have a large vertical momentum uncertainty, Δp. That is to say, the beam will diverge
vertically in the same way that light diffracts when it passes through a narrow slit. This is no accident
as event thoug both electrons and photons are particles, they both also have a wave nature and so
experience diffraction at a single slit.

�����������������
A beam of aluminium (Al) atoms is used to dope a semiconductor. The atoms vertical velocity is
known to within 0.20 m· s-1 (that is to say Δ v = 0.2 m·s-1 ). What is the minimum uncertainty in
vertical position?

The mass of an Aluminium atom is approximately 26.98 u where u is the unified atomic mass unit.
(1 u ≃ 1.66 × 10-27 kg)
Solution
Δv = 0.2 m· s-1
Δp = mAl Δv
Δx Δp ≥ ℏ
Δx ≥
ℏ = ℏ
Δp mAl Δv


Δ� = /� {��� → ������ Δ� → ���} /� �����
� π Δ� ��� �
������� × ��-�

Δx ≥ 1.18 × 10-8 = 12 nm (2 s.f.)

�����������������
Find the minimum energy for an electron confined in an atom of diameter one Ångstrom (1Å). Use a
one-dimensional 1D approximation to this 3D problem.
30 ��� Q.nb

Note that in atomic physics the unit Ångstrom (symbol Å) is commonly used. One Ångstrom is
10-10 metres. So this atom could also be said to have a diameter of 0.1 nm.
Solution
Using a 1D approximation,

Δx = 0.1 nm

Now if the electron has momentum of magnitude p, the motion can be directed to the left or to the
right within the potential well.

Figure 2.30 The electron’s momentum can be in either direction so Δp the change in momentum is
twice the magnitude.
So the uncertainty in the momentum is
Δp = p - (-p) = 2p. — (1)
and
Δp Δx ≥ ℏ the uncertainty principle
2 p Δx ≥ ℏ using (1)
p≥
ℏ = h — (2)
2 Δx 4 π Δx

The kinetic energy


p2 1 h 2
KE = 2 me
= 2 me
 4 π Δx  using (2)

� �
  �
� π Δ�
�� = /� Δ� →  /� �����
� (� ��) ����
��������

Energy of electron confined to atom = 0.955 = 1 eV (1s.f.)

�����������������������������������������
There is also a lower limit on how well we can know a particle’s energy given the length of time it
stays at that energy

ΔE Δt ≥ ℏ — Heisenburg uncertainty principle for Energy and time (24)

where:
ΔE — the uncertainty in the particle’s energy
Δt — the time for which the particle stays at that energy
Q.nb ��� 31

h
ℏ= 2π
— Planck constant divided by 2π

Lecture 5

��������������������������������
Wolfson 34.7
In 1920 Bohr formulated the correspondence principle.

The results of quantum mechanics approach those of classical mechanics in situations where the
size of an individual quantum of energy is negligible (h → 0).

In other words for large orbits and for large energies, quantum calculations agree with classical
calculations.
32 ��� Q.nb

�������������������
������������������

�������������������������������������
Wolfson 35.1
The relationship between waves and a particles is statistical.
◼ For electromagnetic radiation (EMR) the probability of finding a photon is proportional to the
intensity of the electromagnetic wave. Intensity goes with the square of the wave amplitude.

Figure 3.1 The probability of finding a photon is proportional to the intensity or square of the ampli-
tude of the electromagnetic wave.
This relationship is true for all particels.

The probability of finding the particle is proportional to the square of the amplitude, of the matter
wave or wave function. This is true electrons, protons, and neutrons, as well as for photons.

���������������������������������������
Wolfson 35.2
Schrödinger building on the work of de Broglie and using non-relativistic classical mechanics formu-
lated an equation for matter waves.

The one-dimensional time-independent version of the Schrödinger wave equation is

ℏ 2 ⅆ 2 ψ (x)
- + U(x) ψ (x) = E ψ (x) (25)
2m ⅆx 2

where:
ψ(x) — the wave function (Greek letter psi ψ)
m — mass of the particle (for this topic this is me the electron mass)
h
ℏ = 2 π — Planck constant divided by 2π
U(x) — particle’s potential energy
E — particle’s total energy (Ek + U)
Q.nb ��� 33

Schrödinger’s wave equation is based on other similar classical wave equations, and the classical
energy equation
EK + U = E
where
Ek — kinetic energy
U — potential energy
E — total energy
1 1 1
and since Ek = 2
m v2 = 2m
(m v)2 = 2m
p2
p2
2m
+U=E

�����������������������������
The wave function itself is not physically observable. It is the square of the wave function that tells
us the statistical likelihood of finding a particle.
To make this precise, we define the �����������������������������������������

P (x) = ψ2 (x) probability density (26)

So the probability of finding the particle in an interval��������������


prob �� � + Δ�) = � (�) Δ � = ψ2 (x) Δx

Figure 3.2 The probability density P(x) = ψ2 (x) when multiplied by Δx gives the probability of finding
the particle in the interval [x, x + Δx].

�������������������������������������
The wave function Ψ(x, t) is a complex function of position and time. The frequency of Ψ, gives us
the particles energy, and the wavelength of Ψ, gives us the particle’s momentum.

h
E = hf = = ℏω
T Energy and moment of particle from properties of ψ
h
p = = ℏk
λ
where:
E — energy of the particle
34 ��� Q.nb

p — momentum of the particle


f — frequency of ψ the wave function
ω — the angular frequency 2 π f of ψ the wave function
T — period of ψ
λ — wavelength of ψ

k — the wavenumber λ
h — Planck constant (6.63 × 10-34 J·s)
h
ℏ— 2π

�����������������������
So if P(x) is the probability density and we integrated it over all of space we have the probability of
finding the particle anywhere in space. If the particle exists its probability of being somewhere will be
certainty, or 1. That is
+∞
∫-∞ P(x) ⅆx = 1
But we as stated earlier P(x) = ψ2 (x) So

+∞
 ψ2 x ⅆx = 1 Normalisation condition (27)
-∞

������������������������������������������
Wolfson 35.3

����������������������������������������1D������������
To find solutions to the Schrödinger wave equation let’s start with an easy problem; a one-dimen-
sional, 1D, infinite square well.

The potential energy of the electron is infinite except on the interval [0, L], where the potential
energy is zero. Here is what the potential energy function of the electron looks like.
∞ x<0
U(x) =  0 0<x<L potential energy
∞ L<x
and here is a graph of the potential energy function.
Q.nb ��� 35

Figure 3.3 An infinite square well where the potential goes to infinity for x < 0 and for x > L, but is
zero in the interval x ϵ [0, L].
You can picture this as a one-dimensional well with rigid walls, so the electron is confined to stay in
the well.

We impose a boundary condition that ψ is zero in the “walls” of the well. That is
ψ(x) = 0 for x ∉ [0, L] (x is not an element of the interval [0, L])
Since within the well x ϵ [0, L], the potential energy function is zero, we can substitute U(x) = 0 into
the Schrödinger equation (25), and we get
ℏ 2 ⅆ2 ψ (x)
- = E ψ (x) (28)
2m ⅆ x2
However another condition on ψ is that is must be continuous. So a second boundary condition is
that the wave function is zero at the walls of the well, or
ψ (0) = 0, and ψ (L) = 0.
Solutions of this differential equation are of the form
nπx
ψn (x) = A sin   where n = 1, 2, 3, ... (29)
L
36 ��� Q.nb

Figure 3.4 Solutions of SWE for 1D infinite square well.


Solutions are similar to standing waves on a string that is fixed at both ends — Melde’s experiment.
If we substitute Ψn (x) back into the form of the Schrödinger wave equation with U = 0, equation (28),
we find that Ψn (x) is a solution provided that
n2 π 2 ℏ 2 n2 h2
En = =
2mL 8 m L2

n2 h2
En = 8 m L2
energy levels for a 1D infinite square well
where: n = 1, 2, 3 ...

Notice that En for any value of n has a simple relationship to E1 .

En = n2 E1 energy levels for an infinite square well in terms of E1

Lecture 6

��������������������
The following is an energy level diagram. It shows the states of the infinite square well with increas-
ing value of n on the left, and the energy on the right. The height of the line gives its energy.
Q.nb ��� 37

Figure 3.5 An infinite square well energy level diagram showing energy level E1 to E5 and their
associated energy on the right-hand side.
Finally, as is customary on an energy level diagram, we plot the functions Ψn (x) using the energy
level as the x-axis. Note that this is not a conventional graph because each ψn (x) has its own
individual x-axis that is shifted up the energy scale.
38 ��� Q.nb

Figure 3.6 energy level diagram show wave functions associated with each energy En .

�������������������������
The solutions given above in equation 29 are not strictly wave function because we have not yet
normalised them. Remember that a wave function in order that the probability of finding the particle
somewhere is 1, should satisfy the normalisation condition given in equation 27, that is
+∞ L
 ψ 2 ( x) ⅆ x =  ψ 2 ( x) ⅆ x = 1 — since Ψ is zero outside of [0, L]
-∞ 0

We still have the variable A at our disposal to normalise the solutions. However A is constant for a
given solution so we can take it outside of the integral
L nπx
A 2  sin2 ⅆx = 1
0 L
1
But the area under the curve sin 2 x is just 2
over any number of half periods. So
L nπx L
A 2  sin2 ⅆx = A2 = 1
0 L 2
Or re-arranging
2
A= L

So normalised solution to the infinite square well are as follows:

2 nπx
ψn (x) = sin where n = 1, 2, 3, ... (30)
L L

��������������������������������������������
Remember that the wave function itself is not physically observable, so we are more interested in
the probability density function, which is the square of the wave function ψn 2 x. Below for instance
is normalised wave function Ψ1 (x) associated with the ground state E1 and its probability density
function ψ1 2 x. Note the different shape of the functions.

1.4

1.2

1.0

0.8

0.6

0.4

0.2

0.2 0.4 0.6 0.8 1.0

Figure 3.7 normalised solution ψ1 (x) for an infinite square well.


Q.nb ��� 39

2.0

1.5

1.0

0.5

0.2 0.4 0.6 0.8 1.0

Figure 3.8 probability density ψ1 2 x for an infinite square well.


and here is an energy level diagram showing probability density function instead of wave functions.

Figure 3.9 Infinite square well energy level diagram showing probability densities.

�����������������������������������������������������
Have a look at the ground state n = 1 in the energy level diagram showing probability densities in
Figure 3.9. Classically the particle is just as likely to be anywhere in the well (as shown by the
dotted line), quantum mechanics, however, says the particle is much more likely to be near the
centre of the well.

As the energy level increases to n = 15, quantum mechanics starts to move toward the classical
solution with the particle just as likely to be at any place in the well. As the energy level continues to
increase with n = 1000 say, the wavelength would be so small that it would be hard or impossible to
40 ��� Q.nb

detect any place in the well where the particle was excluded. The quantum mechanical solution
then, as required by the correspondence principle, would approach the classical solution.

������������������������������
For a simple harmonic oscillator the restoring force is proportional to displacement F = - k x. This
implies a quadratic potential function. It can be shown that for this shape potential function solutions
to the Schrödinger equation require that

1
En = n + ℏω where n = 0, 1, 2, ... simple harmonic oscillator (31)
2

1
Note that the ground state has n = 0 but is still a non-zero energy 2
ℏ ω.

Figure 3.10 energy levels for a harmonic oscillator


Below is a plot of the probability density function for some small values of n and for the quantum
state n = 10. Note a totally new quantum physics phenomenon called quantum tunneling — the
particle has a non-zero probability of being found outside the potential well.
Q.nb ��� 41

Figure 3.11 Probability density functions for a harmonic oscillator.

���������������������������������������������������������������
A classical simple harmonic oscillator, say a ball in a bowl, or a small amplitude pendulum, spends
more time at the edge of its swing. Classically then it is more likely to find the particle at the edges
of the potential well. The classical probability density is therefore larger at the edges of the simple
harmonic potential well. This is shown by the dotted lines in Figure 3.11.

We again see the correspondence principle holding in that as n becomes larger, the quantum
mechanical solution approaches the classical one, and the probability of finding the particle near the
edge of the well also becomes larger for the quantum mechanical solution.

�������������������������������
Wolfson 35
42 ��� Q.nb

Figure 3.12 This scanning-tunneling microscope image shows a “quantum corral” of 48 iron atoms
on a copper surface. (Wolfson)

You might also like