You are on page 1of 59

F LIMPOPO

UN YO
I VERSIT

Faculty of Science and Agriculture

SCHOOL OF PHYSICAL AND MINERAL SCIENCES

Department of Physics

MODULE OUTLINE
(SPHA031)
Quantum Mechanics

2023
MODULE OUTLINE

Module Title Quantum Mechanics


SPHA031 16
Module Code No. of Credits
Physics Physical & Mineral
Department School Sciences
Pre-requisites SPHA021/SPHB021 Co-requisites SMTA021
Module Code SMTA021 Module Code

Prof TE Mosuang
Module Lecturer
Q-3043
Office Address
thuto.mosuang@ul.ac.za -3576
Email Telephone No.
Module evaluator/ Dr SP Ntoahae (x3341)
Moderator/2nd
Examiner
Monday-Friday Mon: 09H20-11H00
Consultation Time 14H00-17H00 Lecture Periods Fri: 09H20-11H00
WhatsApp group
06/03/2023 test 1 160 hours
Important Dates 29/03/2023 test 2 Learning Hours
1st quarter/1st semester
Quarter/Semester

Module Structure No. of Lectures: 24


No. of Practical Sessions: 5
Assessment
Method Description Weighting
Test 30%
Assignments 10%
Practicals 20%
Exam 40%

MODULE DESCRIPTION
This course is fundamental to the understanding of physical systems at an atomic and sub-atomic
level. Through the wave functions students can build a relationship between the particle and wave
nature of particles. Students are also introduced to probabilities in order to determine physical
quantities like position, momentum, and energy. This course also acts as an introduction to nuclear,
particle, high energy, string theory and quantum theory physics. A strong mathematical background
of the second year level is essential.
MODULE OBJECTIVES
This course is fundamental to the understanding of the principles which describe the physical
systems. It is more fundamental than the classical mechanics and the classical field theory. These
principles are the dual wave-like and particle-like nature of particles and radiation, and the prediction
of probabilities. The connection of the course content with other Physics courses and to on-going
experiments at leading global research facilities shall be emphasized.
Students shall be acquainted with modern phenomenology of quantum mechanics as well as with
necessary theoretical tools to understand them from a fundamental point of view. A continuous
comparison of the classical and quantum approach will always be emphasized. Cumulated theory
will assist students in various theoretical and experimental research.

MODULE CONTENT

1. A Preview of Quantum Mechanics


 Classical mechanics is an approximation of quantum mechanics
 Probabilities of wave functions
 Normalization
2. Planck’s Constant in Action
 Photons
 De Broglie waves
 Atoms
 The uncertainty principle
 Measurement and wave-particle duality
 Measurement and non-locality
3. Schrödinger’s Wave Equation
 Sinusoidal wave
 Linear superposition of sinusoidal waves
 Dispersive and non-dispersive waves
 A wave equation for a free particle
 A wave equation for a particle in a potential energy field
4. Operators, Expectation values and Eigenvalues
 operators,
 Expectation values, <x>, <p>, <E>
5. Fourier Transforms
 Fourier transform as an operator
 Dirac delta functions
6. Particle in a
 Potential well
 Potential barrier
 The tunnelling effect
7. The Harmonic Oscillator
 The classical oscillator
 The quantum oscillator
 Quantum states
 Stationary and non-stationary states
 Diatomic molecules
8. The Hydrogen Atom
 Symmetry
 Radial wave equation
 Angular equations
 Interpretation of quantum numbers
 Introduction to angular momentum

LEARNING OUTCOMES
After complexion of this course students are expected to:
 describe the wave nature and particle nature of particles and radiation in general.
 Know the Schrodinger equation as fundamental equation for solving physical
problems.
 solve problems concerning particles in various potentials.
 make a connection between quantum mechanics and high energy physics
experiments.
 have a good understanding of the fundamental interactions in nature.
 Bridge the gap from classical mechanics to quantum mechanics.

ASSESSMENT CRITERIA
The students should be able to:
 apply wave functions to find particles properties.
 use graphical methods to describe particles properties.
 use the Schrodinger equation in various potential problems.
 use graphical methods to describe particles in various potentials.
 project a quantum mechanical solution to a classical situation.
 distinguish a quantum mechanical from a classical mechanical problem.
 quantify a central potential to different types of atoms.

REFERENCE MATERIALS FOR THE MODULE


1. Introduction to Quantum Mechanics – AC Phillips
2. Quantum Mechanics – F. Mandl
3. Concepts of Modern Physics – A. Beiser
4. en.wikipedia.org/wiki/quantum_mechanics

STUDENT FEEDBACK ON MODULE

Student feedback on the module occurs on written assignments and tests in a form of tutorials every
Wednesday (07H30-09H10).

MODULE POLICY (Including plagiarism, academic honesty, attendance etc.)

 Lecture attendance is compulsory


 Cheating and plagiarism is forbidden
 Submission of assignments and writing of test is mandatory

ADDITIONAL MODULE INFORMATION


I present this QM in powerpoint presentation form in lectures and online through blackboard
collaborate.
I give out notes in class in relation to what I teach and the scope of the syllabus.
Additional information pertaining to this course is obtainable from the blackboard
noticeboard.

TENTATIVE SCHEDULE OF LECTURES

Date Topic/Activity
Jan 30 – Feb 03 A Preview of Quantum Mechanics
 Classical mechanics is an approximation of quantum
mechanics
 Probabilities of wave functions
 Normalization
Planck’s Constant in Action
 Photons
 De Broglie waves
 Atoms
 The uncertainty principle
 Measurement and wave-particle duality
Measurement and non-locality

Feb 06 - 10 Schrödinger’s Wave Equation


 Sinusoidal wave
 Linear superposition of sinusoidal waves
 Dispersive and non-dispersive waves
 A wave equation for a free particle
 A wave equation for a particle in a potential energy field

Feb 13 - 17 Operators, Expectation values and Eigenvalues


 operators,
 Expectation values, <x>, <p>, <E>
Fourier Transforms
 Fourier transform as an operator
 Dirac delta functions

Feb 20 – 24 Particle in a
 Potential well
 Potential barrier
 The tunnelling effect

Feb 27 – Mar 03 The Harmonic Oscillator


 The classical oscillator
 The quantum oscillator
 Quantum states
 Stationary and non-stationary states
 Diatomic molecules

Mar 06 - 10 The Hydrogen Atom


 Symmetry
 Radial wave equation
 Angular equations
 Interpretation of quantum numbers
 Angular momentum
Mar 13 – 17 Identical Particles
 Exchange symmetry
 Exchange symmetry with spin
 Bosons and Fermions

Fundamental Quantum Mechanics concepts

Photons are particle-like quanta of electromagnetic radiation. They travel at the speed of light 𝑐 with
momentum 𝑝 and energy 𝜀 given by:
ℎ ℎ𝑐
𝑝 = 𝜆 and 𝜀 = 𝜆
where 𝜆 is the wavelength of the electromagnetic radiation. In reference to the
classical mechanics, the momentum and energy of a photon are very small.

e.g. the momentum and energy of a visible region photon of wavelength 𝜆 = 663 nm are

𝑝 = 10−27 Js and 𝜀 = 3 × 10−19 J.

It must be noted that the electronvolt, 1 eV = 1.602 x 10-19J is a useful unit for a photon. Visible
region photons have energies in the region of eV, while the X-rays have energies in the region of
keV.

The evidence for the existence of photons emerged in the early 1900s, with AH Compton giving
evidence that the wavelength of an X-ray increases when it is scattered by an electron. This effect
which was later known as the Compton effect, was understood as X-ray-electron collision in which
momentum and energy of this system is conserved. The phenomenon as shown in the Figure 1.1
below, illustrate an incident photon which transfer momentum

Figure 1.1 Schematic representation of the Compton effect

to a stationary electron, so that the scattered photon has a lower momentum and hence longer
wavelength. When a photon is scattered by an electron of mass 𝑚𝑒 at rest through an angle 𝜃, the
wavelength is increased by

∆𝜆 = 𝑚 𝑐 (1 − cos 𝜃). 1.1
𝑒


It must be noted that this increase is set by a constant 𝑚𝑒 𝑐
= 2.43 × 10−12 m called the Compton
wavelength of the electron.

The photon concept provides a natural explanation of the Compton effect and the other particle-like
electromagnetic phenomena like the photo-electric effect. In spite of this it is still unclear how the
photon can account for the wave-like property of the electromagnetic radiation. This difficulty is
illustrated by the Thomas Young’s (1801) double slit experiment used to measure the wavelength of
light. The assembly of the double slit interference experiment is shown in Figure 1.2. When
electromagnetic radiation passes through a double slit an interference pattern occurs on the screen.
This interference pattern occurs simply because the electromagnetic radiation in a form of waves
interferes constructively and destructively on the screen. But a particle approach to the interference
pattern can be that it is innumerable photons which arrive at different points on the screen. The
photons are not behaving like classical particles with well-defined trajectories. Instead, when
presented with two trajectories, photons seem to pass through both slit with equal probability, then
arrive at random points on the screen and accumulate an interference pattern. At first opinion the
particle-like and the wave-like properties of a photon is not usual. But it must be understood that
electrons, protons, atoms, and molecules all behave in this bizarre manner.

Figure 1.2 Schematic representation of the interference patterns through the double slit experiment.

de Broglie waves

The fact that particles of matter like electrons could also have particle-like and wave-like property
was first proposed by Louis de Broglie in 1923. Specifically he proposed that a particle of matter with
momentum 𝑝 could have wavelength defined by:

𝜆 = 𝑝. 1.2

This wavelength is thereafter called the de Broglie wavelength.

Sometime it is useful to write the de Broglie wavelength in terms of the energy of the particle. A
general relativistic relation between the momentum 𝑝 and energy 𝜀 of a particle of mass 𝑚 is

𝜀 2 − 𝑝2 𝑐 2 = 𝑚2 𝑐 4 . 1.3

So the de Broglie wavelength of a particle with relativistic energy 𝜀 is given by


ℎ𝑐
𝜆= . 1.4
√(𝜀−𝑚𝑐 2 )(𝜀+𝑚𝑐 2 )

If the particle is ultra-relativistic, the mass energy 𝑚𝑐 2 term can be neglected to get
ℎ𝑐
𝜆= , 1.5
𝜀

an expression which agree with the energy and the wavelength relation of the photon.

If the particle is non-relativistic, then its energy is

𝜀 = 𝑚𝑐 2 + 𝐾𝐸 , 1.6

𝑝2
and the particles non-relativistic kinetic energy is 𝐾𝐸 = 𝐸 = 2𝑚, and the particles wavelength is


𝜆= . 1.7
√2𝑚𝐸

Practically, the de Broglie wavelength of the particle of matter is small and difficult to measure. It
can be seen from the above equation that particles of smaller mass 𝑚 have longer wavelength. This
implies that the lightest particles of matter, like electrons will be the easiest to measure. The
wavelength of an electron can be obtained by substituting 𝑚 = 𝑚𝑒 = 9.109 × 10−31kg into
equation (1.7). Expressing the kinetic energy 𝐸 in electronvolts we get

1.5
𝜆 = √ 𝐸 nm. 1.8

Example: Consider an electron moving with kinetic energy 15 keV. So the de Broglie wavelength is
given by


𝜆=
√2𝑚𝐸

ℎ2𝑐 2
= √2𝑚𝑐 2 𝐸

(1.240×10−6 eV.m)2
= √2(0.511×106 eV)(15×103 eV)
1.540×10−12 eV2 m2
=√
1.533×1010 eV2

= √1.004 × 10−22 m2

= 1.002 × 10−11 = 0.010 nm

This equation shows that an electron with energy 1.5 eV has a wavelength of 1 nm, electron with
energy of 15 keV has a wavelength of 0.01 nm.

As these wavelengths are comparable with the interatomic distances in crystalline solids, electrons
with energy in the eV to keV range are diffracted by crystal lattices. The first experiments to
demonstrate the wave properties of electrons were crystal diffraction experiments by CJ Davisson
and LH Germer and independently by GP Thomson in 1927. These experiments showed beyond any
reasonable doubt that electrons can behave like waves, with wavelength given by de Broglie
wavelength equation 1.7.

Atoms

Generally, atoms can exist in states with discrete or quantized energy. A simple example is the
energy levels of a hydrogen atom which has an electron and a proton as shown in the Figure 1.3
below.

Figure 1.3 Representation of an electron or proton energy levels in a hydrogen atom.

At a later stage, bound states of electrons and protons in an atom will be shown to be given by
13.6
𝐸=− 𝑛2
eV, 1.9

where 𝑛 is the principal quantum number which takes on the infinite number of values 1, 2, 3, … The
ground state of hydrogen has 𝑛 = 1, the first excited state has 𝑛 = 2, and so on. When the
excitation energy is above 13.6 eV, the electron is no longer bound to the proton. The atom is now
ionized, its energy can in principle take any value in the continuum from 𝐸 = 0 to 𝐸 = ∞.
The existence of quantized energy levels is demonstrated by the existence of electromagnetic
spectra with sharp spectral lines that an atom makes when making a transition between two
quantized energy levels. For example, a transition in a hydrogen atom from a lower state 𝑛𝑖 to a
higher state 𝑛𝑓 , results in a spectral line with wavelength given by

ℎ𝑐
𝜆
= |𝐸𝑛𝑖 − 𝐸𝑛𝑓 |. 1.10

Some of the spectral lines of atomic hydrogen are given in Figure 1.4 below.

Figure 1.4 A diagram showing the spectral lines in a hydrogen atom.

Atoms have a wide variation in chemical properties, but amazingly there is a very small different in
atomic size. For example a mercury (Hg) atom with 80 electrons is only three times bigger than a
hydrogen atom having only one electron.

So to a certain extent, atomic properties can be best described using wave-like property of matter.
For instance, in music, when strings of a guitar vibrate with a definite frequency; they form a
standing wave of a given shape. Similarly, electrons in an atom have well defined a wavelength
which confines them in a specific atom.

Another startling property of electrons in atoms, is that they are entangled, which means generally it
is impossible to differentiate one electron from another. The only solution is that of applying the
Paulis exclusion principle for electron of an atom with say atomic number Z.

All of these ideas will be discussed in details in the coming chapters. But currently the main aim is to
show that the wave nature of atomic electrons can give an accepted explanation of the size of an
atom. Because the de Broglie wavelength of an electron depend on the Planck’s constant ℎ and the
electrons mass 𝑚𝑒 , the size of an atom consisting of wave-like electrons also depends on ℎ and 𝑚𝑒 .
In addition, the atomic size must depend on the force that binds the electron to the nucleus of an
𝑒2
atom: that is proportional to 4𝜋𝜖 , where 𝑒 is the magnitude of charge of an electron and a proton.
0
𝑒2
So specifically, the size of an atom depends on 4𝜋𝜖 , ℎ and 𝑚𝑒 . Moreover, the accepted unit of
0
length at atomic scale is the Bohr radius which is given by

4𝜋𝜖0 ℏ2
𝑎0 = [ ]
𝑒 2 𝑚𝑒
= 0.529 × 10−10 m. 1.11

Form the natural unit of length 𝑎0 , the natural unit of atomic binding energy called the Rydberg
energy is given by

𝑒2
𝐸𝑅 = 8𝜋𝜖 = 13.6 eV. 1.12
0 𝑎0

The uncertainty principle

The uncertainty principle is introduced using a mind experiment formulated by Werner Heisenberg.
In this experiment the uncertainty in position and momentum of a particle is measured using a
microscope. The particle is illuminated by incident light, and the scattered light is collected by the
lens of the microscope as shown in Figure 1.5 below.

Figure 1.5 An illustration of the observation of a particle using Heisenberg miscroscope.

Because of the wave-like properties of light, the microscope has a finite spatial resolution power.
This means the observed particle has an uncertainty in position given by

𝜆
∆𝑥 ≈ sin 𝛼 where 𝜆 is the wavelength of the incident light and 2𝛼 is the angle
subtended by the lens at the particle. The resolution can be increased by decreasing the wavelength
of the incident light on the particle; visible light waves are better than the infrared waves, and the
ultraviolet waves are better than the visible light waves.
On the other hand, due to particle-like properties of light the process involves after scattering the
numerous photon-particle like collision, with the scattered photons entering the lens of the
microscope. To enter the lens, the photons with wavelength 𝜆 and momentum ℎ/𝜆 must have
sideways momentum components

− ℎ⁄𝜆 sin 𝛼 and + ℎ⁄𝜆 sin 𝛼

So the sideways momentum of the scattered photon is uncertain by

∆𝑝 ≈ ℎ⁄𝜆 sin 𝛼. 1.13

Due to the conservation of momentum during the scattering, the observed particles possess the
uncertainty in momentum which is the same as that of the scattered photon.

The uncertainty in the particles momentum can be reduced by increasing the wavelength of the
incident light. But this will cause a poor spatial resolution of the microscope resulting in an increase
in the uncertainty of the particles position. By combining the uncertainty in position with that in
momentum we find the following relation

Δ𝑥Δ𝑝 ≈ ℎ. 1.14

This result is called the Heisenberg uncertainty principle. It asserts that greater accuracy in
measuring a particles position is possible, but only at the expense of greater uncertainty in
measuring the particles momentum and greater accuracy in measuring the particles momentum is
possible, but only at the expense of greater uncertainty in measuring the particles position. In
particular, the principle explains that the fundamental uncertainties in the simultaneous knowledge
of the position and momentum of a particle obey the inequality

ℏ ℎ
∆𝑥∆𝑝 ≥ where ℏ = . 1.15
2 2𝜋

The Heisenberg uncertainty principle suggest that an exact determination of the particles position,
i.e. ∆𝑥 = 0, is possible provided there is a total uncertainty of the particles momentum. Actually, an
analysis of the microscope mind experiment, considering the Compton effect, shows that an exact
determination of the particles position is impossible!

The Free Particle

 a free particle is the one which moves through space without experiencing any force.
 hence it travels in a straight line.
 its potential is everywhere constant and can be assigned zero magnitude!
 the energy states are not quantized, but any value is allowed.
 wavefunction solutions are imaginary exponentials, indicating an oscillating amplitude in
space and time.
The Schrӧdinger Equation

In this course, a non-relativistic quantum mechanics will be thrashed out. In order to achieve that
goal, we need to set up a wave equation required to describe a wave-particle property of quantum
particles. The equation contemplated is the Schrӧdinger wave equation. We need to understand
that the role played by Schrӧdinger equation in quantum mechanics is the same as the role played
by Newton’s Laws equations in classical mechanics. Both equations describe the motion.

Newtons’ 2nd Law is a second order differential equation which describes how a classical particle
moves in space.

𝐹⃗ = 𝑚𝑎⃗ = 𝑚𝑟⃗̈. 2.1

Schrӧdinger equation is a second order partial differential equation which describes how the
wavefunction corresponding to a particle propagates in space.

It must be mentioned that both the equations were postulated and then experiments were
performed to test their validity.

Sinusoidal Waves

The well-designed wave is a sinusoidal wave travelling with wavelength 𝜆 and period 𝜏. In the same
way this can be represented by the wave number 𝑘 = 2𝜋⁄𝜆 and angular frequency 𝜔 = 2𝜋⁄𝜏. Such
a wave is represented by a mathematical function

𝜓(𝑥, 𝑡) = 𝐴 cos(𝑘𝑥 − 𝜔𝑡). 2.2


2𝜋
𝐴 is a constant for the amplitude. At point 𝑥, the function oscillates with amplitude 𝐴 and period 𝜔
.
2𝜋
At time 𝑡, the function undulates with amplitude 𝐴 and wavelength 𝑘
. On top of that these
undulations the wave function exhibit a Mexican wave, in the direction of the increasing 𝑥 with
velocity 𝜔/𝑘. For the case in point, the maximum of 𝜓(𝑥, 𝑡) corresponds to 𝑘𝑥 − 𝜔𝑡 = 0, which
occurs at position 𝑥 = 𝜔𝑡/𝑘. The minimum of 𝜓(𝑥, 𝑡) corresponds to 𝑘𝑥 − 𝜔𝑡 = 𝜋 and it occurs at
𝜆 𝜔𝑡
position 𝑥 = 2 + 𝑘
. In both cases the position moves with velocity 𝜔/𝑘.

The function sin(𝑘𝑥 − 𝜔𝑡), like cos(𝑘𝑥 − 𝜔𝑡), represents a sinusoidal travelling wave with wave
number 𝑘 and angular frequency 𝜔. Since we know that sin(𝑘𝑥 − 𝜔𝑡) = cos(𝑘𝑥 − 𝜔𝑡 − 𝜋/2), the
undulations and oscillations of sin(𝑘𝑥 − 𝜔𝑡) are out of step in relation to those of cos(𝑘𝑥 − 𝜔𝑡). So
the waves of sin(𝑘𝑥 − 𝜔𝑡) and cos(𝑘𝑥 − 𝜔𝑡) have a phase difference of 𝜋/2.

Now the most general sinusoidal travelling wave having wave number 𝑘 and angular frequency 𝜔 is
the linear superposition of the two waves to give:

𝜓(𝑥, 𝑡) = 𝐴 cos(𝑘𝑥 − 𝜔𝑡) + 𝐵 sin(𝑘𝑥 − 𝜔𝑡),

𝐴 and 𝐵 are constants.

Occasionally in classical physics and invariably in quantum physics, sinusoidal travelling waves are
represented by the complex exponential function

𝜓(𝑥, 𝑡) = 𝐴𝑒 𝑖(𝑘𝑥−𝜔𝑡) . 2.3

In classical physics this exponential representation is just for mathematical convenience. For
instance, the pressure in sound waves could be regarded as the real function 𝐴 cos(𝑘𝑥 − 𝜔𝑡), but
the real function can be taken from the real part of a complex exponential function 𝐴𝑒 𝑖(𝑘𝑥−𝜔𝑡)
because

𝐴𝑒 𝑖(𝑘𝑥−𝜔𝑡) = 𝐴[cos(𝑘𝑥 − 𝜔𝑡) + 𝑖 sin(𝑘𝑥 − 𝜔𝑡)].

So in classical physics we have the option of representing a real sinusoidal wave by the real part of a
complex exponential.

In quantum physics, the use of complex exponential function is not an option. Complex exponential
functions provide a natural description of a de Broglie wave.

Linear superposition of sinusoidal waves

Two sinusoidal waves travelling in opposite directions may be combined to form standing waves.
Take for instance, the linear superposition

𝐴 cos(𝑘𝑥 − 𝜔𝑡) + 𝐴 cos(𝑘𝑥 + 𝜔𝑡) give rise to the wave 2𝐴 cos 𝑘𝑥 cos 𝜔𝑡.

This wave oscillates with period 2𝜋/𝜔 and undulates with wavelength 2𝜋/𝑘. But these oscillations
and undulations do not propagate, they are just standing waves. Many sinusoidal waves may be
combined to form a wave packet. For example, the mathematical form of a wave packet formed by
linear superposition of sinusoidal waves with constant amplitude 𝐴 and wave numbers in the range
𝑘 − Δ𝑘 to 𝑘 + Δ𝑘 is

𝑘+Δ𝑘
𝜓(𝑥, 𝑡) = ∫𝑘−Δ𝑘 𝐴 cos(𝑘′𝑥 − 𝜔′𝑡)𝑑𝑘′. 2.4

If 𝑘 is positive, the wave packet travels in the positive 𝑥 direction, and to the negative 𝑥 direction
when 𝑘 in negative.

The initial shape of the wave packet, i.e. a shape at 𝑡 = 0, may be obtained by evaluating the
integral
𝑘+Δ𝑘
𝜓(𝑥, 0) = ∫𝑘−Δ𝑘 𝐴 cos 𝑘′𝑥𝑑𝑘′ 2.5

sin ∆𝑘𝑥
this gives 𝜓(𝑥, 0) = 𝒮(𝑥) cos 𝑘𝑥 where 𝒮(𝑥) = 2𝐴∆𝑘 (∆𝑘𝑥)
.

Dispersive and non-dispersive waves

A good example of non-dispersive waves is electromagnetic waves travelling in a vacuum.

A non-dispersive wave has a dispersion relation of the form 𝜔 = 𝑐𝑘, where 𝑐 is a constant such that
for sinusoidal waves, the velocity, 𝑐 = 𝜔/𝑘, is independent of the wave number 𝑘.

A wave packet formed from such sinusoidal waves travels without changing shape because each
sinusoidal component has the same velocity.

The non-dispersive waves are described by a partial differential equation called classical wave
equation. In one dimension along the 𝑥 direction, the wave equation has the form

𝜕2 𝜓 1 𝜕2 𝜓
= , 2.6
𝜕𝑥 2 𝑐 2 𝜕𝑡 2

1 𝜕2 𝜓 𝜕2 𝜕2 𝜕2
and in three dimensions ∇2 𝜓 = , where ∇2 = + + .
𝑐 2 𝜕𝑡 2 𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

The classical wave equation has infinite number of solutions which corresponds to different wave
forms. For instance, the sinusoidal waves has

𝐴 cos(𝑘𝑥 − 𝜔𝑡), 𝐴 sin(𝑘𝑥 − 𝜔𝑡), or 𝐴𝑒 𝑖(𝑘𝑥−𝜔𝑡) 2.7

as solutions provided 𝜔2 = 𝑐 2 𝑘 2 , which can be shown by direct substitution on the classical wave
equation.

Solutions with 𝑐 = +𝜔/𝑘, describe waves travelling in the positive 𝑥 direction, whilst solutions with
𝑐 = −𝜔/𝑘, describe waves travelling in the negative 𝑥 direction. Since each term representing the
sinusoidal wave is a solution to the classical wave equation, even the linear superposition of these
terms is also a solution to the classical wave equation. The example is the superposition term in
equation (..), which describes a wave packet travelling without changing its shape.

Although, under normal circumstances, most of the waves encountered in classical and quantum
physics are dispersive waves. The partial differential equation describing dispersive waves is more
complicated than the one discussed for classical wave equation, eq (..). The dispersion relation, 𝜔 =
𝑐𝑘 is more complicated so that the velocity, 𝜔/𝑘 of a propagating wave depends on the wave
number, 𝑘. This explains why the packet of the dispersive waves generally does change in shape as
they propagate in space. But, if the packet consists of a narrow range of wave numbers, it can have a
well-defined velocity of propagation. This velocity is called the group velocity and is given by

𝑑𝜔
𝜐𝑔𝑟𝑜𝑢𝑝 = , 2.8
𝑑𝑘

though the velocity of a simple sinusoidal wave is called the phase velocity
𝜔
𝜐𝑝ℎ𝑎𝑠𝑒 = . 2.9
𝑘

To obtain a clear understanding of eq (..), you must note that the group velocity describes the
motion of a localized disturbance resulting from constructive interference of two sinusoidal wave.
Consider a constructive interference resulting from two sinusoidal waves with wave numbers 𝑘1 and
𝑘2 and angular frequencies 𝜔1 and 𝜔2 , mathematically this is

𝑘1 𝑥 − 𝜔1 𝑡 = 𝑘2 𝑥 − 𝜔2 𝑡.

The point of constructive interference is obtained by rearranging the above equation to give

𝜔1 −𝜔2
𝑥=( )𝑡 . 2.10
𝑘1 −𝑘2

So, the point of constructive interference is located at 𝑥 = 0, when 𝑡 = 0 and it travels with a
velocity given by 𝜔1 − 𝜔2 /𝑘1 − 𝑘2 , or eq (..) if |𝑘1 − 𝑘2 | is small. Obviously, for two sinusoidal
waves there are, infinite number of constructive interference points, but for many sinusoidal waves
a single localized region travelling with velocity, 𝑑𝜔⁄𝑑𝑘 can be obtained.

Example: To illustrate how group velocity can be derived from 𝜔 = 𝑐𝑘, consider water waves of
long wavelength obeying the dispersion relation

𝜔 = √𝑔𝑘 , where 𝑔 is the acceleration due to gravity.

The phase velocity of the sinusoidal wave is

𝜔 𝑔
𝜐𝑝ℎ𝑎𝑠𝑒 = =√ .
𝑘 𝑘

The velocity of the wave packet, with a narrow range of wave numbers near 𝑘 is

𝑑𝜔 1 𝑔
𝜐𝑔𝑟𝑜𝑢𝑝 = = √ .
𝑑𝑘 2 𝑘

So for water waves the group velocity is one-half of the phase velocity. In other words, the sinusoidal
waves forming a wave packet travel at twice the speed of the region of maximum disturbance
caused by the interference of these waves. The shape of the disturbance will change as it
propagates, which means it tends to spread out.

Particle Wave Equations

In classical physics, fundamental laws of physics are used to derive the wave equations which
describe the wave-like phenomena of electromagnetic waves. Take for instance; the Maxwell’s laws
of electromagnetism are used to derive the classical wave equation which presides over
electromagnetic waves in a vacuum. It can also be seen that the wave equation which describes
wave-like properties of quantum particles is a fundamental equation, which cannot be derived from
basic physical laws.
Like the inventors of quantum theory, the form of the particle wave equation can only be guessed
and then tested for consistency and agreement with experiments.

Figure 1.6 initial shapes of wave packets given by linear superposition of sinusoidal waves with amplitude A in the wave
number range 𝒌 − ∆𝒌 to 𝒌 + ∆𝒌.

Free particle wave equation

-we need to construct a possible wave equation for a freely moving non-relativistic particle by
considering the properties of the de Broglie’s waves.

- a particle with momentum 𝑝 has de Broglie wavelength given by 𝜆 = ℎ/𝑝.

- this means that a de Broglie wave with wave number 𝑘 = 2𝜋/𝜆 describes a particle with
momentum

𝑝 = ℏ𝑘 , where ℏ = 2𝜋

- to extend this argument, assume that a de Broglie wave packet with a range of wave
numbers from 𝑘 − Δ𝑘 to 𝑘 + Δ𝑘 describes a particle with an uncertain momentum
Δ𝑝 ≈ ℏ∆𝑘
- assume also that the length of this wave packet is a measure of Δ𝑥, the uncertainty in the
position of the particle.
2𝜋
∆𝑥 ≈
∆𝑘
- using the equation Δ𝑝 ≈ ℏ∆𝑘 and the figure …. as a model, we write
- multiplying these two uncertainties we obtain
∆𝑥∆𝑝 ≈ ℎ
- this means a de Broglie wave packet can account for the uncertainties in the position and
momentum of a quantum particle.
- this also means that a de Broglie wave is transformed by a measurement.
- if a precise measurement of the position is made, the new packet describing the particle will
be very short, a superposition of sinusoidal waves with a wide range of wavelengths.
- if a precise measurement of momentum is made, the new wave packet will be long with a
sharply defined wavelength.
- this shows that a wave packet is a delicate entity which is transformed by measurements.
This is still a mystery on how it happens.
- the notion that the wave packet represents a moving quantum particle is being imposed.
- the fact that the group velocity of the packet equals the velocity of a particle with mass 𝑚
and momentum 𝑝 = ℏ𝑘 requires that
𝑑𝜔 ℏ𝑘
=
𝑑𝑘 𝑚
- this equation can be integrated to give the following dispersion relation for the de Broglie
waves describing a free moving quantum particle of mass 𝑚:
ℏ𝑘 2
𝜔= . 2.11
2𝑚
- in order to obtain the above relation the constant of integration is set to zero.
- in any case this constant give rise to no observable consequences in the non-relativistic
quantum mechanics.
- the main goal is to obtain a wave equation with sinusoidal solutions obeying the above
dispersion relation.
- the simplest such a wave equation is called the Schrӧdinger equation.
- for a free particle moving in one dimension, it has the form
𝜕𝜓 ℏ2 𝜕 2 𝜓
𝑖ℏ =− . 2.12
𝜕𝑡 2𝑚 𝜕𝑥 2
- the complex exponential function 𝜓(𝑥, 𝑡) = 𝐴𝑒 𝑖(𝑘𝑥−𝜔𝑡) is the solution of this
ℏ𝑘 2
equation provided it obey the dispersion relation 𝜔 = .
2𝑚
- substituting 𝜓(𝑥, 𝑡) on the left hand side of eq … yields
𝜕𝜓
𝑖ℏ 𝜕𝑡 = 𝑖ℏ(−𝑖𝜔)𝐴𝑒 𝑖(𝑘𝑥−𝜔𝑡) = ℏ𝜔𝐴𝑒 𝑖(𝑘𝑥−𝜔𝑡) = ℏ𝜔𝜓(𝑥, 𝑡)
- the right hand side yields
ℏ2 𝜕2 𝜓 ℏ2 𝜕2 ℏ2 ℏ2 𝑘 2
− 2𝑚 𝜕𝑥 2 = − 2𝑚 𝜕𝑥 2 𝐴𝑒 𝑖(𝑘𝑥−𝜔𝑡) = − 2𝑚 (−𝑘 2 )𝐴𝑒 𝑖(𝑘𝑥−𝜔𝑡) = 2𝑚
𝜓(𝑥, 𝑡)
- the solution can be obtained provided
ℏ2 𝑘 2
ℏ𝜔 = . 2.13
2𝑚
- the sinusoidal wave solution, 𝐴𝑒 𝑖(𝑘𝑥−𝜔𝑡) , describes a wave moving in the x-direction with
wave number 𝑘 and angular frequency 𝜔.
- we assume that the wave represents a free particle moving in the x-direction with
𝑝2
momentum 𝑝 = ℏ𝑘 and energy 𝐸 = = ℏ𝜔. 2.14
2𝑚
- of course, there are many other solutions of the Schrӧdinger equation representing other
states of motion of the particle.
- it is necessary to point out that in order to oblige with the dispersion relation for the de
ℏ𝑘 2
Broglie waves, 𝜔 = 2𝑚
, we arrived at a wave equation, the free-particle Schrӧdinger
𝜕𝜓 ℏ 𝜕2 𝜓
2
equation, 𝑖ℏ 𝜕𝑡 = − 2𝑚 𝜕𝑥 2 , whose solutions are necessarily complex functions of space and
time.
- these complex functions are called wave functions.
- recall that classically waves are often represented by complex functions, as a matter of
mathematical convenience.
- in the practical sense, classical waves normally are real functions of space and time.
- alternatively, Schrӧdinger wave functions are not real functions of space and time.
- they are complex functions which describe the concealed wave-like property of a quantum
particle.
Since each term in the Schrӧdinger equation is linear in the wave function 𝜓, a superposition of
solutions is also a solution. e.g.

𝜓(𝑥, 𝑡) = 𝐴1 𝑒 𝑖(𝑘1 𝑥−𝜔1 𝑡) + 𝐴2 𝑒 𝑖(𝑘2 𝑥−𝜔2 𝑡) 2.15

ℏ2 𝑘12 ℏ2 𝑘22
with ℏ𝜔1 = and ℏ𝜔2 = ,
2𝑚 2𝑚

where 𝐴1 and 𝐴2 are arbitrary constants, is the solution of equation (2.15).

-this can easily be confirmed by direct substitution.

-The most general solution is the superposition of sinusoidal waves with all possible wave numbers
and angular frequencies:

+∞ ′ 𝑥−𝜔′ 𝑡)
𝜓(𝑥, 𝑡) = ∫−∞ 𝐴(𝑘 ′ )𝑒 𝑖(𝑘 𝑑𝑘′ 2.16

ℏ2 𝑘′2
with ℏ𝜔′ = .
2𝑚

Here, 𝐴(𝑘 ′ ) is an arbitrary complex function of 𝑘′, and the integral is over all possible wave numbers
𝑘′.

Wave equation for a particle in a potential energy field

For non-relativistic particles, the interactions can be explained using the potential energy field. For
example the electron in a hydrogen atom can be regarded as moving in a spherical potential energy
field

𝑉(𝑟) = −𝑒 2 /4𝜋𝜖0 𝑟 2.17

of the nucleus in a hydrogen atom.


Classically this means that an electron at a distance 𝑟 from the nucleus experiences an attractive
force 𝑒 2 /4𝜋𝜖0 𝑟 2 due to the nucleus.

Quantum mechanically, this means that the wave equation for an electron is not a simple free
particle wave equation given by equation (2.12).

In 1926, Erwin Schrӧdinger invented wave equation for a quantum particle in the presence of a
potential energy field which led to a successful description of electrons, protons, atoms, molecules,
etc. It is a generalization of a free particle wave equation discussed in the previous section (equation
(2.12)).

The Schrӧdinger equation for a particle moving in a three- dimensional potential energy field 𝑉(𝑟⃗) is
given by

𝜕𝜓 ℏ2
𝑖ℏ = [− ∇2 + 𝑉(𝑟⃗)] 𝜓. 2.18
𝜕𝑡 2𝑚

In a one dimensional potential 𝑉(𝑥) along the 𝑥, Schrӧdinger equation simplifies to

𝜕𝜓 ℏ2 𝜕2
𝑖ℏ 𝜕𝑡 = [− 2𝑚 𝜕𝑥 2 + 𝑉(𝑥)] 𝜓. 2.19

The solutions of this Schrӧdinger wave equation can be obtained easily if the potential energy field is
a constant. For instance, if a particle travels along the 𝑥, with a constant potential 𝑉0 its solution is

𝜓(𝑥, 𝑡) = 𝐴𝑒 𝑖(𝑘𝑥−𝜔𝑡) 2.20

ℏ2 𝑘 2
provided ℏ𝜔 = 2𝑚
+ 𝑉0. 2.21

The wave function represents a well-defined particle having total energy and momentum given by:

𝑝2
𝐸 = 2𝑚 + 𝑉0 and 𝑝 = ℏ𝑘.

Probabilities
The importance of probability in quantum measurements forces the consideration to describe
discrete and continuous random variables in probability distributions. These considerations can be
illustrated by Gaussian, Poisson, and exponential probability distributions.

Discrete random variables

Consider a process or experiment with possible outcomes described by discrete random variables
which take on values 𝑥0 , 𝑥1 , 𝑥2 , 𝑥3 … with probabilities 𝑝0 , 𝑝1 , 𝑝2 , 𝑝3 , … The set of probabilities 𝑝𝑛 is
called a probability distribution. A total probability of all possible distribution must equal to unity.
Therefore the probability distribution 𝑝𝑛 must satisfy the normalization condition

∑𝑛 𝑝𝑛 = 1. 3.1
For example if the probability 𝑝𝑛 that a student who is doing this module is doing other 𝑛 modules
to complete his Physics course is

𝑝0 + 𝑝1 + 𝑝2 + 𝑝3 = 1.

The probability distribution 𝑝𝑛 can be used to evaluate the expectation values for the random
variables 𝑥𝑛 . This is the average value of the possible outcome that may occur when the process or
experiment takes places an infinite number of times. It is given by

〈𝑥〉 = ∑all 𝑛 𝑥𝑛 𝑝𝑛 . 3.2

The spread in the outcomes about this expectation value is given by the standard deviation or the
uncertainty in 𝑥. This uncertainty or standard deviation is denoted by ∆𝑥. The square of the standard
deviation is called the variance and is given by:

(∆𝑥)2 = ∑all 𝑛(𝑥𝑛 − 〈𝑥〉)2 𝑝𝑛 . 3.3

In this expression, (𝑥𝑛 − 〈𝑥〉) is the deviation of 𝑥𝑛 from the expected value; this deviation may be
positive or negative but its average value is zero. However, the variance is the average of the square
of this deviation; it is zero when there is only one possible outcome and it is positive when there is
more than one possible outcomes. In many cases it is useful to write the variance using the following
formulation:

(𝑥𝑛 − 〈𝑥〉)2 = 𝑥𝑛2 − 2𝑥𝑛 〈𝑥〉 + 〈𝑥〉2 , 3.4

recalling that 〈𝑥〉 is a number that does not depend on 𝑛, then

(∆𝑥)2 = ∑all 𝑛 𝑥𝑛2 𝑝𝑛 − 2〈𝑥〉 ∑all 𝑛 𝑥𝑛 𝑝𝑛 + 〈𝑥〉2 ∑all 𝑛 𝑝𝑛

= 〈𝑥 2 〉 − 2〈𝑥〉2 + 〈𝑥〉2

= 〈𝑥 2 〉 − 〈𝑥〉2.

∆𝑥 = √〈𝑥 2 〉 − 〈𝑥〉2 3.5

So in summary this equation states that the variance is the difference between the average of the
square and the square of the average of the random variables.

Continuous random variables

Consider now a process or experiment in which an outcome is described by a continuous variable 𝑥.


The probability of an outcome between 𝑥 and 𝑥 + 𝑑𝑥 is denoted by 𝜌(𝑥)𝑑𝑥. The function 𝜌(𝑥) is
called the probability density. It satisfy the normalization condition

∫all 𝑛 𝜌(𝑥)𝑑𝑥 = 1. 3.6

For example if 𝑥 is the position of a particle confined in a region 0 ≤ 𝑥 ≤ 𝑎, then


𝑎
∫0 𝜌(𝑥)𝑑𝑥 = 1.

The expectation value 〈𝑥〉 in analogy with the discrete variables is given by
〈𝑥〉 = ∫all 𝑛 𝑥𝜌(𝑥)𝑑𝑥. 3.7

Similarly, the expectation value of 𝑥 2 is given by

〈𝑥 2 〉 = ∫all 𝑛 𝑥 2 𝜌(𝑥)𝑑𝑥, 3.8

and the standard deviation or the uncertainty in 𝑥 is given by

(∆𝑥) = √〈𝑥 2 〉 − 〈𝑥〉2 . 3.5

The Born Interpretation of the Wavefunction

The fact that the wavefunction can explain the measurement of position was first proposed by Max
Born in 1926.

This is now called the Born interpretation of the wavefunction.

According to this, the wavefunction is a complex function of space coordinates whose modulus
squared |𝜓(𝑟⃗, 𝑡)|2 , is a measure of the probability of finding a particle at the point 𝑟⃗, in time 𝑡.

The particle can be found anywhere but is more likely to be found where |𝜓(𝑟⃗, 𝑡)|2 is large.

𝑡ℎ𝑒 𝑝𝑟𝑜𝑏𝑎𝑏𝑖𝑙𝑖𝑡𝑦 𝑜𝑓 𝑓𝑖𝑛𝑑𝑖𝑛𝑔 𝑡ℎ𝑒 𝑝𝑎𝑟𝑡𝑖𝑐𝑙𝑒


Specifically |𝜓(𝑟⃗, 𝑡)|2 𝑑3 𝑟⃗ = { } 3.9
𝑎𝑡 𝑡𝑖𝑚𝑒 𝑡 𝑖𝑛 𝑡ℎ𝑒 𝑣𝑜𝑙𝑢𝑚𝑒 𝑒𝑙𝑒𝑚𝑒𝑛𝑡 𝑑3 𝑟⃗

|𝜓(𝑟⃗, 𝑡)|2 , can also be thought as the probability density of position. 𝜓(𝑟⃗, 𝑡), is also regarded as the
probability amplitude for position.

If we integrate the probability density over all possible positions of the particle, we find the
probability of finding a particle somewhere in the universe.

Because the particle is to be found somewhere, this probability must be unity.

So the wavefunction must satisfy the following normalization condition

∫|𝜓(𝑟⃗, 𝑡)|2 𝑑 3 𝑟⃗ = 1 , 3.10

where the integral is over all space.

This equation is less intimidating when a particle is restricted to move in one dimension.

Say a particle moves along the 𝑥 axis then it can be described by the wavefunction 𝜓(𝑥, 𝑡) such that

𝑡ℎ𝑒 𝑝𝑟𝑜𝑏𝑎𝑏𝑖𝑙𝑖𝑡𝑦 𝑜𝑓 𝑓𝑖𝑛𝑑𝑖𝑛𝑔 𝑡ℎ𝑒 𝑝𝑎𝑟𝑡𝑖𝑐𝑙𝑒


|𝜓(𝑥, 𝑡)|2 𝑑𝑥 = { }. 3.11
𝑎𝑡 𝑡𝑖𝑚𝑒 𝑡 𝑏𝑒𝑡𝑤𝑒𝑒𝑛 𝑥 𝑎𝑛𝑑 𝑥 + 𝑑𝑥

Because the particle must be found somewhere between −∞ and +∞, it must also obey the
normalization condition
∫|𝜓(𝑥, 𝑡)|2 𝑑𝑥 = 1

Practically a wavefunction is normalized if the solution to the Schrӧdinger equation is multiplied by


an appropriate constant.

So when a new wavefunction is confronted, normalization procedure will be done, then explore the
potential positions by considering the probability density of position.

Fourier Transforms

 Fourier transform is an operator that matches functions to other functions.


 Loosely, the Fourier transform decomposes a function into a continuous spectrum of its
frequency components.
 The inverse transform synthesizes a function from its frequency components.
+∞
 If |𝑓(𝑥)| can be integrated, i.e. ∫−∞ |𝑓(𝑥)|𝑑𝑥 < ∞, then one can always write a Fourier
expansion
+∞ 𝑑𝑘
𝑓(𝑥) = ∫−∞ √2𝜋
𝑔(𝑘)𝑒 𝑖𝑘𝑥 ............................... 3.12

 𝑔(𝑘) is called the Fourier transform of the function 𝑓(𝑥) and is given by:
+∞ 𝑑𝑥
𝑔(𝑘) = ∫−∞ √2𝜋
𝑓(𝑥)𝑒 −𝑖𝑘𝑥 . ................................ 3.13

Self-consistency:
+∞ 𝑑𝑥
𝑔(𝑘) = ∫−∞ 𝑓(𝑥)𝑒 −𝑖𝑘𝑥
√2𝜋

+∞ 𝑑𝑥 −𝑖𝑘𝑥 +∞ 𝑑𝑘 ′ 𝑖𝑘 ′ 𝑥
= ∫−∞ 𝑒 ∫−∞ 2𝜋 𝑒 𝑔(𝑘 ′ )
√2𝜋 √

+∞ +∞ 𝑑𝑥 ′ −𝑘)𝑥
= ∫−∞ 𝑑𝑘 ′ 𝑔(𝑘 ′ ) [∫−∞ 2𝜋
𝑒 𝑖(𝑘 ] 3.14

+∞ 𝑑𝑥 ′ −𝑘)𝑥
 so what kind of a function is ∫−∞ 2𝜋
𝑒 𝑖(𝑘 ...?

Dirac delta function

a Dirac delta function 𝛿(𝑥 − 𝑥 ′ ) is defined by

(i) 𝛿(𝑥 − 𝑥 ′ ) = 0 if 𝑥 ≠ 𝑥′ and 𝛿(𝑥 − 𝑥 ′ ) = ∞ if 𝑥 = 𝑥′


(ii) ∫ 𝑑𝑥𝛿(𝑥 − 𝑥 ′ ) = 1
(iii) ∫ 𝑑𝑥𝑓(𝑥)𝛿(𝑥 − 𝑥 ′ ) = 𝑓(𝑥 ′ )

+∞ ′ −𝑘)𝑥
Now if ∫−∞ 𝑑𝑥𝑒 𝑖(𝑘 = 2𝜋𝛿(𝑘 ′ − 𝑘) then equation 3.14 is satisfied because

+∞ 1
∫−∞ 𝑑𝑘 ′ 𝑔(𝑘 ′ ) 2𝜋 2𝜋𝛿(𝑘 ′ − 𝑘) = 𝑔(𝑘) 3.15
so all in all we have
+∞ ′ −𝑘)𝑥
∫−∞ 𝑑𝑥𝑒 𝑖(𝑘 = 2𝜋𝛿(𝑘 ′ − 𝑘) 3.16a

+∞ ′ −𝑥)
∫−∞ 𝑑𝑘𝑒 𝑖𝑘(𝑥 = 2𝜋𝛿(𝑥′ − 𝑥) 3.16b

Property of Fourier transform


+∞ +∞ ∞ ∞
𝑑𝑘 𝑑𝑘′
∫ 𝑑𝑥|𝑓(𝑥)|2 = ∫ 𝑑𝑥 (∫ 𝑔∗ (𝑘)𝑒 −𝑖𝑘𝑥 ) (∫ 𝑔(𝑘 ′ )𝑒 𝑖𝑘𝑥 )
−∞ −∞ −∞ √2𝜋 −∞ √2𝜋

∞ 𝑑𝑘 ∞ 𝑑𝑘 ′ ∞ ′ −𝑘)𝑥
= ∫−∞ 𝑔∗ (𝑘) ∫−∞ 𝑔(𝑘 ′ ) ∫−∞ 𝑑𝑥𝑒 𝑖(𝑘
√2𝜋 √2𝜋

∞ 𝑑𝑘 ∞ 𝑑𝑘 ′
= ∫−∞ 𝑔∗ (𝑘) ∫−∞ 𝑔(𝑘 ′ ) 2𝜋𝛿(𝑘 ′ − 𝑘)
√2𝜋 √2𝜋

+∞
= ∫−∞ 𝑑𝑘𝑔∗ (𝑘)𝑔(𝑘)

+∞ +∞
⇒ ∫−∞ 𝑑𝑥|𝑓(𝑥)|2 = ∫−∞ 𝑑𝑘|𝑔(𝑘)|2 3.17

Momentum Probabilities

In this section we aim to explain how momentum properties of a particle can be explained using the
Schrӧdinger wavefunction.

If a particle wavefunction represent a particle with a selection of positions, then it can also represent
a selection of momenta.

Take for instance the wavefunction

𝜓(𝑥, 𝑡) = 𝐴1 𝑒 𝑖(𝑘1 𝑥−𝜔1 𝑡) + 𝐴2 𝑒 𝑖(𝑘2 𝑥−𝜔2 𝑡) , which describes a free particle moving along the
𝑥 direction with momenta and energies

𝑝1 = ℏ𝑘1 , 𝐸1 = ℏ𝜔1 and 𝑝2 = ℏ𝑘2 , 𝐸2 = ℏ𝜔2 . 3.18

This idea is also illustrated by the general solution of a free particle Schrӧdinger equation given by

+∞
𝜓(𝑥, 𝑡) = ∫−∞ 𝐴(𝑘)𝑒 𝑖(𝑘𝑥−𝜔𝑡) 𝑑𝑘 , 3.19
ℏ2 𝑘 2
with ℏ𝜔 = 2𝑚
.

In wave notation, this is a superposition of sinusoidal waves with different wave number 𝑘. The
magnitude of |𝐴(𝑘)|2 , measures the intensity of sinusoidal waves with wave number 𝑘.

According to particles, this range of wave numbers corresponds to a range of possible momenta
𝑝 = ℏ𝑘.

Using the same analogy as with Born interpretation of wavefunction, the most probable momenta
found in a measurement correspond to the value ℏ𝑘 for which the function |𝐴(𝑘)|2 is the largest.

In a general form, one need to treat position and momentum in a symmetrical way using Fourier
transforms.

So any wavefunction moving in one dimension can always be written as the Fourier transform
1 +∞
𝜓(𝑥, 𝑡) = ∫ 𝜓̃(𝑝, 𝑡) 𝑒 +𝑖𝑝𝑥/ℏ 𝑑𝑝.
√2𝜋ℏ −∞
3.20

The inverse Fourier transform is


1 +∞
𝜓̃(𝑝, 𝑡) = ∫ 𝜓(𝑥, 𝑡)𝑒 −𝑖𝑝𝑥/ℏ 𝑑𝑥. 3.21
√2𝜋ℏ −∞

It can also be shown that if 𝜓(𝑥, 𝑡) is normalized, then 𝜓̃(𝑝, 𝑡) is also normalized;

+∞ +∞ 2
i.e. if ∫−∞ |𝜓(𝑥, 𝑡)|2 𝑑𝑥 = 1, then ∫−∞ |𝜓̃(𝑝, 𝑡)| 𝑑𝑝 = 1. 3.22

Through symmetry between the position and momentum, and earlier observation about the
interpretation of the sinusoidal waves the following conclusion can be drawn:

 |𝜓(𝑥, 𝑡)|2 is the probability density for finding a particle with position 𝑥.
2
 |𝜓̃(𝑝, 𝑡)| is the probability density for finding a particle with momentum 𝑝.
In addition, since 𝜓(𝑥, 𝑡) is the probability amplitude for position, its Fourier transform 𝜓̃(𝑝, 𝑡) is the
probability amplitude for momentum.

Through generalization, Born interpretation of wavefunction can be extended to describe momenta


of a particle moving in three dimensions.

A Particle in a Box

In this section position and momentum probability densities of a system are calculated considering
one of the simplest systems in quantum mechanics: a particle of mass 𝑚confined in a one
dimensional region 0 ≤ 𝑥 ≤ 𝑎. Such a particle in a state with quantum number 𝑛 has an energy
given by

ℏ2 𝑘𝑛
2 𝑛𝜋
𝐸𝑛 = 2𝑚
, with 𝑘𝑛 = 𝑎
, 3.23

and a wavefunction
−𝑖𝐸𝑛 𝑡/ℏ
𝜓(𝑥, 𝑡) = {𝑁 sin 𝑘𝑛 𝑥𝑒 0 < 𝑥 < 𝑎. 3.24
0 elsewhere

The normalization constant 𝑁 can be obtained by normalizing the position probability density. This is
given by |𝜓(𝑥, 𝑡)|2 ; this is zero outside the region 0 < 𝑥 < 𝑎, and inside this region it is given by

𝑖𝐸𝑛 𝑡⁄ 𝑖𝐸𝑛 𝑡⁄
|𝜓(𝑥, 𝑡)|2 = 𝑁 ∗ sin 𝑘𝑛 𝑥𝑒 + ℏ 𝑁 sin 𝑘𝑛 𝑥 𝑒 − ℏ = 𝑁 2 sin2 𝑘𝑛 𝑥. 3.25

The total probability of finding a particle at any of its possible locations is given by the normalization
integral

+∞ 𝑎 𝑎
∫−∞ |𝜓(𝑥, 𝑡)|2 𝑑𝑥 = |𝑁|2 ∫0 sin2 𝑘𝑥𝑑𝑥 = |𝑁|2 2 . 3.26

2
By equating this probability to one, 𝑁 = √𝑎 give rise to a normalized probability density and to a
normalized wavefunction.
2
The momentum probability density is given by |𝜓̂𝑛 (𝑝, 𝑡)| , where according to the momentum
probability amplitude

1 +∞
𝜓̂𝑛 (𝑝, 𝑡) = ∫ 𝜓 (𝑥, 𝑡) 𝑒 −𝑖𝑝𝑥/ℏ 𝑑𝑥. 3.27
√2𝜋ℏ −∞ 𝑛

−𝑖𝐸𝑛 𝑡/ℏ
Considering 𝜓(𝑥, 𝑡) = {𝑁 sin 𝑘𝑛 𝑥𝑒 0 < 𝑥 < 𝑎, with , 𝑁 = 𝑎,
√2
0 elsewhere
1 +∞
𝜓̂𝑛 (𝑝, 𝑡) = 𝑒 −𝑖𝐸𝑛 𝑡/ℏ ∫−∞ 𝑒 −𝑖𝑝𝑥/ℏ sin 𝑘𝑛 𝑥 𝑑𝑥 , 3.28
√𝜋ℏ𝑎

and the integral can be easily evaluated using

𝑒 𝑖𝑘𝑛 𝑥 −𝑒 −𝑖𝑘𝑛 𝑥
sin 𝑘𝑛 𝑥 = 2𝑖
. 3.29

In the figure below, position and momentum probabilities densities for a particle confined in a
region 0 < 𝑥 < 𝑎 are shown. Two possible states are shown, the ground state with 𝑛 = 1, and the
second excited state with 𝑛 = 3. The position probability density shown on the left of the figure
shows that the most probable position is 𝑥 = 𝑎/2 when 𝑛 = 1 and when 𝑛 = 3 there are three
most probable positions: at 𝑥 = 𝑎/6, 𝑥 = 𝑎/2, and 𝑥 = 5𝑎/6. The momentum probability densities
shown on the right of the figure indicate that the most probable momentum is zero when 𝑛 = 1 and
that there two most likely momenta at 𝑛 = 3; one at 𝑝 = +3𝜋ℏ/𝑎 and one at 𝑝 = −3𝜋ℏ/𝑎. In fact,
a state with a high value of 𝑛 can be thought as a particle trapped between 𝑥 = 0 and 𝑥 = 𝑎, with
two possible momenta 𝑝 = 𝑛𝜋ℏ/𝑎 and 𝑝 = −𝑛𝜋ℏ/𝑎.
Figure 3.1 : The position and momentum probability densities for a particle confined in a
box region; 𝟎 ≤ 𝒙 ≤ 𝒂. The particles wavefunction defined by equation 3.24 is normalised
with 𝒏 = 𝟏 and 𝒏 = 𝟑.
Particle in a Box II

Consider a particle in a one dimensional box of dimension 0 ≤ 𝑥 ≤ 𝑎 defined by a potential

0 0≤𝑥≤𝑎
𝑉(𝑥) = , 4.2.1
∞ elsewhere

This infinite potential well confines a particle in a one dimensional box of size 𝑎. Classically, the
particle is either lying down the well with zero energy or it is bouncing back and forth between the
walls of the well with energy up to infinity. In quantum mechanics more varied and quantized states
exist with wavefunction defined by the one dimensional Schrödinger equation

𝜕 ℏ2 𝑑 2
𝑖ℏ 𝜕𝑡 Ψ(𝑥, 𝑡) = [− 2𝑚 𝑑𝑥 2 + 𝑉(𝑥)] Ψ(𝑥, 𝑡). 4.2.2

If the particle has a definite energy 𝐸𝑛 then the wavefunction has the form
𝐸𝑛
Ψ(𝑥, 𝑡) = 𝜓(𝑥)𝑒 −𝑖 ℏ 𝑡 , 4.2.3

where the wavefunction 𝜓(𝑥) satisfy the energy eigenvalue equation

ℏ2 𝑑 2
𝐸𝜓(𝑥) = [− 2𝑚 𝑑𝑥 2 + 𝑉(𝑥)] 𝜓(𝑥), 4.2.4
which is often referred to as the time-independent Schrödinger equation. The task now is to find the
physically acceptable solutions to equation 4.2.4. Because the potential rises abruptly to infinity at
𝑥 = 0 and 𝑥 = 𝑎, particle is strictly confined in the region 0 ≤ 𝑥 ≤ 𝑎, beyond this region the
eigenfunction 𝜓(𝑥) is zero! Inside this region, the potential energy is zero, so the eigenfunction is a
solution of equation 4.2.4 with 𝑉(𝑥) = 0. Let us simplify this equation by writing the energy as

ℏ2 𝑘 2
𝐸= , 4.2.5
2𝑚

so that equation 4.2.4 can now appear as

𝑑2
𝑑𝑥 2
𝜓(𝑥) = −𝑘 2 𝜓(𝑥). 4.2.6

Physically acceptable solutions to this partial differential equation are obtained from the general
solution 𝜓(𝑥) = 𝑀 cos 𝑘𝑥 + 𝑁 sin 𝑘𝑥,

where 𝑀 and 𝑁 are constants. The boundary conditions

𝜓(0) = 𝜓(𝑎) = 0 4.2.7

are imposed to ensure that the position probability of the particle does not change abruptly at the
edges. It must be noted the eigenvalue problem of a particle in a one dimensional box is identical to
the eigenvalue problem of a vibrating string. In both cases there are infinite number of
eigenfunctions identified by an integer = 1, 2, 3, … .

They are given by


𝑛𝜋
𝜓𝑛 (𝑥) = 𝑁 sin 𝑘𝑛 𝑥, and 𝑘𝑛 = , where 𝑁 is an arbitrary constant, and they illustrated in Fig 4.2.2
𝑎
below.
In classical physics, eigenfunctions 𝜓𝑛 are used to describe the possible shapes of normal modes of
vibration of a string. In they can be used to describe possible shapes of wavefunctions of a particle in
a box with definite energies 𝐸𝑛 , associated with quantum numbers 𝑛 = 1, 2, 3, … .

In conclusion the possible energy levels of a particle in a one dimensional box with width 𝑎 are given
by

𝑛2 𝜋2 ℏ2
𝐸𝑛 = 2𝑚𝑎 2
, with 𝑛 = 1, 2, 3, …, 4.2.8

And that a particle with energy 𝐸𝑛 , has a wavefunction of the form


𝐸𝑛
Ψ(𝑥, 𝑡) = 𝑁 sin 𝑘𝑛 𝑥 𝑒 −𝑖 ℏ 𝑡 . 4.2.9

Note the following:


𝐸𝑛+1 −𝐸𝑛 2
 The separation between energy levels ⟶ increases with increasing quantum
𝐸𝑛 𝑛
number 𝑛.
𝜋2 ℏ2
 The minimum energy state is not zero but some minimum value 𝐸1 = 2𝑚𝑎2.
 The spacial shape of the wavefunction of a particle in box with energy 𝐸 is identical to the
spacial shape of the normal modes of a vibrating string with angular frequency 𝜔.
 The wavefunction of a particle in a box, unlike the displacement of a vibrating string is not an
observable quantity. But it can be used to construct physical observables like position and
momentum.
Square-well potential

Insights on how quantum particles can be bound or scattered by potential energy fields can be
explained by using the models of square wells and square barriers. In these models, Schrödinger
equation can be solved easily using elementary mathematics. The possible energies of a particle can
be found and the properties of the wavefunction are self-evident.

In this section quantum states of a particle in a one dimensional square-well potential are
considered. It will be shown that when the well is deep enough, bound states with discrete energy
levels are possible.

Consider a particle of mass 𝑚 in a one dimensional potential defined by

−∞ if − ∞ ≤ 𝑥 ≤ 0
𝑉(𝑥) = { −𝑉0 if 0 ≤ 𝑥 ≤ 𝑎 .
0 if 𝑎 ≤ 𝑥 ≤ +∞

As shown in Fig 5.1, the potential energy changes abruptly at 𝑥 = 0 and 𝑥 = 𝑎. There is a potential
well of depth 𝑉0 , which may or may not trap the particle, and an infinite wall at 𝑥 = 0 which repels
the particle.

Figure 5.1: A square-well potential model

The behaviour of a classical particle in this situation can be anticipated. It is known classically, the
energy of the particle is given by the sum of its kinetic energy and potential energy:
𝑝2
𝐸 = 𝐾 + 𝑉(𝑥) = 2𝑚 + 𝑉(𝑥). 5.1

The particle is said to be bound or trapped in a well of depth −𝑉0 , when its energy is negative in the
region −𝑉0 ≤ 𝐸 ≤ 0. Here, the particle bounces back and forth between 𝑥 = 0 and 𝑥 = 𝑎, with
kinetic given by 𝐾 = 𝐸 + 𝑉0 . But when the energy is positive, the particle is unbound. For example, it
can approach the well from 𝑥 = +∞ with kinetic energy 𝐾 = 𝐸, at 𝑥 = 𝑎 kinetic energy rises to
𝐾 = 𝐸 + 𝑉0 , it hits the infinitely high potential wall at 𝑥 = 0, then bounces back to 𝑥 = +∞.

In quantum mechanics, the particles motion is described by the wavefunction Ψ(𝑥, 𝑡) which is a
solution of the Schrödinger wave equation

𝜕 ℏ2 𝑑 2
𝑖ℏ 𝜕𝑡 Ψ(𝑥, 𝑡) = [− 2𝑚 𝑑𝑥 2 + 𝑉(𝑥)] Ψ(𝑥, 𝑡). 5.2

If the particle has definite energy E, the wavefunction takes the form
𝐸⁄ 𝑡
Ψ(𝑥, 𝑡) = 𝜓(𝑥)𝑒 −𝑖 ℏ , 5.3

where 𝜓(𝑥) is the eigenfunction which satisfy the energy eigenvalue equation

ℏ2 𝑑 2
− 2𝑚 𝑑𝑥 2 𝜓(𝑥) + 𝑉(𝑥)𝜓(𝑥) = 𝐸𝜓(𝑥). 5.4

After solving this energy eigenvalue equation for suitable eigenvalues and eigenfunctions, any
quantum states of the particle in a potential can be written as the linear superposition of energy
eigenfunctions.

Bound states

If a bound state exists, it has a negative energy somewhere between 𝐸 = −𝑉0 and 𝐸 = 0. Let 𝐸 =
−𝜖, where 𝜖 is the particles binding energy, and seek the solutions to equation 5.4 above.

 In the region −∞ ≤ 𝑥 ≤ 0
The potential energy is infinite, so the only possible solution to equation 5.4 is 𝜓(𝑥) = 0,
which means the particle will never be found in the negative 𝑥 direction.
 In the region 0 ≤ 𝑥 ≤ 𝑎
The potential energy is 𝑉(𝑥) = −𝑉0 and equation 5.4 has the form
𝑑2 ℏ2 𝑘02
𝜓(𝑥) = −𝑘02 𝜓(𝑥), with 𝐸 = − 𝑉0 . 5.5
𝑑𝑥 2 2𝑚
The general solution to this equation is
𝜓(𝑥) = 𝐶 sin(𝑘0 𝑥 + 𝛾),
where 𝐶 and 𝛾 are arbitrary constants. To ensure continuity of 𝜓(𝑥) at 𝑥 = 0, 𝛾 is set to
zero so that
𝜓(𝑥) = 𝐶 sin 𝑘0 𝑥. 5.6

 In the region 𝑎 ≥ 𝑥 ≥ +∞
The potential energy is zero so equation 5.4 takes the form
𝑑2 𝛼 2 ℏ2
𝑑𝑥 2
𝜓(𝑥) = 𝛼 2 𝜓(𝑥), with 𝐸 = − 2𝑚
. 5.7
The general solution to equation 5.7 is
𝜓(𝑥) = 𝐴𝑒 −𝛼𝑥 + 𝐴′𝑒 +𝛼𝑥 ,
where 𝐴 and 𝐴′ are arbitrary constants. To ensure that the eigenfunction is finite at infinity,
𝐴′ is set to zero to give a solution which falls off exponentially with 𝑥:
𝜓(𝑥) = 𝐴𝑒 −𝛼𝑥 . 5.8

The next task is to join the solution given by equation 5.6 which is valid in the region 0 ≤ 𝑥 ≤ 𝑎 and
a solution given by equation 5.8 which valid in the region 𝑎 ≥ 𝑥 ≥ +∞. To accomplish this the
eigenfunctions and their first derivatives must be continuous at 𝑥 = 𝑎. So, continuity of the
eigenfunction gives

𝐶 sin 𝑘0 𝑎 = 𝐴𝑒 −𝛼𝑎 , 5.9


𝑑
and continuity of 𝑑𝑥 𝜓 gives

𝑘0 𝐶 cos 𝑘0 𝑎 = −𝛼𝐴𝑒 −𝛼𝑎 . 5.10

If equation 5.10 is divided by equation 5.9 results in

𝑘0 cot 𝑘0 𝑎 = −𝛼. 5.11

Equation 5.11 sets a condition for a smooth join at 𝑥 = 𝑎 for the eigenfunctions 𝐶 sin 𝑘0 𝑥 and
𝐴𝑒 −𝛼𝑥 . This is a non-trivial condition which is satisfied only when 𝑘0 and 𝛼 takes on special values.
Having obtained these special values, the binding energy for the bound states can be calculated from
𝛼 2 ℏ2
the equation 𝜖 = .
2𝑚

In order to find the binding energies, it must be noted that 𝑘0 and 𝛼 are not independent
parameters. These are defined by

ℏ2 𝑘02 𝛼 2 ℏ2
𝐸= − 𝑉0 , and 𝐸 = − ,
2𝑚 2𝑚

which imply that

𝑤 2 ℏ2
𝛼 2 + 𝑘02 = 𝑤 2 , where 𝑤 is given by 𝑉0 = 2𝑚
. 5.12

This has resulted in two simultaneous equations 5.11 and 5.12. These equations can be solved
graphically by finding the intersection points of the two curves 𝑘0 cot 𝑘0 𝑎 = −𝛼 and 𝛼 2 + 𝑘02 = 𝑤 2 .
The solution is illustrated in Fig 5.2 below.
Figure 5. 1: The behaviour of simultaneous functions 𝒌𝟎 𝐜𝐨𝐭 𝒌𝟎 𝒂 = −𝜶 and 𝜶𝟐 + 𝒌𝟐𝟎 = 𝒘𝟐 to determine the special
values for calculating the binding energies.

Fig 5.2 shows the number of intersection points, and hence the number of bound states increase as
the potential well becomes deeper! Particularly there are no bound states for a shallow well with
𝜋
𝑤< ,
2𝑎

there is one bound states when


𝜋 3𝜋
2𝑎
< 𝑤 < 2𝑎,

there are two bound states when


3𝜋 5𝜋
<𝑤< ,
2𝑎 2𝑎

and so on.
2𝜋
To illustrate the nature of bound states, consider a potential with well-depth parameter 𝑤 = 𝑎
,
which corresponds to well with depth

2𝜋2 ℏ2
𝑉0 = 𝑚𝑎 2
.

In this case two bound states exist, with ground state energy and the first excited state energy

𝜋2 ℏ2 𝜋2 ℏ2
𝜖1 = 3.26 2𝑚𝑎2, and 𝜖2 = 1.17 2𝑚𝑎2 .

Potential barrier

Consider a barrier potential field in one dimension defined by

0 if − ∞ ≤ 𝑥 ≤ 0
𝑉(𝑥) = 𝑉0 if 0 ≤ 𝑥 ≤ 𝑎, 5.13
0 if 𝑎 ≤ 𝑥 ≤ +∞

which can be illustrated as in Fig 5.3 below

Figure 5.2: The potential step as defined by equation 5.13 is used to illustrate quantum mechanical tunnelling effect.

If the classical particle approaches this step from the left, it gets reflected at 𝑥 = 0 if its energy is less
than the potential step 𝑉0 , and it gets transmitted through if its energy is greater than the potential
barrier.

Quantum mechanically, if a particle approaches the barrier the result is uncertain. The particle may
be reflected or transmitted depending on the state of the incident particle. More importantly, it will
be shown that even if the energy of a particle is less than the potential step, the particle may be
transmitted through. The behaviour of a particle of mass m, in a potential barrier V(x) is described by
the Schrödinger equation

𝜕 ℏ2 𝜕2
𝑖ℏ 𝜕𝑡 Ψ(𝑥, 𝑡) = [− 2𝑚 𝜕𝑥 2 + 𝑉(𝑥)] Ψ(𝑥, 𝑡),
In order to describe an uncertain encounter with a barrier, a wavefunction Ψ(𝑥, 𝑡) which describe an
incident wave and the probability of being reflected or transmitted is required. This incoming
wavefunction must results in an incoming pulse which represent the probability of an incident wave.
After hitting the barrier, the wavefunction should give rise in two pulses, one for the probability that
the wavefunction will be reflected, the other for the probability that wavefunction will be
transmitted. According to the standard interpretation of the wavefunction, the reflection and
transmission probabilities persist at equal chances until the particle is detected. After this
occurrence, the wavefunction Ψ(𝑥, 𝑡) collapses, and one or other of the two options are realized
with probabilities governed by the magnitudes of the reflected and transmitted pulses.

This dynamical description suggests that a time-dependent quantum mechanics problem must be
solved. Even though the time-dependent quantum mechanics is required for the conceptual
understanding of uncertain encounter with the barrier potential, this is not necessary for the
calculation of the reflection and transmission probabilities. In the subsequent discussion it will be
shown how the time-independent quantum mechanics is used to calculate these probabilities.

Stationery states analysis

Provided the uncertainty of the particles energy is small compared to the potential barrier 𝑉(𝑥), the
reflection and transmission probabilities can be calculated by considering stationery states with
definite energy. To acquire this, a time-independent Schrödinger eigenvalue equation

ℏ2 𝜕2
𝐸𝜓(𝑥) = [− + 𝑉(𝑥)] 𝜓(𝑥). 5.14
2𝑚 𝜕𝑥 2

Due to the simple nature of the barrier potential of equation 5.13, it is a straight forward procedure
to find the eigenfunctions which describe the incident, reflected, and transmitted waves. The
solutions along 𝑥 at different regions are first sorted out, then joined smoothly at the boundaries
𝑥 = 0 and 𝑥 = 𝑎. On the left hand side of the barrier, the potential 𝑉(𝑥) = 0, the eigenfunction
𝜓(𝑥) satisfies the differential equation

𝑑2 ℏ2 𝑘 2
𝑑𝑥 2
𝜓(𝑥) = −𝑘 2 𝜓(𝑥), with 𝐸 = 2𝑚
. 5.15

The solution representing an incident wave of intensity |𝐴𝐼 |2 and reflected wave of intensity |𝐴𝑅 |2 is

𝜓(𝑥) = 𝐴𝐼 𝑒 𝑖𝑘𝑥 + 𝐴𝑅 𝑒 −𝑖𝑘𝑥 . 5.16

The form of the eigenfunction inside the barrier depends on whether, the energy of the particle is
greater or less than the barrier potential 𝑉0 . If 𝐸 > 𝑉0, the region (0 ≤ 𝑥 ≤ 𝑎) is classically allowed
and eigenfunctions are explained by the equation

𝑑2 ℏ2 𝑘02
𝜓(𝑥) = −𝑘02 𝜓(𝑥), with 𝐸 = + 𝑉0 . 5.17
𝑑𝑥 2 2𝑚

The general solution involves two arbitrary constants and it undulates with wave number 𝑘0

𝜓(𝑥) = 𝐴𝑒 𝑖𝑘0 𝑥 + 𝐴′𝑒 −𝑖𝑘0 𝑥 . 5.18

When 𝐸 < 𝑉0, the region (0 ≤ 𝑥 ≤ 𝑎) is classically forbidden. The eigenfunctions are described by
the equation
𝑑2 ℏ2 𝛽 2
𝑑𝑥 2
𝜓(𝑥) = 𝛽 2 𝜓(𝑥), with 𝐸 = − 2𝑚
+ 𝑉0 , 5.19

and the general solution is

𝜓(𝑥) = 𝐵𝑒 −𝛽𝑥 + 𝐵′𝑒 𝛽𝑥 . 5.20

Finally, on the right hand side of the barrier the potential energy 𝑉(𝑥) = 0. The eigenfunctions
satisfy the equation 5.15 and the solution representing the transmitted wave of intensity |𝐴 𝑇 |2 is

𝜓(𝑥) = 𝐴 𝑇 𝑒 𝑖𝑘𝑥 . 5.21

Smooth joining of these solutions at 𝑥 = 0 and 𝑥 = 𝑎 helps in evaluation of intensities of reflection


and transmission waves. Particularly, the expressions for the ratios

|𝐴𝑅 |2 |𝐴𝑇 |2
𝑅= |𝐴𝐼 |2
and 𝑇 = |𝐴𝐼 |2
, 5.22

can be derived.

Because the probability of finding a particle at 𝑥 is proportional to |𝜓(𝑥)|2 , these ratios are
probabilities: 𝑅 is the probability that the particle is reflected and 𝑇 is the probability that the
particle is transmitted, and sum of these probabilities is one, i.e.

𝑅 + 𝑇 = 1. 5.23

Tunnelling through the barrier

The eigenfunction of a particle with 𝐸 < 𝑉0 are given by

𝐴𝐼 𝑒 𝑖𝑘𝑥 + 𝐴𝑅 𝑒 −𝑖𝑘𝑥 − ∞ ≤ 𝑥 ≤ 0
−𝛽𝑥
𝜓(𝑥) = 𝐵𝑒 𝛽𝑥 + 𝐵′𝑒 0≤𝑥≤𝑎 , 5.24
𝐴 𝑇 𝑒 𝑖𝑘𝑥 𝑎 ≤ 𝑥 ≤ +∞

With constants 𝐴𝐼 , 𝐴𝑅 , 𝐴 𝑇 , 𝐵, and 𝐵′ having values ensuring 𝜓(𝑥) and its derivative with respect to
position are continuous at 𝑥 = 0 and 𝑥 = 𝑎.
𝑑
For the 𝜓(𝑥) and 𝑑𝑥 𝜓(𝑥) to be continuous at 𝑥 = 0, it require that

𝐴𝐼 + 𝐴𝑅 = 𝐵 + 𝐵′, 5.25(a)

𝑖𝑘𝐴𝐼 − 𝑖𝑘𝐴𝑅 = 𝛽𝐵 − 𝛽𝐵′, 5.25(b)

and at 𝑥 = 𝑎
−𝛽𝑥
𝐵𝑒 𝛽𝑥 + 𝐵′𝑒 = 𝐴 𝑇 𝑒 𝑖𝑘𝑥 , 5.26(a)
−𝛽𝑥
𝛽𝐵𝑒 𝛽𝑥 − 𝛽𝐵′𝑒 = 𝑖𝑘𝐴 𝑇 𝑒 𝑖𝑘𝑥 . 5.26(b)
The main aim is to find the expression for the tunnelling probability. To acquire this, 𝐴𝐼 and 𝐴 𝑇 are
written in terms of other constant so that equations at 𝑥 = 0 give

2𝑖𝑘𝐴𝐼 = −(𝛽 − 𝑖𝑘)𝐵 + (𝛽 + 𝑖𝑘)𝐵′, 5.27

and at 𝑥 = 𝑎 give

2𝛽 (𝛽+𝑖𝑘)
𝐴 𝑇 𝑒 𝑖𝑘𝑥 = (𝛽−𝑖𝑘) 𝐵𝑒 −𝛽𝑎 and 𝐵′ = (𝛽−𝑖𝑘) 𝐵𝑒 −2𝛽𝑎 . 5.28

The algebra is further simplified by assuming that 𝑒 −2𝛽𝑎 ≪ 1, which consequently imply that 𝐵′ ≪
𝐵 so that to a good approximation

2𝑖𝑘𝐴𝐼 ≈ −(𝛽 − 𝑖𝑘)𝐵. 5.29

Combining equation 5.29 and 5.28, results in

4𝑖𝑘𝛽
𝐴 𝑇 𝑒 𝑖𝑘𝑥 ≈ (𝛽−𝑖𝑘)2 𝐴𝐼 𝑒 −𝛽𝑎 . 5.30

This implies that the transmission probability is approximated to

16𝑘 2 𝛽 2 −2𝛽𝑎
𝑇 ≈ (𝛽2 𝑒 . 5.31
+𝑘 2 )2

The Harmonic Oscillator

The harmonic oscillator played a crucial role in the development of quantum mechanics. In 1900,
Planck made a bold assumption that atoms are behaving like harmonic oscillators when they absorb
and emit radiation. This assumption was complemented by Einstein in 1905, when he also assumed
that electromagnetic radiation acted like electromagnetic harmonic oscillator with quantized energy;
further in 1907 he assumed that elastic vibrations in solids behaved like a system of mechanical
oscillators with quantized energy. These assumptions later proved to support the black body
radiation, photoelectric effect and temperature dependence of the specific heat of solids. In so
doing, quantum theory provided a good explanation of the electromagnetic waves and mechanical
harmonic oscillators.

In this section, quantum mechanical behaviour of particle in the harmonic oscillator potential will be
explored. In particular, energy eigenvalues and eigenfunctions of stationery and non-stationery
quantum states. These states are quite useful in molecular, solid state, and nuclear physics, but
more generally in quantum field theory.

The Classical Oscillator


A simple example of a harmonic oscillator is a mass-spring system with the spring constant 𝑘. When
a mass 𝑚 is displaced from equilibrium position by a distance 𝑥, there is a restoring force 𝐹 = −𝑘𝑥
which oppose this displacement. Because the work needed to move the mass from 𝑥 to 𝑥 + 𝑑𝑥 is
𝑘𝑥𝑑𝑥, the potential energy stored by displacing a mass 𝑚 is
𝑥 1
𝑉(𝑥) = ∫0 𝑘𝑥′𝑑𝑥′ = 𝑘𝑥 2 . 6.1
2

This potential energy is converted to kinetic energy when a mass is released. The equation of motion
for the mass 𝑚 is

𝑑2 𝑥
𝑚 𝑑𝑡 2 = −𝑘𝑥 6.2

which is usually written as

𝑥̈ = −𝜔2 𝑥, 6.3

where 𝑥̈ is the acceleration of mass 𝑚 and 𝜔 = √𝑘/𝑚. The motions general solution is

𝑥 = 𝐴 cos(𝜔𝑡 + 𝛿), 6.4

where 𝐴 and 𝛿 are two constants which may be determined by specifying the initial position and
velocity of mass 𝑚; for example if the mass is released from rest at position 𝑥0 , then 𝐴 = 𝑥0 and
𝛿 = 0.

Equation 6.4 describes a simple harmonic motion along the x-direction with amplitude 𝐴, phase 𝛿
and frequency 𝜔 or period 2𝜋/𝜔. In the cause of the motion, the potential energy rises and falls as
the kinetic energy falls and rises. But the total energy 𝐸, of which is the sum of the kinetic and
potential energy remains constant and equal to
1 1 1
𝐸 = 2 𝑚𝑥̇ 2 + 2 𝑘𝑥 2 = 2 𝑚𝜔2 𝐴2. 6.5

In the real quantum situations, simple harmonic motion with definite energy, phase, frequency, and
amplitude never happens. It is going to be shown that oscillators either have definite energy and do
not oscillate or they oscillate with uncertain energy. But in certain circumstances, oscillations are
more or less like the simple harmonic motion.

The Quantum Oscillator

In a quantum system, the Hamiltonian operator is the defining property. In one dimension harmonic
oscillator this Hamiltonian operator is
2
̂ = 𝑝̂ + 1 𝑚𝜔2 𝑥̂ 2 ,
𝐻 6.6
2𝑚 2

or if position and momentum operators are used


2 2
̂ = − ℏ 𝜕 2 + 1 𝑚𝜔2 𝑥 2.
𝐻 6.7
2𝑚 𝜕𝑥 2

The first term represent the kinetic energy of a particle with mass 𝑚, and the second term represent
the potential energy of the particle in a potential well, of which classically will give rise to simple
harmonic motion with angular frequency 𝜔. The particles behaviour in an harmonic oscillator
potential is more varied in quantum mechanics than in classical physics. There are infinite number of
quantum states; some are stationery states with definite energy while some are non-stationery
states with uncertain energy. All of these states are described by the wavefunction 𝜓(𝑥, 𝑡) which
satisfy the Schrӧdinger equation

𝜕
̂ 𝜓(𝑥, 𝑡),
𝑖ℏ 𝜕𝑡 𝜓(𝑥, 𝑡) = 𝐻 6.8

where 𝐻̂ is the Hamiltonian as defined above. The first focus will be on quantum states with definite
energy. From previous discussion, states with energy 𝐸 are defined by

Ψ(𝑥, 𝑡) = 𝜓(𝑥)𝑒 −𝑖𝐸𝑡/ℏ , 6.9

where 𝜓(𝑥) is an eigenfunction belong to an energy eigenvalue 𝐸. If this wavefunction is used


together with the Hamiltonian operator defined above, an energy eigenvalue equation becomes

ℏ2 𝑑 2 1
[− 2𝑚 𝑑𝑥2 + 2 𝑚𝜔2 𝑥 2 ] 𝜓(𝑥) = 𝐸𝜓(𝑥). 6.10

In order to obtain solutions to this equation, a physical condition that a wavefunction must be
normalizable is introduced. This is achieved by letting the eigenfunctions to go to zero at infinity; i.e.

𝜓(𝑥) ⟶ 0 as 𝑥 ⟶ ±∞. 6.11

Because the sides of the harmonic oscillator potential are infinitely high like the infinite square-well
potential, an infinite number of quantized eigenvalues is expected. These are denoted by 𝐸𝑛 , and
the corresponding eigenfunctions by 𝜓𝑛 (𝑥), where 𝑛 is a quantum number. The convention of
labelling the ground state by 𝑛 = 0, and the subsequent first, second, and third excited states by
𝑛 = 1, 2, 3, … is followed.

Diatomic Molecules

To a first approximation a diatomic molecule consist of two nuclei held together by an effective
potential which arises from the Coulomb interaction of electrons and nuclei and quantum behaviour
of electrons. This effective potential determines the strength of the molecular bond between nuclei
and it also governs the vibrational motion of nuclei. The effective potential and the vibrational
energy levels of the simplest diatomic molecule, the hydrogen molecule is shown below:
It must be noted that the effective potential of a diatomic molecule has a minimum, and that near
the minimum the shape of the potential is like a harmonic oscillator potential. Indeed if 𝑟0 denotes
the separation at which the potential has a minimum, and 𝑥 = 𝑟 − 𝑟0 denotes a small displacement
from 𝑟0 then
1
𝑉𝑒 (𝑟) ≈ 2 𝑘𝑥 2 , 6.12

where 𝑘 is a constant. This equation implies that if nuclei are displaced a distance 𝑥 from their
equilibrium separation 𝑟0 , there is a restoring force of magnitude 𝑘𝑥, and the potential energy
1
increases by 𝑘𝑥 2 . The constant 𝑘 is the effective elastic constant which characterises the strength
2
of molecular bond between nuclei in the molecule. If classical physics was applicable, the nuclei
would have energy

𝑝2 𝑝2 1
𝐸classical = 2𝑚1 + 2𝑚2 + 2 𝑘𝑥 2 , 6.13
1 2

where 𝑚1 and 𝑚2 are masses of the nuclei, and 𝑝1 and 𝑝2 are the magnitudes of their momenta. In
the centre of mass frame, 𝑝1 = 𝑝2 = 𝑝, and the reduced mass is
𝑚 𝑚
𝜇 = 𝑚 1+𝑚2 , 6.14
1 2

so that the classical total energy becomes

𝑝2 1
𝐸classical = 2𝜇 + 2 𝑘𝑥 2 . 6.15

This is the same as the energy of a single particle of mass 𝜇 on a spring with elastic constant 𝑘.
Accordingly, the vibrating nuclei in a diatomic molecule act like a harmonic oscillator with classical
frequency 𝜔 = √𝑘/𝜇, where 𝜇 is the reduced mass of the nuclei and 𝑘 is the elastic constant
characterizing the strength of the molecular bond between the nuclei.

The quantum mechanical behaviour of this oscillator is described by a wave function 𝜓(𝑥, 𝑡) which
satisfies the Schrӧdinger equation
𝜕 ℏ2 𝜕2 1
𝑖ℏ 𝜕𝑡 𝜓 = [− 2𝜇 𝜕𝑥 2 + 2 𝑘𝑥 2 ] 𝜓. 6.16

It can also be shown that the energy eigenvalues of the harmonic oscillator with a classical angular
frequency 𝜔 is given by

1
𝐸𝑛 = (𝑛 + 2) ℏ𝜔 with 𝑛 = 0, 1, 2, 3, … 6.17

As shown below, the energy levels have equal spacing ℏ𝜔 and lowest energy level is
1
𝐸0 = 2 ℏ𝜔. 6.18

When 𝑛 is large the harmonic oscillator model of molecular vibrations breaks down. This occurs
when the vibrational energy becomes comparable with the dissociation energy of the molecule. A
transition from one vibrational level to another is accompanied by emission or absorption of
electromagnetic radiation. In diatomic molecules, electrons form an electric dipole which can
strongly absorb or emit electromagnetic radiation. This mechanism leads to the emission and
𝑘
absorption of photons with energy 𝐸 = ℏ√𝜇. These photons give rise to spectral lines having

ℎ𝑐 𝜇
wavelength 𝜆 = 𝐸
= 2𝜋𝑐√𝑘 .

Example:
Carbon monoxide (CO) has reduced mass 𝜇 = 6.85 u
Transitions occur due to infrared radiation of 𝜆 = 4.6 μm

𝜇
Substituting these values in 𝜆 = 2𝜋𝑐√𝑘 we obtain

The strength of the bond as the elastic constant 𝑘 = 1908 Nm-1.

3-D Oscillators

Consider a particle of mass 𝑚 in a three dimensional oscillator potential


1 1
𝑉(𝑟) = 2 𝑘𝑟 2 = 2 𝑘(𝑥 2 + 𝑦 2 + 𝑧 2 ). 6.19

A classical particle at a distance 𝑟 from the origin will experience a central force of magnitude 𝑘𝑟.
When displaced from the origin and released it executes a simple harmonic motion with angular
frequency 𝜔 = √𝑘/𝑚. A more complicated motion occurs if a particle is displaced and also given
transverse velocity.

The behaviour of a quantum particle is governed by the Hamiltonian operator of which in three
dimensions is given by

̂=𝐻
𝐻 ̂𝑥 + 𝐻
̂𝑦 + 𝐻
̂𝑧 , where

ℏ2 𝜕2 1
̂𝑥 = −
𝐻 + 𝑚𝜔2 𝑥 2 , 6.20(a)
2𝑚 𝜕𝑥 2 2

2 2
̂𝑦 = − ℏ 𝜕 2 + 1 𝑚𝜔2 𝑦 2 ,
𝐻 6.20(b)
2𝑚 𝜕𝑦 2

ℏ2 𝜕2 1
̂𝑧 = −
𝐻 + 𝑚𝜔2 𝑧 2. 6.20(c)
2𝑚 𝜕𝑧 2 2

Stationery states with definite energy are given by wave functions of the form

Ψ(𝑥, 𝑦, 𝑧, 𝑡) = 𝜓(𝑥, 𝑦, 𝑧)𝑒 −𝑖𝐸𝑡/ℏ, 6.21

where 𝜓(𝑥, 𝑦, 𝑧) and 𝐸 satisfy the three dimensional eigenvalue equation

̂ 𝜓(𝑥, 𝑦, 𝑧) = 𝐸𝜓(𝑥, 𝑦, 𝑧).


𝐻 6.22

̂𝑥 , 𝐻
These states can be found by using the one dimensional eigenvalue equations according to 𝐻 ̂𝑦 ,
̂𝑧 as
and 𝐻

̂𝑥 𝜓𝑛 (𝑥) = (𝑛𝑥 + 1)ℏ𝜔𝜓𝑛 (𝑥),


𝐻 6.23(a)
𝑥 2 𝑥

̂𝑦 𝜓𝑛 (𝑦) = (𝑛𝑦 + 1)ℏ𝜔𝜓𝑛 (𝑦),


𝐻 6.23(b)
𝑦 2 𝑦

̂𝑧 𝜓𝑛 (𝑧) = (𝑛𝑧 + 1)ℏ𝜔𝜓𝑛 (𝑧),


𝐻 6.23(c)
𝑧 2 𝑧

where the quantum numbers 𝑛𝑥 , 𝑛𝑦 , 𝑛𝑧 can take values 𝑛 = 0, 1, 2, …

These three equations imply that the wave function 𝜓𝑛𝑥 ,𝑛𝑦 ,𝑛𝑧 (𝑥, 𝑦, 𝑧) = 𝜓𝑛𝑥 (𝑥)𝜓𝑛𝑦 (𝑦)𝜓𝑛𝑧 (𝑧)
satisfy the three dimensional eigenvalue equation

̂ 𝜓𝑛 ,𝑛 ,𝑛 (𝑥, 𝑦, 𝑧) = 𝐸𝑛 ,𝑛 ,𝑛 𝜓𝑛 ,𝑛 ,𝑛 (𝑥, 𝑦, 𝑧)
𝐻 6.24
𝑥 𝑦 𝑧 𝑥 𝑦 𝑧 𝑥 𝑦 𝑧

provided that
3
𝐸𝑛𝑥 ,𝑛𝑦 ,𝑛𝑧 = (𝑛𝑥 + 𝑛𝑦 + 𝑛𝑧 + 2)ℏ𝜔. 6.25

Thus the eigenvalues and eigenfunctions of a three dimensional oscillator are labelled by three
quantum numbers 𝑛𝑥 , 𝑛𝑦 , and 𝑛𝑧 each of which can take any integer value from zero to infinity.
The form of the low-lying eigenfunctions can be found using the Table below:

Table 6.1: Normalized eigenstates for the first four lowest states of the one dimensional
harmonic oscillator.

When all three quantum numbers are equal to zero we have the ground state:

3 1 3/2 2 +𝑦 2 +𝑧 2 )/2𝑎 2
𝐸0,0,0 = 2 ℏ𝜔 and 𝜓0,0,0 (𝑥, 𝑦, 𝑧) = (𝑎 𝜋) 𝑒 −(𝑥 , 6.26

where 𝑎 = √ℏ/𝑚𝜔. By one of the quantum numbers from 0 to 1, three excited states with the
same energy are obtained:

5 1 3/2 𝑥 2 +𝑦 2 +𝑧 2 )/2𝑎 2
𝐸1,0,0 = 2 ℏ𝜔 and 𝜓1,0,0 (𝑥, 𝑦, 𝑧) = (𝑎 𝜋) 21/2 (𝑎) 𝑒 −(𝑥 ; 6.27(a)

5 1 3/2 𝑦 2 +𝑦 2 +𝑧 2 )/2𝑎2
𝐸0,1,0 = 2 ℏ𝜔 and 𝜓0,1,0 (𝑥, 𝑦, 𝑧) = (𝑎 𝜋) 21/2 (𝑎) 𝑒 −(𝑥 ; 6.27(b)

5 1 3/2 𝑧 2 +𝑦 2 +𝑧 2 )/2𝑎2
𝐸0,0,1 = 2 ℏ𝜔 and 𝜓0,0,1 (𝑥, 𝑦, 𝑧) = (𝑎 𝜋) 21/2 (𝑎) 𝑒 −(𝑥 . 6.27(c)

Angular Momentum

Let us consider a particle at time 𝑡 with position and momentum vector

𝑟⃗ = (𝑥, 𝑦, 𝑧) and 𝑝⃗ = (𝑝𝑥 , 𝑝𝑦 , 𝑝𝑧 ). 7.1

The angular momentum about the origin of coordinates system is given by

𝑙⃗ = 𝑟⃗ × 𝑝⃗. 7.2

This vector has the following three Cartesian components

𝑙𝑥 = 𝑦𝑝𝑧 − 𝑧𝑝𝑦 , 𝑙𝑦 = 𝑧𝑝𝑥 − 𝑥𝑝𝑧 , 𝑙𝑧 = 𝑥𝑝𝑦 − 𝑦𝑝𝑥 7.3

with the magnitude

|𝑙⃗| = √𝑙𝑥2 + 𝑙𝑦2 + 𝑙𝑧2 . 7.4

Quantum mechanically the angular momentum is described by the operator

𝑙̂ = 𝑟̂ × 𝑝̂ = −𝑖ℏ𝑟⃗ × ∇
⃗⃗ 7.5

this is a vector operator with three Cartesian components

𝜕 𝜕 𝜕 𝜕 𝜕 𝜕
𝑙̂𝑥 = −𝑖ℏ (𝑦 𝜕𝑧 − 𝑧 𝜕𝑦), 𝑙̂𝑦 = −𝑖ℏ (𝑧 𝜕𝑥 − 𝑥 𝜕𝑧), 𝑙̂𝑧 = −𝑖ℏ (𝑥 𝜕𝑦 − 𝑦 𝜕𝑥) 7.6

that act on wave functions representing possible quantum states of a particle.

The spin angular momentum is usually described by an operator 𝑠̂ = (𝑠̂𝑥 , 𝑠̂𝑦 , 𝑠̂𝑧 ) which acts on a
quantum state which includes a description of the spin properties of the particle. Spin operators and
spin quantum states are usually represented by matrices.
Angular momenta in classical situations.

“spin” angular momentum s is actually orbital angular momentum of the object at every point.
Extrinsic orbital angular momentum 𝐿 about an axis.

The moment of inertia tensor 𝑰 and angular velocity 𝜔 (𝐿 is not always parallel to 𝜔).

The total angular momentum (spin plus orbital) is 𝑱. For a quantum particle the interpretations are a
little different; particle spin does not have the above interpretation.

Quantization

Quantum mechanically, angular momentum is quantized, that is it cannot vary continuously, but
only in quantum leaps between certain allowed values. For any system, the following restrictions on
measuring angular momentum applies: here ℏ is the reduced Planck’s constant on either 𝑥̂, 𝑦̂, or 𝑧̂
direction:

if you measure … you get …


𝑳̂𝒙,𝒚,𝒛 … , −2ℏ, −ℏ, 0, ℏ, 2ℏ, …
𝒔̂𝒙,𝒚,𝒛or 𝑱̂𝒙,𝒚,𝒛 3 1 1 3
… , − ℏ, −ℏ, − ℏ, 0, ℏ, ℏ, ℏ, …
2 2 2 2
𝑳̂𝟐 (ℏ2 𝑛(𝑛 + 1)), 𝑛 = 0, 1, 2, …
1 3
𝒔̂𝟐 or 𝑱̂𝟐 (ℏ2 𝑛(𝑛 + 1)), 𝑛 = 0, , 1, , …
2 2

The reduced Planck’s constant ℏ is very small in everyday standards, about 10-34 Js, and therefore
this quantization does not noticeably affect the angular momentum of macroscopic objects. But this
is very crucial in the atomic scale. For example the structure of electron energy levels and sub-
energy levels is significantly affected by the quantization of angular momentum.

Commutators
The role of compatible and non-compatible observables can be made clearer by introducing a
mathematical concept of a commutator of two operators.

The commutator of two operators is defined by

[𝐴̂, 𝐵̂] ≡ 𝐴̂𝐵̂ − 𝐵̂𝐴̂. 7.7

In quantum mechanics operators do not commute i.e. 𝐴̂𝐵̂ ≠ 𝐵̂𝐴̂.

Two observables 𝐴 and 𝐵 described by operators 𝐴̂ and 𝐵̂ are non-compatible if [𝐴̂, 𝐵̂] ≠ 0 and they
are compatible if [𝐴̂, 𝐵̂] = 0 .

This statement is best understood by considering the quantum states of a particle in one dimension
and in three dimensions.

Some important examples about commutators:

ℏ 𝜕 ℏ 𝜕
𝑥̂𝑝̂ 𝜓 = 𝑥 𝑖 𝜕𝑥 𝜓 = 𝑖 𝑥 𝜕𝑥 𝜓

ℏ 𝜕 ℏ ℏ 𝜕
𝑝̂ 𝑥̂𝜓 = 𝑝̂ (𝑥̂𝜓) = 𝑖 𝜕𝑥
(𝑥𝜓) = 𝑖 𝜓 + 𝑖 𝑥 𝜕𝑥 𝜓

(𝑥̂𝑝̂ − 𝑝̂ 𝑥̂)𝜓 = 𝑖ℏ𝜓

therefore [𝑥̂, 𝑝̂ ] = (𝑥̂𝑝̂ − 𝑝̂ 𝑥̂)𝜓 = 𝑖ℏ𝜓

in a similar manner it can be shown that:

[𝑥̂, 𝑝̂𝑥 ] = [𝑦̂, 𝑝̂𝑦 ] = [𝑧̂ , 𝑝̂𝑧 ] = 𝑖ℏ 7.8(a)

[𝑧̂ , 𝑝̂𝑦 ] = [𝑥̂, 𝑝̂𝑧 ] = [𝑦̂, 𝑝̂𝑥 ] = 0 7.8(b)

Properties of commutators

[𝐴̂, 𝐵̂] = −[𝐵̂, 𝐴̂] 7.9(a)

[𝐴̂, 𝐵̂ + 𝐶̂ ] = [𝐴̂, 𝐵̂ ] + [𝐴̂, 𝐶̂ ] 7.9(b)

[𝐴̂, 𝐵̂𝐶̂ ] = [𝐴̂, 𝐵̂]𝐶̂ + 𝐵̂[𝐴̂, 𝐶̂ ] 7.9(c)

[𝐴̂, [𝐵̂, 𝐶̂ ]] = [𝐵̂, [𝐶̂ , 𝐴̂]] + [𝐶̂ , [𝐴̂, 𝐵̂]] = 0 ……. Jacobi identity! 7.9(d)

and note: [𝐴̂, 𝐴̂] = 0 and [𝐴̂, 𝑓(𝐴̂)] = 0

likewise for angular momentum we have:

[𝑙̂𝑥 , 𝑙̂𝑦 ] = 𝑖ℏ𝑙̂𝑧 ; [𝑙̂𝑦 , 𝑙̂𝑧 ] = 𝑖ℏ𝑙̂𝑥 ; [𝑙̂𝑧 , 𝑙̂𝑥 ] = 𝑖ℏ𝑙̂𝑦 7.10
[𝑙̂2 , 𝑙̂𝑥 ] = [𝑙̂2 , 𝑙̂𝑦 ] = [𝑙̂2 , 𝑙̂𝑧 ] = 0 7.11

𝑳̂𝟐 Operator

A new operator 𝐿̂2 is introduced as this operator commutes with individual component of 𝐿, but the
same components of 𝐿 do not commute with each other.

𝐿̂2 is given by 𝐿̂2 = 𝐿̂2𝑥 + 𝐿̂2𝑦 + 𝐿̂2𝑧 .

When a measurement is made we can find the total angular momentum and only one other
component at a time.

For example, if a wave function is an eigenfunction of 𝐿̂𝑧 , then it is not a function of 𝐿̂𝑥 and 𝐿̂𝑦 .

Taking measurements of angular momentum along 𝐿̂𝑧 (applying external field) shows the total
angular momentum direction in figure below

When a particle is under the influence of a central symmetric potential, then 𝐿 commutes with
potential energy 𝑉(𝑟). If 𝐿 commutes with kinetic energy, then 𝐿 is a constant of motion.

̂ ) (kinetic and potential energy operator), then the


If 𝐿 commutes with the Hamiltonian operator (𝐻
angular momentum and the energy can be known at the same time.
Eigenvalues of Angular Momentum

??? …

The Hydrogen Atom

The hydrogen atom provides the first meaningful laws of quantum mechanics. The hydrogen is the
simplest atom with one electron of charge – 𝑒 and one proton in a nucleus of charge +𝑒. To a first
approximation a nucleus with mass far larger than the electron mass can regarded as a fixed object.
This means it would be possible to understand the properties of hydrogen by just solving the one
particle quantum mechanical problem. That is the problem of an electron in Coulomb potential
energy field

𝑒2
𝑉(𝑟) = . 2.17
4𝜋𝜖0 𝑟

In this section we shall try to find the energy eigenvalues and eigenfunction of such an electron and
use them to describe outstanding features of the hydrogen atom.

Central Potentials

Let’s begin by considering a particle in a central potential like a Coulomb potential, which only
depends on the distance of the particles from a fixed origin.

Consider a classical particle of mass 𝑚, with vector position 𝑟⃗, momentum 𝑝⃗, and angular
momentum 𝐿⃗⃗ = 𝑟⃗ × 𝑝⃗, about a fixed origin. If the particle moves in a central potential 𝑉(𝑟), it is
subject to a force
𝑑𝑉
𝐹⃗ = − 𝑑𝑟 𝑒̂𝑟 , 8.1

where 𝑒̂𝑟 is the unit vector in the direction of 𝑟⃗. Because the force acts along the radius vector 𝑟⃗, the
⃗⃗ = 𝑟⃗ × 𝐹⃗ on the particle is zero, and particle moves with constant angular momentum 𝐿⃗⃗.
torque 𝑁
The geometrical implications of a constant angular momentum can be understood with the help of
the Figure 8.1 below,
Figure 8.1: Illustration of the relation between the constant angular momentum and the
area element 𝒅𝑨 ⃗⃗ swept by 𝒓⃗⃗ and position element 𝒅𝒓
⃗⃗.
which shows that the vector area swept by the radius vector 𝑟⃗ is given by

⃗⃗
𝐿
𝑑𝐴⃗ = 2𝑚 𝑑𝑡. 8.2

This implies that, when angular momentum is a constant vector 𝐿⃗⃗, the particle moves in a fixed
plane with a radius vector which sweeps out an area at a constant rate of 𝐿/2𝑚. The momentum 𝑝⃗
of the particle moving in a plane has two independent components which can be conveniently called
the radial and transverse momenta
𝑑𝑟 𝐿
𝑝𝑟 = 𝑚 𝑑𝑡 , and 𝑝𝑡 = 𝑟. 8.3

𝑝2
If we write the kinetic energy in terms of 𝑝𝑟 and 𝑝𝑡 the the constant total energy of the particle
2𝑚
is given by

𝑝2
𝑟 𝐿2
𝐸 = 2𝑚 + 2𝑚𝑟2 + 𝑉(𝑟). 8.4

𝑟 𝑝2
The total energy of the particle can also be viewed as the radial kinetic term 2𝑚 , and the effective
potential

𝐿2
𝑉𝑒𝑓𝑓 (𝑟) = 2𝑚𝑟2 + 𝑉(𝑟). 8.5

The effective force corresponding to this effective potential acts in the radial direction with
magnitude

𝑑𝑉𝑒𝑓𝑓 𝐿2 𝑑𝑉(𝑟)
𝐹=− 𝑑𝑟
= 𝑚𝑟3 − 𝑑𝑟
. 8.6
𝐿2 2
The term 𝑚𝑟 3
, which is the same as the 𝑚𝜐 ⁄𝑟 for a particle moving with speed 𝜐 in a circle of radius
𝐿2
𝑟 with angular momentum 𝐿 = 𝑚𝑟𝜐, is the centrifugal force. So that the term 2𝑚𝑟2 in the effective
potential can be thought of as the centrifugal potential energy or the transverse kinetic energy.

The most important example of a classical motion in a central potential is the planetary motion. A
planet of mass 𝑚 moves around the sun with gravitational potential energy
𝑚𝑀⨀
𝑉(𝑟) = −𝐺 𝑟
, 8.7

where 𝑀⨀ is the mass of the sun, and 𝐺 is the fundamental constant of gravity. Information on
classical mechanics shows that planets revolve around the sun in elliptical orbits. If these planets
could discharge some excess energy, their orbits would be circular. If it acquires extra energy such
that it escapes the sun, the orbit would be parabolic, further acquired energy would result in a
hyperbolic orbit.

Quantum mechanics of a particle in a central potential

Quantum states of a particle in central potential are represented in spherical coordinates by the
wave function 𝜓(𝑟, 𝜃, 𝜙, 𝑡). The focus will be on quantum states with quite definite energy 𝐸, whose
wave functions can be described by

Ψ(𝑟, 𝜃, 𝜙, 𝑡) = 𝜓(𝑟, 𝜃, 𝜙)𝑒 −𝑖𝐸𝑡/ℏ , 8.8

where 𝜓(𝑟, 𝜃, 𝜙) is the energy eigenfunction which satisfy the eigenvalue equation

ℏ2
[− 2𝑚 ∇2 + 𝑉(𝑟)] 𝜓 = 𝐸𝜓. 8.9

This now is partial differential equation in three independent variables 𝑟, 𝜃, and 𝜙. It can be
simplified if in addition to the definite energy 𝐸, definite angular momentum 𝐿 is considered.
Assume that the magnitude of the orbital angular momentum is 𝐿 = √𝑙(𝑙 + 1)ℏ, and the 𝑧-
component of this orbital angular momentum is 𝐿𝑧 = 𝑚𝑙 ℏ, where 𝑙 and 𝑚𝑙 are quantum numbers
which could take on the values 𝑙 = 0, 1, 2, … and 𝑚𝑙 = −𝑙, … , +𝑙. The eigenfunctions have the form

𝜓(𝑟, 𝜃, 𝜙) = 𝑅(𝑟)𝑌𝑙,𝑚𝑙 (𝜃, 𝜙). 8.10

𝑅(𝑟) is the radial part of and 𝑌𝑙,𝑚𝑙 (𝜃, 𝜙) are the spherical harmonics wavefunction part of the
wavefunction 𝜓(𝑟, 𝜃, 𝜙).

In this equation 𝑌𝑙,𝑚𝑙 (𝜃, 𝜙) is a simultaneous eigenfunction of 𝐿̂2 and 𝐿̂𝑧 whose values satisfy
equations

𝐿̂2 𝑌𝑙,𝑚𝑙 (𝜃, 𝜙) = 𝑙(𝑙 + 1)ℏ2 𝑌𝑙,𝑚𝑙 (𝜃, 𝜙) 8.11

and

𝐿̂𝑧 𝑌𝑙,𝑚𝑙 (𝜃, 𝜙) = 𝑚𝑙 ℏ𝑌𝑙,𝑚𝑙 (𝜃, 𝜙). 8.12


𝑅(𝑟) is an unknown function of 𝑟. If we substitute 𝜓(𝑟, 𝜃, 𝜙) into the eigenvalue equation we use
identities

1 𝜕2 (𝑟𝜓) 1 𝜕2 𝜓 cos 𝜃 𝜕𝜓 1 𝜕2 𝜓
∇2 𝜓 = 𝑟 2 𝜕𝑟 2
+ 𝑟2 [ 𝜕𝜃2 + sin 𝜃 𝜕𝜃 + sin2 𝜃 𝜕𝜙2 ] 8.13

and

𝜕 2cos 𝜃 𝜕 1 𝜕 2
𝐿̂2 = −ℏ2 (𝜕𝜃2 + sin 𝜃 𝜕𝜃 + sin2 𝜃 𝜕𝜙2 ), 8.14

together with the angular momentum identities in order to obtain the ordinary differential equation
for 𝑅(𝑟):

ℏ2 𝑑 2 (𝑟𝑅(𝑟)) 𝑙(𝑙+1)ℏ2
− 2𝑚𝑟 𝑑𝑟 2
+[ 2𝑚𝑟 2
+ 𝑉(𝑟)] 𝑅(𝑟) = 𝐸𝑅(𝑟). 8.15

By introducing a radial function defined by

𝑢(𝑟)
𝑅(𝑟) = 𝑟
, 8.16

we obtain

ℏ2 𝑑 2 𝑢(𝑟) 𝑙(𝑙+1)ℏ2
− 2𝑚 𝑑𝑟 2
+[ 2𝑚𝑟 2
+ 𝑉(𝑟)] 𝑢(𝑟) = 𝐸𝑢(𝑟). 8.17

This important equation is called the radial Schrӧdinger equation. It describes a particle with
angular momentum 𝐿 = √𝑙(𝑙 + 1)ℏ, in a 1-D effective potential of the form

𝑙(𝑙+1)ℏ2
𝑉𝑒𝑓𝑓 (𝑟) = 2𝑚𝑟 2
+ 𝑉(𝑟). 8.18

𝑙(𝑙+1)ℏ2
Comparing this effective potential with the classical one, it can be deduced that the term 2𝑚𝑟 2
can
be described as the transverse kinetic energy or as the centrifugal potential energy . When seeking
solutions of the radial Schrӧdinger equation, the boundary conditions

𝑢(𝑟) = 0 at 𝑟 = 0 8.19

𝑢(𝑟)⁄
must be imposed to ensure that the function 𝑅(𝑟) = 𝑟. Actually, this is in line with fact that
the 3-D eigenfunction 𝜓(𝑟, 𝜃, 𝜙) is finite at the origin. In addition, bound states which ensures that
the particle does not escape to infinity, must satisfy the boundary condition

𝑢(𝑟) ⟶ 0 as 𝑟 ⟶ ∞. 8.20

Bound states only exist if the effective potential 𝑉𝑒𝑓𝑓 (𝑟) is sufficiently attractive. These states are
labelled by the quantum number 𝑛𝑟 = 0, 1, 2, … which could be shown to be equal to number of
nodes of the eigenfunction 𝑢(𝑟) between 𝑟 = 0 and 𝑟 = ∞ . This means that a bound state of a
particle in a central potential can always be specified by three quantum numbers 𝑛𝑟 , 𝑙, and 𝑚𝑙 and
that eigenfunction has the form
𝑢𝑛𝑟 ,𝑙 (𝑟)
𝜓𝑛𝑟 ,𝑙,𝑚𝑙 (𝑟, 𝜃, 𝜙) = 𝑟
𝑌𝑙,𝑚𝑙 (𝜃, 𝜙). 8.21
2
By using the normalization condition ∫|𝑌𝑙,𝑚𝑙 | 𝑑Ω = 1, with 𝑑Ω = sin 𝜃 𝑑𝜃 𝑑𝜙 being the solid angle
of spherical harmonics it can be shown that the eigenfunction 𝜓𝑛𝑟,𝑙,𝑚𝑙 (𝑟, 𝜃, 𝜙) is normalized
provided the radial function 𝑢(𝑟) satisfy the condition

∞ 2
∫0 |𝑢𝑛𝑟,𝑙 (𝑟)| 𝑑𝑟 = 1. 8.22

The energy of these bound states will be denoted by 𝐸𝑛𝑟,𝑙 . By considering this energy being
constituted by the average radial kinetic energy, the average transverse kinetic energy, and the
average Coulomb energy, it can be seen that the states with higher 𝑛𝑟 and 𝑙 have higher energies.
The energy 𝐸𝑛𝑟 ,𝑙 increases with increasing 𝑙 because the average transverse kinetic energy is given
by

∞ 𝑙(𝑙+1)ℏ2
∫0 𝑢𝑛∗ 𝑟 ,𝑙 (𝑟) ( 2𝑚𝑟 2
) 𝑢𝑛𝑟,𝑙 (𝑟)𝑑𝑟; 8.23

and the energy 𝐸𝑛𝑟 ,𝑙 also increases with increasing 𝑛𝑟 because the radial kinetic energy is given by

∞ ℏ2 𝑑 2
∫0 𝑢𝑛∗ 𝑟,𝑙 (𝑟) (− 2𝑚 𝑑𝑟 2) 𝑢𝑛𝑟,𝑙 (𝑟)𝑑𝑟, 8.24

increases as radial nodes increases.

Quantum Mechanics of a Hydrogen Atom

The hydrogen atom is essentially an electron in a Coulomb potential, and a Coulomb potential is a
central potential. Therefore bound states of atomic hydrogen can be taken to have definite orbital
angular momentum properties given by 𝐿 = √𝑙(𝑙 + 1)ℏ and 𝐿𝑧 = 𝑚𝑙 ℏ. These states have wave
function of the form

𝜓𝑛𝑟,𝑙,𝑚𝑙 (𝑟, 𝜃, 𝜙, 𝑡) = 𝜓𝑛𝑟 ,𝑙,𝑚𝑙 (𝑟, 𝜃, 𝜙)𝑒 −𝑖𝐸𝑡/ℏ 8.25

with

𝑢𝑛𝑟 ,𝑙 (𝑟)
𝜓𝑛𝑟,𝑙,𝑚𝑙 (𝑟, 𝜃, 𝜙) = 𝑟
𝑌𝑙,𝑚𝑙 (𝜃, 𝜙), 8.26

where 𝑢𝑛𝑟 ,𝑙 (𝑟) is the eigenfunction for the radial Schrӧdinger equation for an electron in a Coulomb
potential

𝑒2
𝑉(𝑟) = − .
4𝜋𝜖0 𝑟

Specifically the radial function 𝑢𝑛𝑟 ,𝑙 (𝑟) is a solution to a differential equation

ℏ2 𝑑 2 𝑢𝑛𝑟 ,𝑙 (𝑟) 𝑙(𝑙+1)ℏ2 𝑒2


2𝑚𝑒 𝑑𝑟 2
+[ 2𝑚𝑒𝑟 2
− 4𝜋𝜖 𝑟] 𝑢𝑛𝑟,𝑙 (𝑟) = 𝐸𝑛𝑟 ,𝑙 𝑢𝑛𝑟,𝑙 (𝑟) 8.27
0

which satisfy the boundary conditions:


𝑢𝑛𝑟 ,𝑙 (𝑟) = 0 at 𝑟 = 0 and at 𝑟 = ∞. 8.28

The qualities of the energy levels given by the eigenvalue equation above may be deduced by
considering the effective potential

𝑙(𝑙+1)ℏ2 𝑒2
𝑉𝑒𝑓𝑓 (𝑟) = 2𝑚𝑒𝑟 2
− 4𝜋𝜖 𝑟. 8.29
0

The shape of this potential for an electron with different values of orbital angular momentum 𝑙 is
shown below in Figure 8.2:

Figure 8.2: Illustration of the effective potential energy of the electron in a hydrogen atom
with the orbital angular momentum quantum numbers l = 0, 1, 2, 3.
For non-zero values of 𝑙, the effective potential is attractive for large values of 𝑟 and repulsive for
𝑑𝑉
small values of 𝑟. By setting 𝑒𝑓𝑓⁄𝑑𝑟 equal to zero, it can be shown that 𝑉𝑒𝑓𝑓 (𝑟) has a minimum
value of

𝑅 𝐸
𝑉𝑒𝑓𝑓 (𝑟) = − 𝑙(𝑙+1) at 𝑟 = 𝑙(𝑙 + 1)𝑎0 , 8.30

where 𝑎0 and 𝐸𝑅 are the natural units of length and energy in atomic physics. These units are
defined as follows

4𝜋𝜖0 ℏ2
Bohr radius 𝑎0 = [ ]
𝑒 2 𝑚𝑒
= 0.529 × 10−10 m.
𝑒2
Rydbergy energy 𝐸𝑅 = 8𝜋𝜖 = 13.6 eV.
0 𝑎0

𝑅 𝐸
The minimum energy of 𝑉𝑒𝑓𝑓 (𝑟) = − 𝑙(𝑙+1) implies that the bound states with angular momentum
𝐸𝑅
𝐿 = √𝑙(𝑙 + 1)ℏ have energies between 𝐸 = − ⁄𝑙(𝑙 + 1) and 𝐸 = 0. It also implies that the
spatial extent of the bound states of eigenfunction with low angular momentum is in the order of 𝑎0
and that the eigenfunctions extend to larger distances when the angular momentum increases.
When the angular momentum greatly exceeds ℏ, many bound states with closely spaced energy
levels corresponding to circular and elliptic orbits of classical mechanics are expected.

It can also be seen from the above figure that the purely attractive effective potential exists only for
an electron with the zero angular momentum. In classical mechanics such an electron just rushes
into a proton and there are no stable bound states. Quantum mechanically, there such stable bound
states whose existence can be explained using the uncertainty principle.

The uncertainty principle explains that an electron localized in the region of size 𝑟, has uncertain
momentum of the order of ℏ/𝑟, and the average kinetic energy which at least of the order of
ℏ2⁄ . This means that the least energy an electron with zero orbital angular momentum in the
2𝑚𝑒 𝑟 2
region of size 𝑟 can roughly be

ℏ2 𝑒2
𝐸 ≈ 2𝑚 2 − 4𝜋𝜖 . 8.31
𝑒𝑟 0𝑟

As the region of localization decreases, this energy decreases because the potential energy
decreases, but eventually the kinetic energy of localization increases more rapidly. This results in
total energy having a minimum of

𝐸 ≈ −𝐸𝑅 at 𝑟 ≈ 𝑎0 . 8.32

This minimum provides an estimate of the lowest possible energy of an electron with zero orbital
angular momentum in a Coulomb potential and suggest that there are bound states with energies in
the range 𝐸 ≈ −𝐸𝑅 and 𝐸 = 0.

The energy levels and eigenfunction of an electron in a Coulomb potential are obtained by solving
the eigenvalue equation

ℏ2 𝑑 2 𝑢𝑛𝑟 ,𝑙 (𝑟) 𝑙(𝑙+1)ℏ2 𝑒2


2𝑚𝑒 𝑑𝑟 2
+[ 2𝑚𝑒 𝑟 2
− 4𝜋𝜖 𝑟] 𝑢𝑛𝑟 ,𝑙 (𝑟) = 𝐸𝑛𝑟,𝑙 𝑢𝑛𝑟 ,𝑙 (𝑟), 8.33
0

satisfying the boundary conditions 𝑢𝑛𝑟 ,𝑙 (𝑟) = 0 at 𝑟 = 0 and at 𝑟 = ∞. In order to focus on the
physical properties of the hydrogen atom, the results of mathematical problem are considered.

It can be shown that an electron in a Coulomb potential has infinite number of bound states with
energies given by
𝐸𝑅
𝐸 = − (𝑛 2 with 𝑛𝑟 = 0, 1, 2, 3, … 8.34
𝑟 +𝑙+1)

The quantum number 𝑛𝑟 is called the radial quantum number. These energy levels are illustrated
below:
Figure 8.3: Illustration of the energy levels according to the equation 8.34 for the bound
states of an electron in a hydrogen atom.
As expected there are bound states with zero and non-zero angular momentum.

The energy levels are closely spaced with very large angular momentum, signifying a resemblance
with the continuum of classical bound states. Many of the energy levels in the figure, with the same
value of 𝑛𝑟 + 𝑙, have the same energy. Due to this degeneracy, energy levels of the hydrogen atom
are given as
𝐸
𝐸 = − 𝑛𝑅2 , 8.35

where 𝐸𝑅 is the Rydberg energy and 𝑛 is the quantum number given by 𝑛 = 𝑛𝑟 + 𝑙 + 1.

This quantum number is the principal quantum number, and can take on values 𝑛 = 1, 2, 3, …

It can also be shown that the radial eigenfunction 𝑢𝑛𝑟,𝑙 (𝑟) belonging to the eigenvalue 𝐸𝑛𝑟 ,𝑙 has
the following three characteristics:

(1) In the square-well potential example, it has been found that the eigenfunction of a particle
2 2
with binding energy 𝐸 = ℏ 𝛼 ⁄2𝑚 falls off exponentially like 𝑒 −𝛼𝑥 . In a similar way the
radial eigenfunction of an electron in a Coulomb potential with binding energy
𝐸𝑅 ℏ2 1
𝐸= 𝑛2
= 2𝑚 𝑛2 𝑎02
, falls off exponentially at large 𝑟 like 𝑢𝑛𝑟 ,𝑙 (𝑟) ∝ 𝑒 −𝑟/𝑛𝑎0 .
𝑒
(2) Because of the singular nature at 𝑟 = 0 of the centrifugal potential energy
𝑙(𝑙 + 1)ℏ2⁄
, the behaviour of the eigenfunction at small 𝑟 governed by the orbital
2𝑚𝑒 𝑟 2
angular momentum and is given by 𝑢𝑛𝑟 ,𝑙 (𝑟) ∝ 𝑟 𝑙+1.
(3) The radial quantum number 𝑛𝑟 denotes the number of nodes between 𝑟 = 0 and 𝑟 = ∞,
the eigenfunction 𝑢𝑛𝑟 ,𝑙 (𝑟) is proportional to a polynomial with 𝑛𝑟 zeros. If this polynomial
is denoted by 𝑝𝑛𝑟,𝑙 (𝑟), we have
𝑢𝑛𝑟,𝑙 (𝑟) ∝ 𝑝𝑛𝑟 ,𝑙 (𝑟).

Combining these three characteristics we arrive at a radial eigenfunction of the form

∞ 2
𝑢𝑛𝑟,𝑙 (𝑟) = 𝑁𝑝𝑛𝑟 ,𝑙 (𝑟)𝑟 𝑙+1 𝑒 −𝑟/𝑛𝑎0 , where 𝑁 is the constant ensuring that the ∫0 |𝑢𝑛𝑟 ,𝑙 (𝑟)| 𝑑𝑟 = 1
normalization condition is satisfied.

Expressions of the radial eigenfunctions with low values of the angular momentum quantum number
𝑙 and low values of the radial quantum number 𝑛𝑟 are shown in the table below:
Table 8.1: Presentation of the normalised radial eigenfunctions for the low-lying states of
the hydrogen atom.
Some of these eigenfunctions are illustrated in the figure below:

Figure 8.4: A diagram demonstrating the radial eigenfunctions 𝒖𝒏,𝒓 (𝒓) for an electron in a
hydrogen atom with radial quantum numbers 𝒏𝒓 = 𝟎, 𝟏, 𝐚𝐧𝐝 𝟐 and with angular
momentum quantum numbers l = 0 and 1.
To conform to the convention of atomic physics, these eigenfunctions are labelled using the
spectroscopic notation. The notation uses the principal quantum number 𝑛 = 𝑛𝑟 + 𝑙 + 1 and a
letter to designate 𝑙; the letter 𝑠 is used for 𝑙 = 0, 𝑝 for 𝑙 = 1, 𝑑 for 𝑙 = 2, and 𝑓 for 𝑙 = 3.

You might also like