Download as pdf or txt
Download as pdf or txt
You are on page 1of 57

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

2022
MODULE OUTLINE

Module Title Quantum Mechanics


SPHA031 16
Module Code No. of
Credits
Physics Physical &
Department School Mineral
Pre-requisites SPHA021/SPHB021 Co- SMTA021
Module Code SMTA021 requisites
Module
Code
Prof TE Mosuang
Module Lecturer
Q-3049
Office Address
thuto.mosuang@ul.ac.za -3576
Email Telephone
No.
Module Dr PS Ntoahae (-3341)
evaluator/
Moderator/2nd
Examiner

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
Fundamental Quantum Mechanics concepts

Photons are particle-like quanta of electromagnetic radiation. They travel at the speed of light 𝑐𝑐 with
momentum 𝑝𝑝 and energy 𝜀𝜀 given by:
ℎ ℎ𝑐𝑐
𝑝𝑝 = and 𝜀𝜀 = where 𝜆𝜆 is the wavelength of the electromagnetic radiation. In reference to the
𝜆𝜆 𝜆𝜆
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 𝜆𝜆 = 663 nm are

𝑝𝑝 = 10−27 Js and 𝜀𝜀 = 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 𝑚𝑚𝑒𝑒 at rest through an angle 𝜃𝜃, the
wavelength is increased by

∆𝜆𝜆 = (1 − cos 𝜃𝜃). 1.1
𝑚𝑚𝑒𝑒 𝑐𝑐


It must be noted that this increase is set by a constant = 2.43 × 10−12 m called the Compton
𝑚𝑚𝑒𝑒 𝑐𝑐
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 𝑝𝑝 could have wavelength defined by:

𝜆𝜆 = . 1.2
𝑝𝑝

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 𝑝𝑝 and energy 𝜀𝜀 of a particle of mass 𝑚𝑚 is

𝜀𝜀 2 − 𝑝𝑝2 𝑐𝑐 2 = 𝑚𝑚2 𝑐𝑐 4. 1.3

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


ℎ𝑐𝑐
𝜆𝜆 = . 1.4
�(𝜀𝜀−𝑚𝑚𝑐𝑐 2 )(𝜀𝜀+𝑚𝑚𝑐𝑐 2 )

If the particle is ultra-relativistic, the mass energy 𝑚𝑚𝑐𝑐 2 term can be neglected to get
ℎ𝑐𝑐
𝜆𝜆 = , 1.5
𝜀𝜀

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

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

𝜀𝜀 = 𝑚𝑚𝑐𝑐 2 + 𝐾𝐾𝐾𝐾 , 1.6

𝑝𝑝2
and the particles non-relativistic kinetic energy is 𝐾𝐾𝐾𝐾 = 𝐸𝐸 = 2𝑚𝑚, and the particles wavelength is


𝜆𝜆 = . 1.7
√2𝑚𝑚𝑚𝑚

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 𝑚𝑚 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 𝑚𝑚 = 𝑚𝑚𝑒𝑒 = 9.109 × 10−31kg into
equation (1.7). Expressing the kinetic energy 𝐸𝐸 in electronvolts we get

1.5
𝜆𝜆 = � nm. 1.8
𝐸𝐸

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


𝜆𝜆 =
√2𝑚𝑚𝑚𝑚

ℎ 2 𝑐𝑐 2
= �2𝑚𝑚𝑐𝑐 2 𝐸𝐸

(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
𝐸𝐸 = − eV, 1.9
𝑛𝑛2

where 𝑛𝑛 is the principal quantum number which takes on the infinite number of values 1, 2, 3, … The
ground state of hydrogen has 𝑛𝑛 = 1, the first excited state has 𝑛𝑛 = 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 𝐸𝐸 = 0 to 𝐸𝐸 = ∞.

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 𝑛𝑛𝑖𝑖 to a
higher state 𝑛𝑛𝑓𝑓 , results in a spectral line with wavelength given by

ℎ𝑐𝑐
= �𝐸𝐸𝑛𝑛𝑖𝑖 − 𝐸𝐸𝑛𝑛𝑓𝑓 �. 1.10
𝜆𝜆

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 𝑚𝑚𝑒𝑒 , the size of an atom consisting of wave-like electrons also depends on ℎ and 𝑚𝑚𝑒𝑒 .
In addition, the atomic size must depend on the force that binds the electron to the nucleus of an
𝑒𝑒 2
atom: that is proportional to , where 𝑒𝑒 is the magnitude of charge of an electron and a proton.
4𝜋𝜋𝜖𝜖0
𝑒𝑒 2
So specifically, the size of an atom depends on 4𝜋𝜋𝜖𝜖 , ℎ and 𝑚𝑚𝑒𝑒 . Moreover, the accepted unit of
0
length at atomic scale is the Bohr radius which is given by

4𝜋𝜋𝜖𝜖0 ℏ2
𝑎𝑎0 = � �
𝑒𝑒 2 𝑚𝑚𝑒𝑒
= 0.529 × 10−10 m. 1.11

Form the natural unit of length 𝑎𝑎0 , the natural unit of atomic binding energy called the Rydberg
energy is given by
𝑒𝑒 2
𝐸𝐸𝑅𝑅 = 8𝜋𝜋𝜖𝜖 = 13.6 eV. 1.12
0 𝑎𝑎0

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

𝜆𝜆
∆𝑥𝑥 ≈ where 𝜆𝜆 is the wavelength of the incident light and 2𝛼𝛼 is the angle
sin 𝛼𝛼
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 𝜆𝜆 and momentum ℎ/𝜆𝜆 must have
sideways momentum components

− ℎ�𝜆𝜆 sin 𝛼𝛼 and + ℎ�𝜆𝜆 sin 𝛼𝛼

So the sideways momentum of the scattered photon is uncertain by

∆𝑝𝑝 ≈ ℎ�𝜆𝜆 sin 𝛼𝛼. 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

Δ𝑥𝑥Δ𝑝𝑝 ≈ 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
∆𝑥𝑥∆𝑝𝑝 ≥ where ℏ = . 1.15
2 2𝜋𝜋

The Heisenberg uncertainty principle suggest that an exact determination of the particles position,
i.e. ∆𝑥𝑥 = 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.

𝐹𝐹⃗ = 𝑚𝑚𝑎𝑎⃗ = 𝑚𝑚𝑟𝑟⃗̈. 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 𝜆𝜆 and period 𝜏𝜏. In the same
way this can be represented by the wave number 𝑘𝑘 = 2𝜋𝜋�𝜆𝜆 and angular frequency 𝜔𝜔 = 2𝜋𝜋�𝜏𝜏. Such
a wave is represented by a mathematical function

𝜓𝜓(𝑥𝑥, 𝑡𝑡) = 𝐴𝐴 cos(𝑘𝑘𝑥𝑥 − 𝜔𝜔𝜔𝜔). 2.2


2𝜋𝜋
𝐴𝐴 is a constant for the amplitude. At point 𝑥𝑥, the function oscillates with amplitude 𝐴𝐴 and period .
𝜔𝜔
2𝜋𝜋
At time 𝑡𝑡, the function undulates with amplitude 𝐴𝐴 and wavelength 𝑘𝑘
. On top of that these
undulations the wave function exhibit a Mexican wave, in the direction of the increasing 𝑥𝑥 with
velocity 𝜔𝜔/𝑘𝑘. For the case in point, the maximum of 𝜓𝜓(𝑥𝑥, 𝑡𝑡) corresponds to 𝑘𝑘𝑘𝑘 − 𝜔𝜔𝜔𝜔 = 0, which
occurs at position 𝑥𝑥 = 𝜔𝜔𝜔𝜔/𝑘𝑘. The minimum of 𝜓𝜓(𝑥𝑥, 𝑡𝑡) corresponds to 𝑘𝑘𝑘𝑘 − 𝜔𝜔𝜔𝜔 = 𝜋𝜋 and it occurs at
𝜆𝜆 𝜔𝜔𝜔𝜔
position 𝑥𝑥 = + . In both cases the position moves with velocity 𝜔𝜔/𝑘𝑘.
2 𝑘𝑘

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

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

𝜓𝜓(𝑥𝑥, 𝑡𝑡) = 𝐴𝐴 cos(𝑘𝑘𝑘𝑘 − 𝜔𝜔𝜔𝜔) + 𝐵𝐵 sin(𝑘𝑘𝑘𝑘 − 𝜔𝜔𝜔𝜔),

𝐴𝐴 and 𝐵𝐵 are constants.

Occasionally in classical physics and invariably in quantum physics, sinusoidal travelling waves are
represented by the complex exponential function
𝜓𝜓(𝑥𝑥, 𝑡𝑡) = 𝐴𝐴𝑒𝑒 𝑖𝑖(𝑘𝑘𝑘𝑘−𝜔𝜔𝜔𝜔) . 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 𝐴𝐴 cos(𝑘𝑘𝑘𝑘 − 𝜔𝜔𝜔𝜔), but
the real function can be taken from the real part of a complex exponential function 𝐴𝐴𝑒𝑒 𝑖𝑖(𝑘𝑘𝑘𝑘−𝜔𝜔𝜔𝜔)
because

𝐴𝐴𝑒𝑒 𝑖𝑖(𝑘𝑘𝑘𝑘−𝜔𝜔𝜔𝜔) = 𝐴𝐴[cos(𝑘𝑘𝑘𝑘 − 𝜔𝜔𝜔𝜔) + 𝑖𝑖 sin(𝑘𝑘𝑘𝑘 − 𝜔𝜔𝜔𝜔)].

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

𝐴𝐴 cos(𝑘𝑘𝑘𝑘 − 𝜔𝜔𝜔𝜔) + 𝐴𝐴 cos(𝑘𝑘𝑘𝑘 + 𝜔𝜔𝜔𝜔) give rise to the wave 2𝐴𝐴 cos 𝑘𝑘𝑘𝑘 cos 𝜔𝜔𝜔𝜔.

This wave oscillates with period 2𝜋𝜋/𝜔𝜔 and undulates with wavelength 2𝜋𝜋/𝑘𝑘. 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 𝐴𝐴 and wave numbers in the range
𝑘𝑘 − Δ𝑘𝑘 to 𝑘𝑘 + Δ𝑘𝑘 is

𝑘𝑘+Δ𝑘𝑘
𝜓𝜓(𝑥𝑥, 𝑡𝑡) = ∫𝑘𝑘−Δ𝑘𝑘 𝐴𝐴 cos(𝑘𝑘′𝑥𝑥 − 𝜔𝜔′𝑡𝑡)𝑑𝑑𝑑𝑑′. 2.4

If 𝑘𝑘 is positive, the wave packet travels in the positive 𝑥𝑥 direction, and to the negative 𝑥𝑥 direction
when 𝑘𝑘 in negative.

The initial shape of the wave packet, i.e. a shape at 𝑡𝑡 = 0, may be obtained by evaluating the
integral

𝑘𝑘+Δ𝑘𝑘
𝜓𝜓(𝑥𝑥, 0) = ∫𝑘𝑘−Δ𝑘𝑘 𝐴𝐴 cos 𝑘𝑘′𝑥𝑥𝑥𝑥𝑥𝑥′ 2.5

sin ∆𝑘𝑘𝑘𝑘
this gives 𝜓𝜓(𝑥𝑥, 0) = 𝒮𝒮(𝑥𝑥) cos 𝑘𝑘𝑘𝑘 where 𝒮𝒮(𝑥𝑥) = 2𝐴𝐴∆𝑘𝑘 .
(∆𝑘𝑘𝑘𝑘)

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 𝜔𝜔 = 𝑐𝑐𝑐𝑐, where 𝑐𝑐 is a constant such that
for sinusoidal waves, the velocity, 𝑐𝑐 = 𝜔𝜔/𝑘𝑘, is independent of the wave number 𝑘𝑘.

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 𝑥𝑥 direction, the wave equation has the form

𝜕𝜕2 𝜓𝜓 1 𝜕𝜕2 𝜓𝜓
= , 2.6
𝜕𝜕𝜕𝜕 2 𝑐𝑐 2 𝜕𝜕𝜕𝜕 2

1 𝜕𝜕2 𝜓𝜓 𝜕𝜕2 𝜕𝜕2 𝜕𝜕2


and in three dimensions ∇2 𝜓𝜓 = , where ∇2 = + + .
𝑐𝑐 2 𝜕𝜕𝜕𝜕 2 𝜕𝜕𝜕𝜕 2 𝜕𝜕𝜕𝜕 2 𝜕𝜕𝜕𝜕 2

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

𝐴𝐴 cos(𝑘𝑘𝑘𝑘 − 𝜔𝜔𝜔𝜔), 𝐴𝐴 sin(𝑘𝑘𝑘𝑘 − 𝜔𝜔𝜔𝜔), or 𝐴𝐴𝑒𝑒 𝑖𝑖(𝑘𝑘𝑘𝑘−𝜔𝜔𝜔𝜔) 2.7

as solutions provided 𝜔𝜔2 = 𝑐𝑐 2 𝑘𝑘 2 , which can be shown by direct substitution on the classical wave
equation.

Solutions with 𝑐𝑐 = +𝜔𝜔/𝑘𝑘, describe waves travelling in the positive 𝑥𝑥 direction, whilst solutions with
𝑐𝑐 = −𝜔𝜔/𝑘𝑘, describe waves travelling in the negative 𝑥𝑥 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, 𝜔𝜔 =
𝑐𝑐𝑐𝑐 is more complicated so that the velocity, 𝜔𝜔/𝑘𝑘 of a propagating wave depends on the wave
number, 𝑘𝑘. 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

𝑑𝑑𝑑𝑑
𝜐𝜐𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔 = , 2.8
𝑑𝑑𝑑𝑑

though the velocity of a simple sinusoidal wave is called the phase velocity
𝜔𝜔
𝜐𝜐𝑝𝑝ℎ𝑎𝑎𝑎𝑎𝑎𝑎 = . 2.9
𝑘𝑘

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 𝑘𝑘1 and
𝑘𝑘2 and angular frequencies 𝜔𝜔1 and 𝜔𝜔2 , mathematically this is

𝑘𝑘1 𝑥𝑥 − 𝜔𝜔1 𝑡𝑡 = 𝑘𝑘2 𝑥𝑥 − 𝜔𝜔2 𝑡𝑡.

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

𝜔𝜔1 −𝜔𝜔2
𝑥𝑥 = � � 𝑡𝑡 . 2.10
𝑘𝑘1 −𝑘𝑘2
So, the point of constructive interference is located at 𝑥𝑥 = 0, when 𝑡𝑡 = 0 and it travels with a
velocity given by 𝜔𝜔1 − 𝜔𝜔2 /𝑘𝑘1 − 𝑘𝑘2 , or eq (..) if |𝑘𝑘1 − 𝑘𝑘2 | 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, 𝑑𝑑𝑑𝑑�𝑑𝑑𝑑𝑑 can be obtained.

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

𝜔𝜔 = �𝑔𝑔𝑔𝑔 , where 𝑔𝑔 is the acceleration due to gravity.

The phase velocity of the sinusoidal wave is

𝜔𝜔 𝑔𝑔
𝜐𝜐𝑝𝑝ℎ𝑎𝑎𝑎𝑎𝑎𝑎 = =� .
𝑘𝑘 𝑘𝑘

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

𝑑𝑑𝑑𝑑 1 𝑔𝑔
𝜐𝜐𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔 = = � .
𝑑𝑑𝑑𝑑 2 𝑘𝑘

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 𝒌𝒌 − ∆𝒌𝒌 to 𝒌𝒌 + ∆𝒌𝒌.

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 𝑝𝑝 has de Broglie wavelength given by 𝜆𝜆 = ℎ/𝑝𝑝.

- this means that a de Broglie wave with wave number 𝑘𝑘 = 2𝜋𝜋/𝜆𝜆 describes a particle with
momentum

𝑝𝑝 = ℏ𝑘𝑘 , where ℏ=
2𝜋𝜋

- to extend this argument, assume that a de Broglie wave packet with a range of wave
numbers from 𝑘𝑘 − Δ𝑘𝑘 to 𝑘𝑘 + Δ𝑘𝑘 describes a particle with an uncertain momentum
Δ𝑝𝑝 ≈ ℏ∆𝑘𝑘
- assume also that the length of this wave packet is a measure of Δ𝑥𝑥, the uncertainty in the
position of the particle.
2𝜋𝜋
∆𝑥𝑥 ≈
∆𝑘𝑘
- using the equation Δ𝑝𝑝 ≈ ℏ∆𝑘𝑘 and the figure …. as a model, we write
- multiplying these two uncertainties we obtain
∆𝑥𝑥∆𝑝𝑝 ≈ 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 𝑚𝑚
and momentum 𝑝𝑝 = ℏ𝑘𝑘 requires that
𝑑𝑑𝑑𝑑 ℏ𝑘𝑘
=
𝑑𝑑𝑑𝑑 𝑚𝑚
- this equation can be integrated to give the following dispersion relation for the de Broglie
waves describing a free moving quantum particle of mass 𝑚𝑚:
ℏ𝑘𝑘 2
𝜔𝜔 = . 2.11
2𝑚𝑚
- 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
𝜕𝜕𝜕𝜕 ℏ2 𝜕𝜕2 𝜓𝜓
𝑖𝑖ℏ =− . 2.12
𝜕𝜕𝜕𝜕 2𝑚𝑚 𝜕𝜕𝜕𝜕 2
- the complex exponential function 𝜓𝜓(𝑥𝑥, 𝑡𝑡) = 𝐴𝐴𝑒𝑒 𝑖𝑖(𝑘𝑘𝑘𝑘−𝜔𝜔𝜔𝜔) is the solution of this
ℏ𝑘𝑘 2
equation provided it obey the dispersion relation 𝜔𝜔 = .
2𝑚𝑚
- substituting 𝜓𝜓(𝑥𝑥, 𝑡𝑡) on the left hand side of eq … yields
𝜕𝜕𝜕𝜕
𝑖𝑖ℏ = 𝑖𝑖ℏ(−𝑖𝑖𝑖𝑖)𝐴𝐴𝑒𝑒 𝑖𝑖(𝑘𝑘𝑘𝑘−𝜔𝜔𝜔𝜔) = ℏ𝜔𝜔𝜔𝜔𝑒𝑒 𝑖𝑖(𝑘𝑘𝑘𝑘−𝜔𝜔𝜔𝜔) = ℏ𝜔𝜔𝜔𝜔(𝑥𝑥, 𝑡𝑡)
𝜕𝜕𝜕𝜕
- the right hand side yields
ℏ2 𝜕𝜕2 𝜓𝜓 ℏ2 𝜕𝜕2 ℏ2 ℏ2 𝑘𝑘 2
− 2𝑚𝑚 = − 2𝑚𝑚 𝜕𝜕𝜕𝜕 2 𝐴𝐴𝑒𝑒 𝑖𝑖(𝑘𝑘𝑘𝑘−𝜔𝜔𝜔𝜔) = − 2𝑚𝑚 (−𝑘𝑘 2 )𝐴𝐴𝑒𝑒 𝑖𝑖(𝑘𝑘𝑘𝑘−𝜔𝜔𝜔𝜔) = 𝜓𝜓(𝑥𝑥, 𝑡𝑡)
𝜕𝜕𝜕𝜕 2 2𝑚𝑚
- the solution can be obtained provided
ℏ2 𝑘𝑘 2
ℏ𝜔𝜔 = . 2.13
2𝑚𝑚
- the sinusoidal wave solution, 𝐴𝐴𝑒𝑒 𝑖𝑖(𝑘𝑘𝑘𝑘−𝜔𝜔𝜔𝜔) , describes a wave moving in the x-direction with
wave number 𝑘𝑘 and angular frequency 𝜔𝜔.
- we assume that the wave represents a free particle moving in the x-direction with
𝑝𝑝2
momentum 𝑝𝑝 = ℏ𝑘𝑘 and energy 𝐸𝐸 = = ℏ𝜔𝜔. 2.14
2𝑚𝑚
- 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
ℏ𝑘𝑘 2
Broglie waves, 𝜔𝜔 = , we arrived at a wave equation, the free-particle Schrӧdinger
2𝑚𝑚
𝜕𝜕𝜕𝜕 ℏ 𝜕𝜕2 𝜓𝜓
2
equation, 𝑖𝑖ℏ 𝜕𝜕𝜕𝜕 = − 2𝑚𝑚 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 𝜓𝜓, a superposition of
solutions is also a solution. e.g.

𝜓𝜓(𝑥𝑥, 𝑡𝑡) = 𝐴𝐴1 𝑒𝑒 𝑖𝑖(𝑘𝑘1 𝑥𝑥−𝜔𝜔1 𝑡𝑡) + 𝐴𝐴2 𝑒𝑒 𝑖𝑖(𝑘𝑘2 𝑥𝑥−𝜔𝜔2 𝑡𝑡) 2.15

ℏ2 𝑘𝑘12 ℏ2 𝑘𝑘22
with ℏ𝜔𝜔1 = and ℏ𝜔𝜔2 = ,
2𝑚𝑚 2𝑚𝑚

where 𝐴𝐴1 and 𝐴𝐴2 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:

+∞ ′ 𝑥𝑥−𝜔𝜔′ 𝑡𝑡)
𝜓𝜓(𝑥𝑥, 𝑡𝑡) = ∫−∞ 𝐴𝐴(𝑘𝑘 ′ )𝑒𝑒 𝑖𝑖(𝑘𝑘 𝑑𝑑𝑑𝑑′ 2.16

ℏ2 𝑘𝑘′2
with ℏ𝜔𝜔′ = 2𝑚𝑚
.

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

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

𝑉𝑉(𝑟𝑟) = −𝑒𝑒 2 /4𝜋𝜋𝜖𝜖0 𝑟𝑟 2.17

of the nucleus in a hydrogen atom.

Classically this means that an electron at a distance 𝑟𝑟 from the nucleus experiences an attractive
force 𝑒𝑒 2 /4𝜋𝜋𝜖𝜖0 𝑟𝑟 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 𝑉𝑉(𝑟𝑟⃗) is
given by

𝜕𝜕𝜓𝜓 ℏ2
𝑖𝑖ℏ = �− ∇2 + 𝑉𝑉(𝑟𝑟⃗)� 𝜓𝜓. 2.18
𝜕𝜕𝜕𝜕 2𝑚𝑚

In a one dimensional potential 𝑉𝑉(𝑥𝑥) along the 𝑥𝑥, Schrӧdinger equation simplifies to

𝜕𝜕𝜕𝜕 ℏ2 𝜕𝜕2
𝑖𝑖ℏ = �− + 𝑉𝑉(𝑥𝑥)� 𝜓𝜓. 2.19
𝜕𝜕𝜕𝜕 2𝑚𝑚 𝜕𝜕𝜕𝜕 2

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 𝑥𝑥, with a constant potential 𝑉𝑉0 its solution is

𝜓𝜓(𝑥𝑥, 𝑡𝑡) = 𝐴𝐴𝑒𝑒 𝑖𝑖(𝑘𝑘𝑘𝑘−𝜔𝜔𝜔𝜔) 2.20

ℏ2 𝑘𝑘 2
provided ℏ𝜔𝜔 = 2𝑚𝑚
+ 𝑉𝑉0. 2.21

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

𝑝𝑝2
𝐸𝐸 = + 𝑉𝑉0 and 𝑝𝑝 = ℏ𝑘𝑘.
2𝑚𝑚

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 𝑥𝑥0 , 𝑥𝑥1 , 𝑥𝑥2 , 𝑥𝑥3 … with probabilities 𝑝𝑝0 , 𝑝𝑝1 , 𝑝𝑝2 , 𝑝𝑝3 , … The set of probabilities 𝑝𝑝𝑛𝑛 is
called a probability distribution. A total probability of all possible distribution must equal to unity.
Therefore the probability distribution 𝑝𝑝𝑛𝑛 must satisfy the normalization condition

∑𝑛𝑛 𝑝𝑝𝑛𝑛 = 1. 3.1

For example if the probability 𝑝𝑝𝑛𝑛 that a student who is doing this module is doing other 𝑛𝑛 modules
to complete his Physics course is

𝑝𝑝0 + 𝑝𝑝1 + 𝑝𝑝2 + 𝑝𝑝3 = 1.


The probability distribution 𝑝𝑝𝑛𝑛 can be used to evaluate the expectation values for the random
variables 𝑥𝑥𝑛𝑛 . 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

〈𝑥𝑥〉 = ∑all 𝑛𝑛 𝑥𝑥𝑛𝑛 𝑝𝑝𝑛𝑛 . 3.2

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

(∆𝑥𝑥)2 = ∑all 𝑛𝑛(𝑥𝑥𝑛𝑛 − 〈𝑥𝑥〉)2 𝑝𝑝𝑛𝑛 . 3.3

In this expression, (𝑥𝑥𝑛𝑛 − 〈𝑥𝑥〉) is the deviation of 𝑥𝑥𝑛𝑛 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:

(𝑥𝑥𝑛𝑛 − 〈𝑥𝑥〉)2 = 𝑥𝑥𝑛𝑛2 − 2𝑥𝑥𝑛𝑛 〈𝑥𝑥〉 + 〈𝑥𝑥〉2 , 3.4

recalling that 〈𝑥𝑥〉 is a number that does not depend on 𝑛𝑛, then

(∆𝑥𝑥)2 = ∑all 𝑛𝑛 𝑥𝑥𝑛𝑛2 𝑝𝑝𝑛𝑛 − 2〈𝑥𝑥〉 ∑all 𝑛𝑛 𝑥𝑥𝑛𝑛 𝑝𝑝𝑛𝑛 + 〈𝑥𝑥〉2 ∑all 𝑛𝑛 𝑝𝑝𝑛𝑛

= 〈𝑥𝑥 2 〉 − 2〈𝑥𝑥〉2 + 〈𝑥𝑥〉2

= 〈𝑥𝑥 2 〉 − 〈𝑥𝑥〉2.

∆𝑥𝑥 = �〈𝑥𝑥 2 〉 − 〈𝑥𝑥〉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 𝑥𝑥.
The probability of an outcome between 𝑥𝑥 and 𝑥𝑥 + 𝑑𝑑𝑑𝑑 is denoted by 𝜌𝜌(𝑥𝑥)𝑑𝑑𝑑𝑑. The function 𝜌𝜌(𝑥𝑥) is
called the probability density. It satisfy the normalization condition

∫all 𝑛𝑛 𝜌𝜌(𝑥𝑥)𝑑𝑑𝑑𝑑 = 1. 3.6

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


𝑎𝑎
∫0 𝜌𝜌(𝑥𝑥)𝑑𝑑𝑑𝑑 = 1.

The expectation value 〈𝑥𝑥〉 in analogy with the discrete variables is given by

〈𝑥𝑥〉 = ∫all 𝑛𝑛 𝑥𝑥𝑥𝑥(𝑥𝑥)𝑑𝑑𝑑𝑑. 3.7

Similarly, the expectation value of 𝑥𝑥 2 is given by


〈𝑥𝑥 2 〉 = ∫all 𝑛𝑛 𝑥𝑥 2 𝜌𝜌(𝑥𝑥)𝑑𝑑𝑑𝑑, 3.8

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

(∆𝑥𝑥) = �〈𝑥𝑥 2 〉 − 〈𝑥𝑥〉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 |𝜓𝜓(𝑟𝑟⃗, 𝑡𝑡)|2 , is a measure of the probability of finding a particle at the point 𝑟𝑟⃗, in time 𝑡𝑡.

The particle can be found anywhere but is more likely to be found where |𝜓𝜓(𝑟𝑟⃗, 𝑡𝑡)|2 is large.

𝑡𝑡ℎ𝑒𝑒 𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝 𝑜𝑜𝑜𝑜 𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓 𝑡𝑡ℎ𝑒𝑒 𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝


Specifically |𝜓𝜓(𝑟𝑟⃗, 𝑡𝑡)|2 𝑑𝑑3 𝑟𝑟⃗ = � � 3.9
𝑎𝑎𝑎𝑎 𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 𝑡𝑡 𝑖𝑖𝑖𝑖 𝑡𝑡ℎ𝑒𝑒 𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣 𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 𝑑𝑑 3 𝑟𝑟⃗

|𝜓𝜓(𝑟𝑟⃗, 𝑡𝑡)|2 , can also be thought as the probability density of position. 𝜓𝜓(𝑟𝑟⃗, 𝑡𝑡), 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

∫|𝜓𝜓(𝑟𝑟⃗, 𝑡𝑡)|2 𝑑𝑑 3 𝑟𝑟⃗ = 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 𝑥𝑥 axis then it can be described by the wavefunction 𝜓𝜓(𝑥𝑥, 𝑡𝑡) such that

𝑡𝑡ℎ𝑒𝑒 𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝 𝑜𝑜𝑜𝑜 𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓 𝑡𝑡ℎ𝑒𝑒 𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝


|𝜓𝜓(𝑥𝑥, 𝑡𝑡)|2 𝑑𝑑𝑑𝑑 = � �. 3.11
𝑎𝑎𝑎𝑎 𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 𝑡𝑡 𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏 𝑥𝑥 𝑎𝑎𝑎𝑎𝑎𝑎 𝑥𝑥 + 𝑑𝑑𝑑𝑑

Because the particle must be found somewhere between −∞ and +∞, it must also obey the
normalization condition

∫|𝜓𝜓(𝑥𝑥, 𝑡𝑡)|2 𝑑𝑑𝑑𝑑 = 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 |𝑓𝑓(𝑥𝑥)| can be integrated, i.e. ∫−∞ |𝑓𝑓(𝑥𝑥)|𝑑𝑑𝑑𝑑 < ∞, then one can always write a Fourier
expansion
+∞ 𝑑𝑑𝑑𝑑
𝑓𝑓(𝑥𝑥) = ∫−∞ 𝑔𝑔(𝑘𝑘)𝑒𝑒 𝑖𝑖𝑖𝑖𝑖𝑖 ............................... 3.12
√2𝜋𝜋

 𝑔𝑔(𝑘𝑘) is called the Fourier transform of the function 𝑓𝑓(𝑥𝑥) and is given by:
+∞ 𝑑𝑑𝑑𝑑
𝑔𝑔(𝑘𝑘) = ∫−∞ 𝑓𝑓(𝑥𝑥)𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖 . ................................ 3.13
√2𝜋𝜋

Self-consistency:
+∞ 𝑑𝑑𝑑𝑑
𝑔𝑔(𝑘𝑘) = ∫−∞ 𝑓𝑓(𝑥𝑥)𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖
√2𝜋𝜋

+∞ 𝑑𝑑𝑑𝑑 −𝑖𝑖𝑖𝑖𝑖𝑖 +∞ 𝑑𝑑𝑘𝑘 ′ 𝑖𝑖𝑘𝑘 ′ 𝑥𝑥


= ∫−∞ 𝑒𝑒 ∫−∞ √2𝜋𝜋 𝑒𝑒 𝑔𝑔(𝑘𝑘 ′ )
√2𝜋𝜋

+∞ +∞ 𝑑𝑑𝑑𝑑 ′ −𝑘𝑘�𝑥𝑥
= ∫−∞ 𝑑𝑑𝑘𝑘 ′ 𝑔𝑔(𝑘𝑘 ′ ) �∫−∞ 2𝜋𝜋
𝑒𝑒 𝑖𝑖�𝑘𝑘 � 3.14

+∞ 𝑑𝑑𝑑𝑑 ′ −𝑘𝑘�𝑥𝑥
 so what kind of a function is ∫−∞ 𝑒𝑒 𝑖𝑖�𝑘𝑘 ...?
2𝜋𝜋

Dirac delta function

a Dirac delta function 𝛿𝛿(𝑥𝑥 − 𝑥𝑥 ′ ) is defined by

(i) 𝛿𝛿(𝑥𝑥 − 𝑥𝑥 ′ ) = 0 if 𝑥𝑥 ≠ 𝑥𝑥′ and 𝛿𝛿(𝑥𝑥 − 𝑥𝑥 ′ ) = ∞ if 𝑥𝑥 = 𝑥𝑥′


(ii) ∫ 𝑑𝑑𝑑𝑑𝑑𝑑(𝑥𝑥 − 𝑥𝑥 ′ ) = 1
(iii) ∫ 𝑑𝑑𝑑𝑑𝑑𝑑(𝑥𝑥)𝛿𝛿(𝑥𝑥 − 𝑥𝑥 ′ ) = 𝑓𝑓(𝑥𝑥 ′ )

+∞ ′ −𝑘𝑘�𝑥𝑥
Now if ∫−∞ 𝑑𝑑𝑑𝑑𝑒𝑒 𝑖𝑖�𝑘𝑘 = 2𝜋𝜋𝜋𝜋(𝑘𝑘 ′ − 𝑘𝑘) then equation 3.14 is satisfied because

+∞ 1
∫−∞ 𝑑𝑑𝑘𝑘 ′ 𝑔𝑔(𝑘𝑘 ′ ) 2𝜋𝜋 2𝜋𝜋𝜋𝜋(𝑘𝑘 ′ − 𝑘𝑘) = 𝑔𝑔(𝑘𝑘) 3.15
so all in all we have
+∞ ′ −𝑘𝑘�𝑥𝑥
∫−∞ 𝑑𝑑𝑑𝑑𝑒𝑒 𝑖𝑖�𝑘𝑘 = 2𝜋𝜋𝜋𝜋(𝑘𝑘 ′ − 𝑘𝑘) 3.16a

+∞ ′ −𝑥𝑥�
∫−∞ 𝑑𝑑𝑑𝑑𝑒𝑒 𝑖𝑖𝑖𝑖�𝑥𝑥 = 2𝜋𝜋𝜋𝜋(𝑥𝑥′ − 𝑥𝑥) 3.16b

Property of Fourier transform


+∞ +∞ ∞ ∞
𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑′
� 𝑑𝑑𝑑𝑑|𝑓𝑓(𝑥𝑥)|2 = � 𝑑𝑑𝑑𝑑 �� 𝑔𝑔∗ (𝑘𝑘)𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖 � �� 𝑔𝑔(𝑘𝑘 ′ )𝑒𝑒 𝑖𝑖𝑖𝑖𝑖𝑖 �
−∞ −∞ −∞ √2𝜋𝜋 −∞ √2𝜋𝜋

∞ 𝑑𝑑𝑑𝑑 ∞ 𝑑𝑑𝑘𝑘 ′ ∞ ′ −𝑘𝑘�𝑥𝑥


= ∫−∞ 𝑔𝑔∗ (𝑘𝑘) ∫−∞ 𝑔𝑔(𝑘𝑘 ′ ) ∫−∞ 𝑑𝑑𝑑𝑑𝑒𝑒 𝑖𝑖�𝑘𝑘
√2𝜋𝜋 √2𝜋𝜋

∞ 𝑑𝑑𝑑𝑑 ∞ 𝑑𝑑𝑘𝑘 ′
= ∫−∞ 𝑔𝑔∗ (𝑘𝑘) ∫−∞ 𝑔𝑔(𝑘𝑘 ′ ) 2𝜋𝜋𝜋𝜋(𝑘𝑘 ′ − 𝑘𝑘)
√2𝜋𝜋 √2𝜋𝜋

+∞
= ∫−∞ 𝑑𝑑𝑑𝑑𝑔𝑔∗ (𝑘𝑘)𝑔𝑔(𝑘𝑘)

+∞ +∞
⇒ ∫−∞ 𝑑𝑑𝑑𝑑|𝑓𝑓(𝑥𝑥)|2 = ∫−∞ 𝑑𝑑𝑑𝑑|𝑔𝑔(𝑘𝑘)|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

𝜓𝜓(𝑥𝑥, 𝑡𝑡) = 𝐴𝐴1 𝑒𝑒 𝑖𝑖(𝑘𝑘1 𝑥𝑥−𝜔𝜔1 𝑡𝑡) + 𝐴𝐴2 𝑒𝑒 𝑖𝑖(𝑘𝑘2 𝑥𝑥−𝜔𝜔2 𝑡𝑡) , which describes a free particle moving along the
𝑥𝑥 direction with momenta and energies

𝑝𝑝1 = ℏ𝑘𝑘1 , 𝐸𝐸1 = ℏ𝜔𝜔1 and 𝑝𝑝2 = ℏ𝑘𝑘2 , 𝐸𝐸2 = ℏ𝜔𝜔2 . 3.18

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

+∞
𝜓𝜓(𝑥𝑥, 𝑡𝑡) = ∫−∞ 𝐴𝐴(𝑘𝑘)𝑒𝑒 𝑖𝑖(𝑘𝑘𝑘𝑘−𝜔𝜔𝜔𝜔) 𝑑𝑑𝑑𝑑 , 3.19

ℏ2 𝑘𝑘 2
with ℏ𝜔𝜔 = 2𝑚𝑚
.
In wave notation, this is a superposition of sinusoidal waves with different wave number 𝑘𝑘. The
magnitude of |𝐴𝐴(𝑘𝑘)|2 , measures the intensity of sinusoidal waves with wave number 𝑘𝑘.

According to particles, this range of wave numbers corresponds to a range of possible momenta
𝑝𝑝 = ℏ𝑘𝑘.

Using the same analogy as with Born interpretation of wavefunction, the most probable momenta
found in a measurement correspond to the value ℏ𝑘𝑘 for which the function |𝐴𝐴(𝑘𝑘)|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 +∞
𝜓𝜓(𝑥𝑥, 𝑡𝑡) = ∫−∞
𝜓𝜓�(𝑝𝑝, 𝑡𝑡) 𝑒𝑒 +𝑖𝑖𝑖𝑖𝑖𝑖/ℏ 𝑑𝑑𝑑𝑑. 3.20
√2𝜋𝜋ℏ

The inverse Fourier transform is


1 +∞
𝜓𝜓�(𝑝𝑝, 𝑡𝑡) = ∫ 𝜓𝜓(𝑥𝑥, 𝑡𝑡)𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖/ℏ 𝑑𝑑𝑑𝑑. 3.21
√2𝜋𝜋ℏ −∞

It can also be shown that if 𝜓𝜓(𝑥𝑥, 𝑡𝑡) is normalized, then 𝜓𝜓�(𝑝𝑝, 𝑡𝑡) is also normalized;

+∞ +∞ 2
i.e. if ∫−∞ |𝜓𝜓(𝑥𝑥, 𝑡𝑡)|2 𝑑𝑑𝑑𝑑 = 1, then ∫−∞ �𝜓𝜓�(𝑝𝑝, 𝑡𝑡)� 𝑑𝑑𝑝𝑝 = 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:

• |𝜓𝜓(𝑥𝑥, 𝑡𝑡)|2 is the probability density for finding a particle with position 𝑥𝑥.
2
• �𝜓𝜓�(𝑝𝑝, 𝑡𝑡)� is the probability density for finding a particle with momentum 𝑝𝑝.
In addition, since 𝜓𝜓(𝑥𝑥, 𝑡𝑡) is the probability amplitude for position, its Fourier transform 𝜓𝜓�(𝑝𝑝, 𝑡𝑡) 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 𝑚𝑚confined in a one
dimensional region 0 ≤ 𝑥𝑥 ≤ 𝑎𝑎. Such a particle in a state with quantum number 𝑛𝑛 has an energy
given by

ℏ2 𝑘𝑘𝑛𝑛
2 𝑛𝑛𝑛𝑛
𝐸𝐸𝑛𝑛 = 2𝑚𝑚
, with 𝑘𝑘𝑛𝑛 = , 3.23
𝑎𝑎

and a wavefunction

−𝑖𝑖𝐸𝐸𝑛𝑛 𝑡𝑡/ℏ
𝜓𝜓(𝑥𝑥, 𝑡𝑡) = �𝑁𝑁 sin 𝑘𝑘𝑛𝑛 𝑥𝑥𝑒𝑒 0 < 𝑥𝑥 < 𝑎𝑎. 3.24
0 elsewhere
The normalization constant 𝑁𝑁 can be obtained by normalizing the position probability density. This is
given by |𝜓𝜓(𝑥𝑥, 𝑡𝑡)|2 ; this is zero outside the region 0 < 𝑥𝑥 < 𝑎𝑎, and inside this region it is given by

𝑖𝑖𝐸𝐸𝑛𝑛 𝑡𝑡� 𝑖𝑖𝐸𝐸𝑛𝑛 𝑡𝑡�


|𝜓𝜓(𝑥𝑥, 𝑡𝑡)|2 = 𝑁𝑁 ∗ sin 𝑘𝑘𝑛𝑛 𝑥𝑥𝑒𝑒 + ℏ 𝑁𝑁 sin 𝑘𝑘𝑛𝑛 𝑥𝑥 𝑒𝑒 − ℏ = 𝑁𝑁 2 sin2 𝑘𝑘𝑛𝑛 𝑥𝑥. 3.25

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

+∞ 𝑎𝑎 𝑎𝑎
∫−∞ |𝜓𝜓(𝑥𝑥, 𝑡𝑡)|2 𝑑𝑑𝑑𝑑 = |𝑁𝑁|2 ∫0 sin2 𝑘𝑘𝑘𝑘𝑘𝑘𝑘𝑘 = |𝑁𝑁|2 2 . 3.26

2
By equating this probability to one, 𝑁𝑁 = � give rise to a normalized probability density and to a
𝑎𝑎
normalized wavefunction.
2
The momentum probability density is given by �𝜓𝜓�𝑛𝑛 (𝑝𝑝, 𝑡𝑡)� , where according to the momentum
probability amplitude

1 +∞
𝜓𝜓�𝑛𝑛 (𝑝𝑝, 𝑡𝑡) = ∫−∞
𝜓𝜓𝑛𝑛 (𝑥𝑥, 𝑡𝑡) 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖/ℏ 𝑑𝑑𝑑𝑑. 3.27
√2𝜋𝜋ℏ

−𝑖𝑖𝐸𝐸𝑛𝑛 𝑡𝑡/ℏ
Considering 𝜓𝜓(𝑥𝑥, 𝑡𝑡) = �𝑁𝑁 sin 𝑘𝑘𝑛𝑛 𝑥𝑥𝑒𝑒 0 < 𝑥𝑥 < 𝑎𝑎, with , 𝑁𝑁 = �𝑎𝑎,
0 elsewhere 2

1 +∞
𝜓𝜓�𝑛𝑛 (𝑝𝑝, 𝑡𝑡) = 𝑒𝑒 −𝑖𝑖𝐸𝐸𝑛𝑛 𝑡𝑡/ℏ ∫−∞ 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖/ℏ sin 𝑘𝑘𝑛𝑛 𝑥𝑥 𝑑𝑑𝑑𝑑 , 3.28
√𝜋𝜋ℏ𝑎𝑎

and the integral can be easily evaluated using

𝑒𝑒 𝑖𝑖𝑘𝑘𝑛𝑛 𝑥𝑥 −𝑒𝑒 −𝑖𝑖𝑘𝑘𝑛𝑛 𝑥𝑥


sin 𝑘𝑘𝑛𝑛 𝑥𝑥 = . 3.29
2𝑖𝑖

In the figure below, position and momentum probabilities densities for a particle confined in a
region 0 < 𝑥𝑥 < 𝑎𝑎 are shown. Two possible states are shown, the ground state with 𝑛𝑛 = 1, and the
second excited state with 𝑛𝑛 = 3. The position probability density shown on the left of the figure
shows that the most probable position is 𝑥𝑥 = 𝑎𝑎/2 when 𝑛𝑛 = 1 and when 𝑛𝑛 = 3 there are three
most probable positions: at 𝑥𝑥 = 𝑎𝑎/6, 𝑥𝑥 = 𝑎𝑎/2, and 𝑥𝑥 = 5𝑎𝑎/6. The momentum probability densities
shown on the right of the figure indicate that the most probable momentum is zero when 𝑛𝑛 = 1 and
that there two most likely momenta at 𝑛𝑛 = 3; one at 𝑝𝑝 = +3𝜋𝜋ℏ/𝑎𝑎 and one at 𝑝𝑝 = −3𝜋𝜋ℏ/𝑎𝑎. In fact,
a state with a high value of 𝑛𝑛 can be thought as a particle trapped between 𝑥𝑥 = 0 and 𝑥𝑥 = 𝑎𝑎, with
two possible momenta 𝑝𝑝 = 𝑛𝑛𝑛𝑛ℏ/𝑎𝑎 and 𝑝𝑝 = −𝑛𝑛𝑛𝑛ℏ/𝑎𝑎.
Figure 3.1 : The position and momentum probability densities for a particle confined in a
box region; 𝟎𝟎 ≤ 𝒙𝒙 ≤ 𝒂𝒂. The particles wavefunction defined by equation 3.24 is normalised
with 𝒏𝒏 = 𝟏𝟏 and 𝒏𝒏 = 𝟑𝟑.
Particle in a Box II

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

0 0 ≤ 𝑥𝑥 ≤ 𝑎𝑎
𝑉𝑉(𝑥𝑥) = , 4.2.1
∞ elsewhere

This infinite potential well confines a particle in a one dimensional box of size 𝑎𝑎. 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 𝑑𝑑 2
𝑖𝑖ℏ Ψ(𝑥𝑥, 𝑡𝑡) = �− 2𝑚𝑚 𝑑𝑑𝑑𝑑 2 + 𝑉𝑉(𝑥𝑥)� Ψ(𝑥𝑥, 𝑡𝑡). 4.2.2
𝜕𝜕𝜕𝜕

If the particle has a definite energy 𝐸𝐸𝑛𝑛 then the wavefunction has the form
𝐸𝐸𝑛𝑛
Ψ(𝑥𝑥, 𝑡𝑡) = 𝜓𝜓(𝑥𝑥)𝑒𝑒 −𝑖𝑖 ℏ 𝑡𝑡 , 4.2.3

where the wavefunction 𝜓𝜓(𝑥𝑥) satisfy the energy eigenvalue equation

ℏ2 𝑑𝑑 2
𝐸𝐸𝐸𝐸(𝑥𝑥) = �− 2𝑚𝑚 𝑑𝑑𝑑𝑑 2 + 𝑉𝑉(𝑥𝑥)� 𝜓𝜓(𝑥𝑥), 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
𝑥𝑥 = 0 and 𝑥𝑥 = 𝑎𝑎, particle is strictly confined in the region 0 ≤ 𝑥𝑥 ≤ 𝑎𝑎, beyond this region the
eigenfunction 𝜓𝜓(𝑥𝑥) is zero! Inside this region, the potential energy is zero, so the eigenfunction is a
solution of equation 4.2.4 with 𝑉𝑉(𝑥𝑥) = 0. Let us simplify this equation by writing the energy as

ℏ2 𝑘𝑘 2
𝐸𝐸 = 2𝑚𝑚
, 4.2.5

so that equation 4.2.4 can now appear as

𝑑𝑑 2
𝑑𝑑𝑑𝑑 2
𝜓𝜓(𝑥𝑥) = −𝑘𝑘 2 𝜓𝜓(𝑥𝑥). 4.2.6

Physically acceptable solutions to this partial differential equation are obtained from the general
solution 𝜓𝜓(𝑥𝑥) = 𝑀𝑀 cos 𝑘𝑘𝑘𝑘 + 𝑁𝑁 sin 𝑘𝑘𝑘𝑘,

where 𝑀𝑀 and 𝑁𝑁 are constants. The boundary conditions

𝜓𝜓(0) = 𝜓𝜓(𝑎𝑎) = 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


𝑛𝑛𝑛𝑛
𝜓𝜓𝑛𝑛 (𝑥𝑥) = 𝑁𝑁 sin 𝑘𝑘𝑛𝑛 𝑥𝑥, and 𝑘𝑘𝑛𝑛 = , where 𝑁𝑁 is an arbitrary constant, and they illustrated in Fig 4.2.2
𝑎𝑎
below.
In classical physics, eigenfunctions 𝜓𝜓𝑛𝑛 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 𝐸𝐸𝑛𝑛 , associated with quantum numbers 𝑛𝑛 = 1, 2, 3, … .

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

𝑛𝑛2 𝜋𝜋2 ℏ2
𝐸𝐸𝑛𝑛 = 2𝑚𝑚𝑎𝑎 2
, with 𝑛𝑛 = 1, 2, 3, …, 4.2.8

And that a particle with energy 𝐸𝐸𝑛𝑛 , has a wavefunction of the form
𝐸𝐸𝑛𝑛
Ψ(𝑥𝑥, 𝑡𝑡) = 𝑁𝑁 sin 𝑘𝑘𝑛𝑛 𝑥𝑥 𝑒𝑒 −𝑖𝑖 ℏ 𝑡𝑡 . 4.2.9

Note the following:


𝐸𝐸𝑛𝑛+1 −𝐸𝐸𝑛𝑛 2
 The separation between energy levels 𝐸𝐸𝑛𝑛
⟶ increases with increasing quantum
𝑛𝑛
number 𝑛𝑛.
𝜋𝜋2 ℏ2
 The minimum energy state is not zero but some minimum value 𝐸𝐸1 = 2𝑚𝑚𝑎𝑎2.
 The spacial shape of the wavefunction of a particle in box with energy 𝐸𝐸 is identical to the
spacial shape of the normal modes of a vibrating string with angular frequency 𝜔𝜔.
 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 𝑚𝑚 in a one dimensional potential defined by

−∞ if − ∞ ≤ 𝑥𝑥 ≤ 0
𝑉𝑉(𝑥𝑥) = � −𝑉𝑉0 if 0 ≤ 𝑥𝑥 ≤ 𝑎𝑎 .
0 if 𝑎𝑎 ≤ 𝑥𝑥 ≤ +∞

As shown in Fig 5.1, the potential energy changes abruptly at 𝑥𝑥 = 0 and 𝑥𝑥 = 𝑎𝑎. There is a potential
well of depth 𝑉𝑉0 , which may or may not trap the particle, and an infinite wall at 𝑥𝑥 = 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:
𝑝𝑝2
𝐸𝐸 = 𝐾𝐾 + 𝑉𝑉(𝑥𝑥) = 2𝑚𝑚 + 𝑉𝑉(𝑥𝑥). 5.1

The particle is said to be bound or trapped in a well of depth −𝑉𝑉0 , when its energy is negative in the
region −𝑉𝑉0 ≤ 𝐸𝐸 ≤ 0. Here, the particle bounces back and forth between 𝑥𝑥 = 0 and 𝑥𝑥 = 𝑎𝑎, with
kinetic given by 𝐾𝐾 = 𝐸𝐸 + 𝑉𝑉0 . But when the energy is positive, the particle is unbound. For example, it
can approach the well from 𝑥𝑥 = +∞ with kinetic energy 𝐾𝐾 = 𝐸𝐸, at 𝑥𝑥 = 𝑎𝑎 kinetic energy rises to
𝐾𝐾 = 𝐸𝐸 + 𝑉𝑉0 , it hits the infinitely high potential wall at 𝑥𝑥 = 0, then bounces back to 𝑥𝑥 = +∞.

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

𝜕𝜕 ℏ2 𝑑𝑑 2
𝑖𝑖ℏ Ψ(𝑥𝑥, 𝑡𝑡) = �− + 𝑉𝑉(𝑥𝑥)� Ψ(𝑥𝑥, 𝑡𝑡). 5.2
𝜕𝜕𝜕𝜕 2𝑚𝑚 𝑑𝑑𝑑𝑑 2

If the particle has definite energy E, the wavefunction takes the form
𝐸𝐸� 𝑡𝑡
Ψ(𝑥𝑥, 𝑡𝑡) = 𝜓𝜓(𝑥𝑥)𝑒𝑒 −𝑖𝑖 ℏ , 5.3

where 𝜓𝜓(𝑥𝑥) is the eigenfunction which satisfy the energy eigenvalue equation

ℏ2 𝑑𝑑 2
− 2𝑚𝑚 𝑑𝑑𝑑𝑑 2 𝜓𝜓(𝑥𝑥) + 𝑉𝑉(𝑥𝑥)𝜓𝜓(𝑥𝑥) = 𝐸𝐸𝐸𝐸(𝑥𝑥). 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 𝐸𝐸 = −𝑉𝑉0 and 𝐸𝐸 = 0. Let 𝐸𝐸 =
−𝜖𝜖, where 𝜖𝜖 is the particles binding energy, and seek the solutions to equation 5.4 above.

• In the region −∞ ≤ 𝑥𝑥 ≤ 0
The potential energy is infinite, so the only possible solution to equation 5.4 is 𝜓𝜓(𝑥𝑥) = 0,
which means the particle will never be found in the negative 𝑥𝑥 direction.
• In the region 0 ≤ 𝑥𝑥 ≤ 𝑎𝑎
The potential energy is 𝑉𝑉(𝑥𝑥) = −𝑉𝑉0 and equation 5.4 has the form
𝑑𝑑 2 ℏ2 𝑘𝑘02
𝜓𝜓(𝑥𝑥) = −𝑘𝑘02 𝜓𝜓(𝑥𝑥), with 𝐸𝐸 = − 𝑉𝑉0 . 5.5
𝑑𝑑𝑑𝑑 2 2𝑚𝑚
The general solution to this equation is
𝜓𝜓(𝑥𝑥) = 𝐶𝐶 sin(𝑘𝑘0 𝑥𝑥 + 𝛾𝛾),
where 𝐶𝐶 and 𝛾𝛾 are arbitrary constants. To ensure continuity of 𝜓𝜓(𝑥𝑥) at 𝑥𝑥 = 0, 𝛾𝛾 is set to
zero so that
𝜓𝜓(𝑥𝑥) = 𝐶𝐶 sin 𝑘𝑘0 𝑥𝑥. 5.6

• In the region 𝑎𝑎 ≥ 𝑥𝑥 ≥ +∞
The potential energy is zero so equation 5.4 takes the form
𝑑𝑑 2 𝛼𝛼 2 ℏ2
𝑑𝑑𝑑𝑑 2
𝜓𝜓(𝑥𝑥) = 𝛼𝛼 2 𝜓𝜓(𝑥𝑥), with 𝐸𝐸 = − 2𝑚𝑚
. 5.7
The general solution to equation 5.7 is
𝜓𝜓(𝑥𝑥) = 𝐴𝐴𝑒𝑒 −𝛼𝛼𝛼𝛼 + 𝐴𝐴′𝑒𝑒 +𝛼𝛼𝛼𝛼 ,
where 𝐴𝐴 and 𝐴𝐴′ are arbitrary constants. To ensure that the eigenfunction is finite at infinity,
𝐴𝐴′ is set to zero to give a solution which falls off exponentially with 𝑥𝑥:
𝜓𝜓(𝑥𝑥) = 𝐴𝐴𝑒𝑒 −𝛼𝛼𝛼𝛼 . 5.8

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

𝐶𝐶 sin 𝑘𝑘0 𝑎𝑎 = 𝐴𝐴𝑒𝑒 −𝛼𝛼𝛼𝛼 , 5.9


𝑑𝑑
and continuity of 𝜓𝜓 gives
𝑑𝑑𝑑𝑑

𝑘𝑘0 𝐶𝐶 cos 𝑘𝑘0 𝑎𝑎 = −𝛼𝛼𝛼𝛼𝑒𝑒 −𝛼𝛼𝛼𝛼 . 5.10

If equation 5.10 is divided by equation 5.9 results in

𝑘𝑘0 cot 𝑘𝑘0 𝑎𝑎 = −𝛼𝛼. 5.11

Equation 5.11 sets a condition for a smooth join at 𝑥𝑥 = 𝑎𝑎 for the eigenfunctions 𝐶𝐶 sin 𝑘𝑘0 𝑥𝑥 and
𝐴𝐴𝑒𝑒 −𝛼𝛼𝛼𝛼 . This is a non-trivial condition which is satisfied only when 𝑘𝑘0 and 𝛼𝛼 takes on special values.
Having obtained these special values, the binding energy for the bound states can be calculated from
𝛼𝛼 2 ℏ2
the equation 𝜖𝜖 = .
2𝑚𝑚

In order to find the binding energies, it must be noted that 𝑘𝑘0 and 𝛼𝛼 are not independent
parameters. These are defined by

ℏ2 𝑘𝑘02 𝛼𝛼 2 ℏ2
𝐸𝐸 = 2𝑚𝑚
− 𝑉𝑉0 , and 𝐸𝐸 = − 2𝑚𝑚
,

which imply that

𝑤𝑤 2 ℏ2
𝛼𝛼 2 + 𝑘𝑘02 = 𝑤𝑤 2 , where 𝑤𝑤 is given by 𝑉𝑉0 = 2𝑚𝑚
. 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 𝑘𝑘0 cot 𝑘𝑘0 𝑎𝑎 = −𝛼𝛼 and 𝛼𝛼 2 + 𝑘𝑘02 = 𝑤𝑤 2 .
The solution is illustrated in Fig 5.2 below.
Figure 5. 1: The behaviour of simultaneous functions 𝒌𝒌𝟎𝟎 𝐜𝐜𝐜𝐜𝐜𝐜 𝒌𝒌𝟎𝟎 𝒂𝒂 = −𝜶𝜶 and 𝜶𝜶𝟐𝟐 + 𝒌𝒌𝟐𝟐𝟎𝟎 = 𝒘𝒘𝟐𝟐 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
𝜋𝜋
𝑤𝑤 < ,
2𝑎𝑎

there is one bound states when


𝜋𝜋 3𝜋𝜋
< 𝑤𝑤 < ,
2𝑎𝑎 2𝑎𝑎

there are two bound states when


3𝜋𝜋 5𝜋𝜋
< 𝑤𝑤 < ,
2𝑎𝑎 2𝑎𝑎

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

2𝜋𝜋2 ℏ2
𝑉𝑉0 = 𝑚𝑚𝑎𝑎 2
.

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

𝜋𝜋2 ℏ2 𝜋𝜋2 ℏ2
𝜖𝜖1 = 3.26 , and 𝜖𝜖2 = 1.17 .
2𝑚𝑚𝑎𝑎 2 2𝑚𝑚𝑎𝑎 2

Potential barrier

Consider a barrier potential field in one dimension defined by

0 if − ∞ ≤ 𝑥𝑥 ≤ 0
𝑉𝑉(𝑥𝑥) = 𝑉𝑉0 if 0 ≤ 𝑥𝑥 ≤ 𝑎𝑎, 5.13
0 if 𝑎𝑎 ≤ 𝑥𝑥 ≤ +∞

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 𝑥𝑥 = 0 if its energy is less
than the potential step 𝑉𝑉0 , 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
𝑖𝑖ℏ Ψ(𝑥𝑥, 𝑡𝑡) = �− 2𝑚𝑚 𝜕𝜕𝜕𝜕 2 + 𝑉𝑉(𝑥𝑥)� Ψ(𝑥𝑥, 𝑡𝑡),
𝜕𝜕𝜕𝜕
In order to describe an uncertain encounter with a barrier, a wavefunction Ψ(𝑥𝑥, 𝑡𝑡) 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(𝑥𝑥, 𝑡𝑡) 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 𝑉𝑉(𝑥𝑥), 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
𝐸𝐸𝐸𝐸(𝑥𝑥) = �− 2𝑚𝑚 𝜕𝜕𝜕𝜕 2 + 𝑉𝑉(𝑥𝑥)� 𝜓𝜓(𝑥𝑥). 5.14

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 𝑥𝑥 at different regions are first sorted out, then joined smoothly at the boundaries
𝑥𝑥 = 0 and 𝑥𝑥 = 𝑎𝑎. On the left hand side of the barrier, the potential 𝑉𝑉(𝑥𝑥) = 0, the eigenfunction
𝜓𝜓(𝑥𝑥) satisfies the differential equation

𝑑𝑑 2 ℏ2 𝑘𝑘 2
𝜓𝜓(𝑥𝑥) = −𝑘𝑘 2 𝜓𝜓(𝑥𝑥), with 𝐸𝐸 = . 5.15
𝑑𝑑𝑑𝑑 2 2𝑚𝑚

The solution representing an incident wave of intensity |𝐴𝐴𝐼𝐼 |2 and reflected wave of intensity |𝐴𝐴𝑅𝑅 |2 is

𝜓𝜓(𝑥𝑥) = 𝐴𝐴𝐼𝐼 𝑒𝑒 𝑖𝑖𝑖𝑖𝑖𝑖 + 𝐴𝐴𝑅𝑅 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖 . 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 𝑉𝑉0 . If 𝐸𝐸 > 𝑉𝑉0, the region (0 ≤ 𝑥𝑥 ≤ 𝑎𝑎) is classically allowed
and eigenfunctions are explained by the equation

𝑑𝑑 2 ℏ2 𝑘𝑘02
𝑑𝑑𝑑𝑑 2
𝜓𝜓(𝑥𝑥) = −𝑘𝑘02 𝜓𝜓(𝑥𝑥), with 𝐸𝐸 = 2𝑚𝑚
+ 𝑉𝑉0 . 5.17

The general solution involves two arbitrary constants and it undulates with wave number 𝑘𝑘0

𝜓𝜓(𝑥𝑥) = 𝐴𝐴𝑒𝑒 𝑖𝑖𝑘𝑘0 𝑥𝑥 + 𝐴𝐴′𝑒𝑒 −𝑖𝑖𝑘𝑘0 𝑥𝑥 . 5.18

When 𝐸𝐸 < 𝑉𝑉0, the region (0 ≤ 𝑥𝑥 ≤ 𝑎𝑎) is classically forbidden. The eigenfunctions are described by
the equation
𝑑𝑑 2 ℏ2 𝛽𝛽 2
𝑑𝑑𝑑𝑑 2
𝜓𝜓(𝑥𝑥) = 𝛽𝛽 2 𝜓𝜓(𝑥𝑥), with 𝐸𝐸 = − 2𝑚𝑚
+ 𝑉𝑉0 , 5.19

and the general solution is

𝜓𝜓(𝑥𝑥) = 𝐵𝐵𝑒𝑒 −𝛽𝛽𝛽𝛽 + 𝐵𝐵′𝑒𝑒 𝛽𝛽𝛽𝛽 . 5.20

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

𝜓𝜓(𝑥𝑥) = 𝐴𝐴 𝑇𝑇 𝑒𝑒 𝑖𝑖𝑖𝑖𝑖𝑖 . 5.21

Smooth joining of these solutions at 𝑥𝑥 = 0 and 𝑥𝑥 = 𝑎𝑎 helps in evaluation of intensities of reflection


and transmission waves. Particularly, the expressions for the ratios

|𝐴𝐴𝑅𝑅 |2 |𝐴𝐴𝑇𝑇 |2
𝑅𝑅 = |𝐴𝐴𝐼𝐼 |2
and 𝑇𝑇 = |𝐴𝐴𝐼𝐼 |2
, 5.22

can be derived.

Because the probability of finding a particle at 𝑥𝑥 is proportional to |𝜓𝜓(𝑥𝑥)|2 , these ratios are
probabilities: 𝑅𝑅 is the probability that the particle is reflected and 𝑇𝑇 is the probability that the
particle is transmitted, and sum of these probabilities is one, i.e.

𝑅𝑅 + 𝑇𝑇 = 1. 5.23

Tunnelling through the barrier

The eigenfunction of a particle with 𝐸𝐸 < 𝑉𝑉0 are given by

𝐴𝐴𝐼𝐼 𝑒𝑒 𝑖𝑖𝑖𝑖𝑖𝑖 + 𝐴𝐴𝑅𝑅 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖 − ∞ ≤ 𝑥𝑥 ≤ 0


𝜓𝜓(𝑥𝑥) = 𝐵𝐵𝑒𝑒 𝛽𝛽𝛽𝛽 + 𝐵𝐵′𝑒𝑒
−𝛽𝛽𝛽𝛽
0 ≤ 𝑥𝑥 ≤ 𝑎𝑎 , 5.24
𝐴𝐴 𝑇𝑇 𝑒𝑒 𝑖𝑖𝑖𝑖𝑖𝑖 𝑎𝑎 ≤ 𝑥𝑥 ≤ +∞

With constants 𝐴𝐴𝐼𝐼 , 𝐴𝐴𝑅𝑅 , 𝐴𝐴 𝑇𝑇 , 𝐵𝐵, and 𝐵𝐵′ having values ensuring 𝜓𝜓(𝑥𝑥) and its derivative with respect to
position are continuous at 𝑥𝑥 = 0 and 𝑥𝑥 = 𝑎𝑎.
𝑑𝑑
For the 𝜓𝜓(𝑥𝑥) and 𝜓𝜓(𝑥𝑥) to be continuous at 𝑥𝑥 = 0, it require that
𝑑𝑑𝑑𝑑

𝐴𝐴𝐼𝐼 + 𝐴𝐴𝑅𝑅 = 𝐵𝐵 + 𝐵𝐵′, 5.25(a)

𝑖𝑖𝑖𝑖𝐴𝐴𝐼𝐼 − 𝑖𝑖𝑖𝑖𝐴𝐴𝑅𝑅 = 𝛽𝛽𝛽𝛽 − 𝛽𝛽𝛽𝛽′, 5.25(b)

and at 𝑥𝑥 = 𝑎𝑎
−𝛽𝛽𝛽𝛽
𝐵𝐵𝑒𝑒 𝛽𝛽𝛽𝛽 + 𝐵𝐵′𝑒𝑒 = 𝐴𝐴𝑇𝑇 𝑒𝑒 𝑖𝑖𝑖𝑖𝑖𝑖 , 5.26(a)
−𝛽𝛽𝛽𝛽
𝛽𝛽𝛽𝛽𝑒𝑒 𝛽𝛽𝛽𝛽 − 𝛽𝛽𝛽𝛽′𝑒𝑒 = 𝑖𝑖𝑖𝑖𝑖𝑖𝑇𝑇 𝑒𝑒 𝑖𝑖𝑖𝑖𝑖𝑖 . 5.26(b)
The main aim is to find the expression for the tunnelling probability. To acquire this, 𝐴𝐴𝐼𝐼 and 𝐴𝐴 𝑇𝑇 are
written in terms of other constant so that equations at 𝑥𝑥 = 0 give

2𝑖𝑖𝑖𝑖𝐴𝐴𝐼𝐼 = −(𝛽𝛽 − 𝑖𝑖𝑖𝑖)𝐵𝐵 + (𝛽𝛽 + 𝑖𝑖𝑖𝑖)𝐵𝐵′, 5.27

and at 𝑥𝑥 = 𝑎𝑎 give

2𝛽𝛽 (𝛽𝛽+𝑖𝑖𝑖𝑖)
𝐴𝐴 𝑇𝑇 𝑒𝑒 𝑖𝑖𝑖𝑖𝑖𝑖 = 𝐵𝐵𝑒𝑒 −𝛽𝛽𝛽𝛽 and 𝐵𝐵′ = 𝐵𝐵𝑒𝑒 −2𝛽𝛽𝛽𝛽 . 5.28
(𝛽𝛽−𝑖𝑖𝑖𝑖) (𝛽𝛽−𝑖𝑖𝑖𝑖)

The algebra is further simplified by assuming that 𝑒𝑒 −2𝛽𝛽𝛽𝛽 ≪ 1, which consequently imply that 𝐵𝐵′ ≪
𝐵𝐵 so that to a good approximation

2𝑖𝑖𝑖𝑖𝐴𝐴𝐼𝐼 ≈ −(𝛽𝛽 − 𝑖𝑖𝑖𝑖)𝐵𝐵. 5.29

Combining equation 5.29 and 5.28, results in

4𝑖𝑖𝑖𝑖𝑖𝑖
𝐴𝐴 𝑇𝑇 𝑒𝑒 𝑖𝑖𝑖𝑖𝑖𝑖 ≈ 𝐴𝐴 𝑒𝑒 −𝛽𝛽𝛽𝛽 . 5.30
(𝛽𝛽−𝑖𝑖𝑖𝑖)2 𝐼𝐼

This implies that the transmission probability is approximated to

16𝑘𝑘 2 𝛽𝛽 2
𝑇𝑇 ≈ (𝛽𝛽2 +𝑘𝑘2 )2 𝑒𝑒 −2𝛽𝛽𝛽𝛽 . 5.31

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 𝑘𝑘. When
a mass 𝑚𝑚 is displaced from equilibrium position by a distance 𝑥𝑥, there is a restoring force 𝐹𝐹 = −𝑘𝑘𝑘𝑘
which oppose this displacement. Because the work needed to move the mass from 𝑥𝑥 to 𝑥𝑥 + 𝑑𝑑𝑑𝑑 is
𝑘𝑘𝑘𝑘𝑘𝑘𝑘𝑘, the potential energy stored by displacing a mass 𝑚𝑚 is
𝑥𝑥 1
𝑉𝑉(𝑥𝑥) = ∫0 𝑘𝑘𝑘𝑘′𝑑𝑑𝑑𝑑′ = 𝑘𝑘𝑥𝑥 2 . 6.1
2

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

𝑑𝑑 2 𝑥𝑥
𝑚𝑚 = −𝑘𝑘𝑘𝑘 6.2
𝑑𝑑𝑡𝑡 2

which is usually written as

𝑥𝑥̈ = −𝜔𝜔2 𝑥𝑥, 6.3

where 𝑥𝑥̈ is the acceleration of mass 𝑚𝑚 and 𝜔𝜔 = �𝑘𝑘/𝑚𝑚. The motions general solution is

𝑥𝑥 = 𝐴𝐴 cos(𝜔𝜔𝜔𝜔 + 𝛿𝛿), 6.4

where 𝐴𝐴 and 𝛿𝛿 are two constants which may be determined by specifying the initial position and
velocity of mass 𝑚𝑚; for example if the mass is released from rest at position 𝑥𝑥0 , then 𝐴𝐴 = 𝑥𝑥0 and
𝛿𝛿 = 0.

Equation 6.4 describes a simple harmonic motion along the x-direction with amplitude 𝐴𝐴, phase 𝛿𝛿
and frequency 𝜔𝜔 or period 2𝜋𝜋/𝜔𝜔. In the cause of the motion, the potential energy rises and falls as
the kinetic energy falls and rises. But the total energy 𝐸𝐸, of which is the sum of the kinetic and
potential energy remains constant and equal to
1 1 1
𝐸𝐸 = 𝑚𝑚𝑥𝑥̇ 2 + 𝑘𝑘𝑥𝑥 2 = 𝑚𝑚𝜔𝜔2 𝐴𝐴2. 6.5
2 2 2

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
� = 𝑝𝑝� + 1 𝑚𝑚𝜔𝜔2 𝑥𝑥� 2 ,
𝐻𝐻 6.6
2𝑚𝑚 2

or if position and momentum operators are used


2 2
� = − ℏ 𝜕𝜕 2 + 1 𝑚𝑚𝜔𝜔2 𝑥𝑥 2.
𝐻𝐻 6.7
2𝑚𝑚 𝜕𝜕𝜕𝜕 2

The first term represent the kinetic energy of a particle with mass 𝑚𝑚, 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 𝜔𝜔. 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 𝜓𝜓(𝑥𝑥, 𝑡𝑡) which
satisfy the Schrӧdinger equation

𝜕𝜕
𝑖𝑖ℏ 𝜓𝜓(𝑥𝑥, 𝑡𝑡) � 𝜓𝜓(𝑥𝑥, 𝑡𝑡),
= 𝐻𝐻 6.8
𝜕𝜕𝜕𝜕

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

Ψ(𝑥𝑥, 𝑡𝑡) = 𝜓𝜓(𝑥𝑥)𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖/ℏ , 6.9

where 𝜓𝜓(𝑥𝑥) is an eigenfunction belong to an energy eigenvalue 𝐸𝐸. If this wavefunction is used
together with the Hamiltonian operator defined above, an energy eigenvalue equation becomes

ℏ2 𝑑𝑑 2 1
�− 2𝑚𝑚 𝑑𝑑𝑑𝑑2 + 2 𝑚𝑚𝜔𝜔2 𝑥𝑥 2 � 𝜓𝜓(𝑥𝑥) = 𝐸𝐸𝐸𝐸(𝑥𝑥). 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.

𝜓𝜓(𝑥𝑥) ⟶ 0 as 𝑥𝑥 ⟶ ±∞. 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 𝐸𝐸𝑛𝑛 , and
the corresponding eigenfunctions by 𝜓𝜓𝑛𝑛 (𝑥𝑥), where 𝑛𝑛 is a quantum number. The convention of
labelling the ground state by 𝑛𝑛 = 0, and the subsequent first, second, and third excited states by
𝑛𝑛 = 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 𝑟𝑟0 denotes
the separation at which the potential has a minimum, and 𝑥𝑥 = 𝑟𝑟 − 𝑟𝑟0 denotes a small displacement
from 𝑟𝑟0 then
1
𝑉𝑉𝑒𝑒 (𝑟𝑟) ≈ 𝑘𝑘𝑥𝑥 2 , 6.12
2

where 𝑘𝑘 is a constant. This equation implies that if nuclei are displaced a distance 𝑥𝑥 from their
equilibrium separation 𝑟𝑟0 , there is a restoring force of magnitude 𝑘𝑘𝑘𝑘, and the potential energy
1
increases by 2 𝑘𝑘𝑥𝑥 2 . The constant 𝑘𝑘 is the effective elastic constant which characterises the strength
of molecular bond between nuclei in the molecule. If classical physics was applicable, the nuclei
would have energy

𝑝𝑝2 𝑝𝑝2 1
𝐸𝐸classical = 2𝑚𝑚1 + 2𝑚𝑚2 + 𝑘𝑘𝑥𝑥 2 , 6.13
1 2 2

where 𝑚𝑚1 and 𝑚𝑚2 are masses of the nuclei, and 𝑝𝑝1 and 𝑝𝑝2 are the magnitudes of their momenta. In
the centre of mass frame, 𝑝𝑝1 = 𝑝𝑝2 = 𝑝𝑝, and the reduced mass is
𝑚𝑚 𝑚𝑚
𝜇𝜇 = 𝑚𝑚 1+𝑚𝑚2 , 6.14
1 2

so that the classical total energy becomes

𝑝𝑝2 1
𝐸𝐸classical = 2𝜇𝜇 + 𝑘𝑘𝑥𝑥 2 . 6.15
2

This is the same as the energy of a single particle of mass 𝜇𝜇 on a spring with elastic constant 𝑘𝑘.
Accordingly, the vibrating nuclei in a diatomic molecule act like a harmonic oscillator with classical
frequency 𝜔𝜔 = �𝑘𝑘/𝜇𝜇, where 𝜇𝜇 is the reduced mass of the nuclei and 𝑘𝑘 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 𝜓𝜓(𝑥𝑥, 𝑡𝑡) which
satisfies the Schrӧdinger equation
𝜕𝜕 ℏ2 𝜕𝜕2 1
𝑖𝑖ℏ 𝜓𝜓 = �− 2𝜇𝜇 𝜕𝜕𝜕𝜕 2 + 𝑘𝑘𝑥𝑥 2 � 𝜓𝜓. 6.16
𝜕𝜕𝜕𝜕 2

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

1
𝐸𝐸𝑛𝑛 = �𝑛𝑛 + � ℏ𝜔𝜔 with 𝑛𝑛 = 0, 1, 2, 3, … 6.17
2

As shown below, the energy levels have equal spacing ℏ𝜔𝜔 and lowest energy level is
1
𝐸𝐸0 = ℏ𝜔𝜔. 6.18
2

When 𝑛𝑛 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
𝑘𝑘
absorption of photons with energy 𝐸𝐸 = ℏ� . These photons give rise to spectral lines having
𝜇𝜇

ℎ𝑐𝑐 𝜇𝜇
wavelength 𝜆𝜆 = = 2𝜋𝜋𝜋𝜋� .
𝐸𝐸 𝑘𝑘

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

𝜇𝜇
Substituting these values in 𝜆𝜆 = 2𝜋𝜋𝜋𝜋� we obtain
𝑘𝑘

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

3-D Oscillators

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


1 1
𝑉𝑉(𝑟𝑟) = 𝑘𝑘𝑟𝑟 2 = 𝑘𝑘(𝑥𝑥 2 + 𝑦𝑦 2 + 𝑧𝑧 2 ). 6.19
2 2

A classical particle at a distance 𝑟𝑟 from the origin will experience a central force of magnitude 𝑘𝑘𝑘𝑘.
When displaced from the origin and released it executes a simple harmonic motion with angular
frequency 𝜔𝜔 = �𝑘𝑘/𝑚𝑚. 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

� = 𝐻𝐻
𝐻𝐻 �𝑦𝑦 + 𝐻𝐻
�𝑥𝑥 + 𝐻𝐻 �𝑧𝑧 , where

ℏ2 𝜕𝜕2 1
�𝑥𝑥 = −
𝐻𝐻 + 𝑚𝑚𝜔𝜔2 𝑥𝑥 2 , 6.20(a)
2𝑚𝑚 𝜕𝜕𝜕𝜕 2 2

2 2
�𝑦𝑦 = − ℏ 𝜕𝜕 2 + 1 𝑚𝑚𝜔𝜔2 𝑦𝑦 2 ,
𝐻𝐻 6.20(b)
2𝑚𝑚 𝜕𝜕𝜕𝜕 2

2 2
�𝑧𝑧 = − ℏ 𝜕𝜕 2 + 1 𝑚𝑚𝜔𝜔2 𝑧𝑧 2.
𝐻𝐻 6.20(c)
2𝑚𝑚 𝜕𝜕𝜕𝜕 2

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

Ψ(𝑥𝑥, 𝑦𝑦, 𝑧𝑧, 𝑡𝑡) = 𝜓𝜓(𝑥𝑥, 𝑦𝑦, 𝑧𝑧)𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖/ℏ, 6.21

where 𝜓𝜓(𝑥𝑥, 𝑦𝑦, 𝑧𝑧) and 𝐸𝐸 satisfy the three dimensional eigenvalue equation

� 𝜓𝜓(𝑥𝑥, 𝑦𝑦, 𝑧𝑧) = 𝐸𝐸𝐸𝐸(𝑥𝑥, 𝑦𝑦, 𝑧𝑧).


𝐻𝐻 6.22

�𝑥𝑥 , 𝐻𝐻
These states can be found by using the one dimensional eigenvalue equations according to 𝐻𝐻 �𝑦𝑦 ,
�𝑧𝑧 as
and 𝐻𝐻

�𝑥𝑥 𝜓𝜓𝑛𝑛 (𝑥𝑥) = (𝑛𝑛𝑥𝑥 + 1)ℏ𝜔𝜔𝜓𝜓𝑛𝑛 (𝑥𝑥),


𝐻𝐻 6.23(a)
𝑥𝑥 𝑥𝑥
2

�𝑦𝑦 𝜓𝜓𝑛𝑛 (𝑦𝑦) = (𝑛𝑛𝑦𝑦 + 1)ℏ𝜔𝜔𝜓𝜓𝑛𝑛 (𝑦𝑦),


𝐻𝐻 6.23(b)
𝑦𝑦 𝑦𝑦
2

�𝑧𝑧 𝜓𝜓𝑛𝑛 (𝑧𝑧) = (𝑛𝑛𝑧𝑧 + 1)ℏ𝜔𝜔𝜓𝜓𝑛𝑛 (𝑧𝑧),


𝐻𝐻 6.23(c)
𝑧𝑧 𝑧𝑧
2

where the quantum numbers 𝑛𝑛𝑥𝑥 , 𝑛𝑛𝑦𝑦 , 𝑛𝑛𝑧𝑧 can take values 𝑛𝑛 = 0, 1, 2, …

These three equations imply that the wave function 𝜓𝜓𝑛𝑛𝑥𝑥 ,𝑛𝑛𝑦𝑦 ,𝑛𝑛𝑧𝑧 (𝑥𝑥, 𝑦𝑦, 𝑧𝑧) = 𝜓𝜓𝑛𝑛𝑥𝑥 (𝑥𝑥)𝜓𝜓𝑛𝑛𝑦𝑦 (𝑦𝑦)𝜓𝜓𝑛𝑛𝑧𝑧 (𝑧𝑧)
satisfy the three dimensional eigenvalue equation

� 𝜓𝜓𝑛𝑛 ,𝑛𝑛 ,𝑛𝑛 (𝑥𝑥, 𝑦𝑦, 𝑧𝑧) = 𝐸𝐸𝑛𝑛 ,𝑛𝑛 ,𝑛𝑛 𝜓𝜓𝑛𝑛 ,𝑛𝑛 ,𝑛𝑛 (𝑥𝑥, 𝑦𝑦, 𝑧𝑧)
𝐻𝐻 6.24
𝑥𝑥 𝑦𝑦 𝑧𝑧 𝑥𝑥 𝑦𝑦 𝑧𝑧 𝑥𝑥 𝑦𝑦 𝑧𝑧

provided that
3
𝐸𝐸𝑛𝑛𝑥𝑥 ,𝑛𝑛𝑦𝑦 ,𝑛𝑛𝑧𝑧 = (𝑛𝑛𝑥𝑥 + 𝑛𝑛𝑦𝑦 + 𝑛𝑛𝑧𝑧 + )ℏ𝜔𝜔. 6.25
2

Thus the eigenvalues and eigenfunctions of a three dimensional oscillator are labelled by three
quantum numbers 𝑛𝑛𝑥𝑥 , 𝑛𝑛𝑦𝑦 , and 𝑛𝑛𝑧𝑧 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 +𝑦𝑦 2 +𝑧𝑧 2 )/2𝑎𝑎 2


𝐸𝐸0,0,0 = ℏ𝜔𝜔 and 𝜓𝜓0,0,0 (𝑥𝑥, 𝑦𝑦, 𝑧𝑧) = � � 𝑒𝑒 −(𝑥𝑥 , 6.26
2 𝑎𝑎√𝜋𝜋

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

5 1 3/2 𝑥𝑥 2 +𝑦𝑦 2 +𝑧𝑧 2 )/2𝑎𝑎 2


𝐸𝐸1,0,0 = ℏ𝜔𝜔 and 𝜓𝜓1,0,0 (𝑥𝑥, 𝑦𝑦, 𝑧𝑧) = � � 21/2 � � 𝑒𝑒 −(𝑥𝑥 ; 6.27(a)
2 𝑎𝑎√𝜋𝜋 𝑎𝑎

5 1 3/2 1/2 𝑦𝑦 2 2 2 2
𝐸𝐸0,1,0 = ℏ𝜔𝜔 and 𝜓𝜓0,1,0 (𝑥𝑥, 𝑦𝑦, 𝑧𝑧) = � � 2 � � 𝑒𝑒 −(𝑥𝑥 +𝑦𝑦 +𝑧𝑧 )/2𝑎𝑎 ; 6.27(b)
2 𝑎𝑎√𝜋𝜋 𝑎𝑎

5 1 3/2 𝑧𝑧 2 +𝑦𝑦 2 +𝑧𝑧 2 )/2𝑎𝑎2


𝐸𝐸0,0,1 = ℏ𝜔𝜔 and 𝜓𝜓0,0,1 (𝑥𝑥, 𝑦𝑦, 𝑧𝑧) = � � 21/2 � � 𝑒𝑒 −(𝑥𝑥 . 6.27(c)
2 𝑎𝑎√𝜋𝜋 𝑎𝑎
Angular Momentum

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

𝑟𝑟⃗ = (𝑥𝑥, 𝑦𝑦, 𝑧𝑧) and 𝑝𝑝⃗ = (𝑝𝑝𝑥𝑥 , 𝑝𝑝𝑦𝑦 , 𝑝𝑝𝑧𝑧 ). 7.1

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

𝑙𝑙⃗ = 𝑟𝑟⃗ × 𝑝𝑝⃗. 7.2

This vector has the following three Cartesian components

𝑙𝑙𝑥𝑥 = 𝑦𝑦𝑝𝑝𝑧𝑧 − 𝑧𝑧𝑝𝑝𝑦𝑦 , 𝑙𝑙𝑦𝑦 = 𝑧𝑧𝑝𝑝𝑥𝑥 − 𝑥𝑥𝑝𝑝𝑧𝑧 , 𝑙𝑙𝑧𝑧 = 𝑥𝑥𝑝𝑝𝑦𝑦 − 𝑦𝑦𝑝𝑝𝑥𝑥 7.3

with the magnitude

�𝑙𝑙⃗� = �𝑙𝑙𝑥𝑥2 + 𝑙𝑙𝑦𝑦2 + 𝑙𝑙𝑧𝑧2 . 7.4

Quantum mechanically the angular momentum is described by the operator

𝑙𝑙̂ = 𝑟𝑟̂ × 𝑝𝑝̂ = −𝑖𝑖ℏ𝑟𝑟⃗ × �∇⃗ 7.5

this is a vector operator with three Cartesian components

𝜕𝜕 𝜕𝜕 𝜕𝜕 𝜕𝜕 𝜕𝜕 𝜕𝜕
𝑙𝑙̂𝑥𝑥 = −𝑖𝑖ℏ �𝑦𝑦 − 𝑧𝑧 �, 𝑙𝑙̂𝑦𝑦 = −𝑖𝑖ℏ �𝑧𝑧 − 𝑥𝑥 �, 𝑙𝑙̂𝑧𝑧 = −𝑖𝑖ℏ �𝑥𝑥 − 𝑦𝑦 � 7.6
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕

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

The spin angular momentum is usually described by an operator 𝑠𝑠̂ = (𝑠𝑠̂𝑥𝑥 , 𝑠𝑠̂𝑦𝑦 , 𝑠𝑠̂𝑧𝑧 ) 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 𝐿𝐿 about an axis.

The moment of inertia tensor 𝑰𝑰 and angular velocity 𝜔𝜔 (𝐿𝐿 is not always parallel to 𝜔𝜔).

The total angular momentum (spin plus orbital) is 𝑱𝑱. 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 𝑥𝑥�, 𝑦𝑦�, or 𝑧𝑧̂
direction:

if you measure … you get …


𝑳𝑳�𝒙𝒙,𝒚𝒚,𝒛𝒛 … , −2ℏ, −ℏ, 0, ℏ, 2ℏ, …
𝒔𝒔�𝒙𝒙,𝒚𝒚,𝒛𝒛or 𝑱𝑱�𝒙𝒙,𝒚𝒚,𝒛𝒛 3 1 1 3
… , − ℏ, −ℏ, − ℏ, 0, ℏ, ℏ, ℏ, …
2 2 2 2
𝑳𝑳�𝟐𝟐 (ℏ2 𝑛𝑛(𝑛𝑛 + 1)), 𝑛𝑛 = 0, 1, 2, …
𝒔𝒔�𝟐𝟐 or 𝑱𝑱�𝟐𝟐 1 3
(ℏ2 𝑛𝑛(𝑛𝑛 + 1)), 𝑛𝑛 = 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

�𝐴𝐴̂, 𝐵𝐵�� ≡ 𝐴𝐴̂𝐵𝐵� − 𝐵𝐵�𝐴𝐴̂. 7.7

In quantum mechanics operators do not commute i.e. 𝐴𝐴̂𝐵𝐵� ≠ 𝐵𝐵�𝐴𝐴̂.

Two observables 𝐴𝐴 and 𝐵𝐵 described by operators 𝐴𝐴̂ and 𝐵𝐵� are non-compatible if �𝐴𝐴̂, 𝐵𝐵�� ≠ 0 and they
are compatible if �𝐴𝐴̂, 𝐵𝐵�� = 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 𝜕𝜕
𝑥𝑥�𝑝𝑝̂ 𝜓𝜓 = 𝑥𝑥 𝜓𝜓 = 𝑥𝑥 𝜓𝜓
𝑖𝑖 𝜕𝜕𝜕𝜕 𝑖𝑖 𝜕𝜕𝜕𝜕

ℏ 𝜕𝜕 ℏ ℏ 𝜕𝜕
𝑝𝑝̂ 𝑥𝑥�𝜓𝜓 = 𝑝𝑝̂ (𝑥𝑥�𝜓𝜓) = (𝑥𝑥𝑥𝑥) = 𝑖𝑖 𝜓𝜓 + 𝑖𝑖 𝑥𝑥 𝜕𝜕𝜕𝜕 𝜓𝜓
𝑖𝑖 𝜕𝜕𝜕𝜕

(𝑥𝑥�𝑝𝑝̂ − 𝑝𝑝̂ 𝑥𝑥�)𝜓𝜓 = 𝑖𝑖ℏ𝜓𝜓

therefore [𝑥𝑥�, 𝑝𝑝̂ ] = (𝑥𝑥�𝑝𝑝̂ − 𝑝𝑝̂ 𝑥𝑥�)𝜓𝜓 = 𝑖𝑖ℏ𝜓𝜓

in a similar manner it can be shown that:

[𝑥𝑥�, 𝑝𝑝̂𝑥𝑥 ] = �𝑦𝑦�, 𝑝𝑝̂𝑦𝑦 � = [𝑧𝑧̂ , 𝑝𝑝̂𝑧𝑧 ] = 𝑖𝑖ℏ 7.8(a)

�𝑧𝑧̂ , 𝑝𝑝̂𝑦𝑦 � = [𝑥𝑥�, 𝑝𝑝̂𝑧𝑧 ] = [𝑦𝑦�, 𝑝𝑝̂𝑥𝑥 ] = 0 7.8(b)

Properties of commutators

�𝐴𝐴̂, 𝐵𝐵�� = −�𝐵𝐵�, 𝐴𝐴̂� 7.9(a)

�𝐴𝐴̂, 𝐵𝐵� + 𝐶𝐶̂ � = �𝐴𝐴̂, 𝐵𝐵�� + �𝐴𝐴̂, 𝐶𝐶̂ � 7.9(b)

�𝐴𝐴̂, 𝐵𝐵�𝐶𝐶̂ � = �𝐴𝐴̂, 𝐵𝐵��𝐶𝐶̂ + 𝐵𝐵��𝐴𝐴̂, 𝐶𝐶̂ � 7.9(c)

�𝐴𝐴̂, �𝐵𝐵�, 𝐶𝐶̂ �� = �𝐵𝐵�, �𝐶𝐶̂ , 𝐴𝐴̂�� + �𝐶𝐶̂ , �𝐴𝐴̂, 𝐵𝐵��� = 0 ……. Jacobi identity! 7.9(d)

and note: �𝐴𝐴̂, 𝐴𝐴̂� = 0 and �𝐴𝐴̂, 𝑓𝑓(𝐴𝐴̂)� = 0

likewise for angular momentum we have:

�𝑙𝑙̂𝑥𝑥 , 𝑙𝑙̂𝑦𝑦 � = 𝑖𝑖ℏ𝑙𝑙̂𝑧𝑧 ; �𝑙𝑙̂𝑦𝑦 , 𝑙𝑙̂𝑧𝑧 � = 𝑖𝑖ℏ𝑙𝑙̂𝑥𝑥 ; �𝑙𝑙̂𝑧𝑧 , 𝑙𝑙̂𝑥𝑥 � = 𝑖𝑖ℏ𝑙𝑙̂𝑦𝑦 7.10
�𝑙𝑙̂2 , 𝑙𝑙̂𝑥𝑥 � = �𝑙𝑙̂2 , 𝑙𝑙̂𝑦𝑦 � = �𝑙𝑙̂2 , 𝑙𝑙̂𝑧𝑧 � = 0 7.11

𝑳𝑳�𝟐𝟐 Operator

A new operator 𝐿𝐿�2 is introduced as this operator commutes with individual component of 𝐿𝐿, but the
same components of 𝐿𝐿 do not commute with each other.

𝐿𝐿�2 is given by 𝐿𝐿�2 = 𝐿𝐿�2𝑥𝑥 + 𝐿𝐿�2𝑦𝑦 + 𝐿𝐿�2𝑧𝑧 .

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 𝐿𝐿�𝑧𝑧 , then it is not a function of 𝐿𝐿�𝑥𝑥 and 𝐿𝐿�𝑦𝑦 .

Taking measurements of angular momentum along 𝐿𝐿�𝑧𝑧 (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 𝐿𝐿 commutes with
potential energy 𝑉𝑉(𝑟𝑟). If 𝐿𝐿 commutes with kinetic energy, then 𝐿𝐿 is a constant of motion.

� ) (kinetic and potential energy operator), then the


If 𝐿𝐿 commutes with the Hamiltonian operator (𝐻𝐻
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 – 𝑒𝑒 and one proton in a nucleus of charge +𝑒𝑒. 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

𝑒𝑒 2
𝑉𝑉(𝑟𝑟) = 4𝜋𝜋𝜖𝜖 𝑟𝑟. 2.17
0

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 𝑚𝑚, with vector position 𝑟𝑟⃗, momentum 𝑝𝑝⃗, and angular
momentum 𝐿𝐿 �⃗ = 𝑟𝑟⃗ × 𝑝𝑝⃗, about a fixed origin. If the particle moves in a central potential 𝑉𝑉(𝑟𝑟), it is
subject to a force
𝑑𝑑𝑑𝑑
𝐹𝐹⃗ = − 𝑒𝑒̂𝑟𝑟 , 8.1
𝑑𝑑𝑑𝑑

where 𝑒𝑒̂𝑟𝑟 is the unit vector in the direction of 𝑟𝑟⃗. Because the force acts along the radius vector 𝑟𝑟⃗, the
�⃗ = 𝑟𝑟⃗ × 𝐹𝐹⃗ on the particle is zero, and particle moves with constant angular momentum 𝐿𝐿
torque 𝑁𝑁 �⃗.
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 𝒅𝒅𝑨𝑨�⃗ swept by 𝒓𝒓
�⃗ and position element 𝒅𝒅𝒓𝒓
�⃗.
which shows that the vector area swept by the radius vector 𝑟𝑟⃗ is given by

�⃗
𝐿𝐿
𝑑𝑑𝐴𝐴⃗ = 2𝑚𝑚 𝑑𝑑𝑑𝑑. 8.2

This implies that, when angular momentum is a constant vector 𝐿𝐿�⃗, the particle moves in a fixed
plane with a radius vector which sweeps out an area at a constant rate of 𝐿𝐿/2𝑚𝑚. The momentum 𝑝𝑝⃗
of the particle moving in a plane has two independent components which can be conveniently called
the radial and transverse momenta
𝑑𝑑𝑑𝑑 𝐿𝐿
𝑝𝑝𝑟𝑟 = 𝑚𝑚 , and 𝑝𝑝𝑡𝑡 = . 8.3
𝑑𝑑𝑑𝑑 𝑟𝑟

𝑝𝑝2
If we write the kinetic energy 2𝑚𝑚 in terms of 𝑝𝑝𝑟𝑟 and 𝑝𝑝𝑡𝑡 the the constant total energy of the particle
is given by

𝑝𝑝2 𝐿𝐿2
𝑟𝑟
𝐸𝐸 = 2𝑚𝑚 + 2𝑚𝑚𝑚𝑚2 + 𝑉𝑉(𝑟𝑟). 8.4

𝑝𝑝2
The total energy of the particle can also be viewed as the radial kinetic term 2𝑚𝑚
𝑟𝑟
, and the effective
potential

𝐿𝐿2
𝑉𝑉𝑒𝑒𝑒𝑒𝑒𝑒 (𝑟𝑟) = 2𝑚𝑚𝑚𝑚2 + 𝑉𝑉(𝑟𝑟). 8.5

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

𝑑𝑑𝑉𝑉𝑒𝑒𝑒𝑒𝑒𝑒 𝐿𝐿2 𝑑𝑑𝑑𝑑(𝑟𝑟)


𝐹𝐹 = − 𝑑𝑑𝑑𝑑
= 𝑚𝑚𝑟𝑟3 − . 8.6
𝑑𝑑𝑑𝑑
𝐿𝐿2 2
The term 𝑚𝑚𝑟𝑟 3
, which is the same as the 𝑚𝑚𝜐𝜐 �𝑟𝑟 for a particle moving with speed 𝜐𝜐 in a circle of radius
𝐿𝐿2
𝑟𝑟 with angular momentum 𝐿𝐿 = 𝑚𝑚𝑚𝑚𝑚𝑚, is the centrifugal force. So that the term in the effective
2𝑚𝑚𝑚𝑚 2
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 𝑚𝑚 moves around the sun with gravitational potential energy
𝑚𝑚𝑀𝑀⨀
𝑉𝑉(𝑟𝑟) = −𝐺𝐺 , 8.7
𝑟𝑟

where 𝑀𝑀⨀ is the mass of the sun, and 𝐺𝐺 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 𝜓𝜓(𝑟𝑟, 𝜃𝜃, 𝜙𝜙, 𝑡𝑡). The focus will be on quantum states with quite definite energy 𝐸𝐸, whose
wave functions can be described by

Ψ(𝑟𝑟, 𝜃𝜃, 𝜙𝜙, 𝑡𝑡) = 𝜓𝜓(𝑟𝑟, 𝜃𝜃, 𝜙𝜙)𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖/ℏ , 8.8

where 𝜓𝜓(𝑟𝑟, 𝜃𝜃, 𝜙𝜙) is the energy eigenfunction which satisfy the eigenvalue equation

ℏ2
�− 2𝑚𝑚 ∇2 + 𝑉𝑉(𝑟𝑟)� 𝜓𝜓 = 𝐸𝐸𝐸𝐸. 8.9

This now is partial differential equation in three independent variables 𝑟𝑟, 𝜃𝜃, and 𝜙𝜙. It can be
simplified if in addition to the definite energy 𝐸𝐸, definite angular momentum 𝐿𝐿 is considered.
Assume that the magnitude of the orbital angular momentum is 𝐿𝐿 = �𝑙𝑙(𝑙𝑙 + 1)ℏ, and the 𝑧𝑧-
component of this orbital angular momentum is 𝐿𝐿𝑧𝑧 = 𝑚𝑚𝑙𝑙 ℏ, where 𝑙𝑙 and 𝑚𝑚𝑙𝑙 are quantum numbers
which could take on the values 𝑙𝑙 = 0, 1, 2, … and 𝑚𝑚𝑙𝑙 = −𝑙𝑙, … , +𝑙𝑙. The eigenfunctions have the form

𝜓𝜓(𝑟𝑟, 𝜃𝜃, 𝜙𝜙) = 𝑅𝑅(𝑟𝑟)𝑌𝑌𝑙𝑙,𝑚𝑚𝑙𝑙 (𝜃𝜃, 𝜙𝜙). 8.10

𝑅𝑅(𝑟𝑟) is the radial part of and 𝑌𝑌𝑙𝑙,𝑚𝑚𝑙𝑙 (𝜃𝜃, 𝜙𝜙) are the spherical harmonics wavefunction part of the
wavefunction 𝜓𝜓(𝑟𝑟, 𝜃𝜃, 𝜙𝜙).

In this equation 𝑌𝑌𝑙𝑙,𝑚𝑚𝑙𝑙 (𝜃𝜃, 𝜙𝜙) is a simultaneous eigenfunction of 𝐿𝐿�2 and 𝐿𝐿�𝑧𝑧 whose values satisfy
equations

𝐿𝐿�2 𝑌𝑌𝑙𝑙,𝑚𝑚𝑙𝑙 (𝜃𝜃, 𝜙𝜙) = 𝑙𝑙(𝑙𝑙 + 1)ℏ2 𝑌𝑌𝑙𝑙,𝑚𝑚𝑙𝑙 (𝜃𝜃, 𝜙𝜙) 8.11

and

𝐿𝐿�𝑧𝑧 𝑌𝑌𝑙𝑙,𝑚𝑚𝑙𝑙 (𝜃𝜃, 𝜙𝜙) = 𝑚𝑚𝑙𝑙 ℏ𝑌𝑌𝑙𝑙,𝑚𝑚𝑙𝑙 (𝜃𝜃, 𝜙𝜙). 8.12


𝑅𝑅(𝑟𝑟) is an unknown function of 𝑟𝑟. If we substitute 𝜓𝜓(𝑟𝑟, 𝜃𝜃, 𝜙𝜙) into the eigenvalue equation we use
identities

1 𝜕𝜕2 (𝑟𝑟𝑟𝑟) 1 𝜕𝜕2 𝜓𝜓 cos 𝜃𝜃 𝜕𝜕𝜕𝜕 1 𝜕𝜕2 𝜓𝜓


∇2 𝜓𝜓 = + � + + � 8.13
𝑟𝑟 2 𝜕𝜕𝜕𝜕 2 𝑟𝑟 2 𝜕𝜕𝜕𝜕2 sin 𝜃𝜃 𝜕𝜕𝜕𝜕 sin2 𝜃𝜃 𝜕𝜕𝜕𝜕2

and

𝜕𝜕 2
cos 𝜃𝜃 𝜕𝜕 1 𝜕𝜕2
𝐿𝐿�2 = −ℏ2 �𝜕𝜕𝜕𝜕2 + + �, 8.14
sin 𝜃𝜃 𝜕𝜕𝜕𝜕 sin 𝜃𝜃 𝜕𝜕𝜕𝜕2
2

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

ℏ2 𝑑𝑑 2 (𝑟𝑟𝑟𝑟(𝑟𝑟)) 𝑙𝑙(𝑙𝑙+1)ℏ2
− 2𝑚𝑚𝑚𝑚 𝑑𝑑𝑑𝑑 2
+� 2𝑚𝑚𝑟𝑟 2
+ 𝑉𝑉(𝑟𝑟)� 𝑅𝑅(𝑟𝑟) = 𝐸𝐸𝐸𝐸(𝑟𝑟). 8.15

By introducing a radial function defined by


𝑢𝑢(𝑟𝑟)
𝑅𝑅(𝑟𝑟) = , 8.16
𝑟𝑟

we obtain

ℏ2 𝑑𝑑 2 𝑢𝑢(𝑟𝑟) 𝑙𝑙(𝑙𝑙+1)ℏ2
− 2𝑚𝑚 +� + 𝑉𝑉(𝑟𝑟)� 𝑢𝑢(𝑟𝑟) = 𝐸𝐸𝑢𝑢(𝑟𝑟). 8.17
𝑑𝑑𝑑𝑑 2 2𝑚𝑚𝑟𝑟 2

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

𝑙𝑙(𝑙𝑙+1)ℏ2
𝑉𝑉𝑒𝑒𝑒𝑒𝑒𝑒 (𝑟𝑟) = 2𝑚𝑚𝑟𝑟 2
+ 𝑉𝑉(𝑟𝑟). 8.18

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

𝑢𝑢(𝑟𝑟) = 0 at 𝑟𝑟 = 0 8.19

𝑢𝑢(𝑟𝑟)�
must be imposed to ensure that the function 𝑅𝑅(𝑟𝑟) = 𝑟𝑟. Actually, this is in line with fact that
the 3-D eigenfunction 𝜓𝜓(𝑟𝑟, 𝜃𝜃, 𝜙𝜙) is finite at the origin. In addition, bound states which ensures that
the particle does not escape to infinity, must satisfy the boundary condition

𝑢𝑢(𝑟𝑟) ⟶ 0 as 𝑟𝑟 ⟶ ∞. 8.20

Bound states only exist if the effective potential 𝑉𝑉𝑒𝑒𝑒𝑒𝑒𝑒 (𝑟𝑟) is sufficiently attractive. These states are
labelled by the quantum number 𝑛𝑛𝑟𝑟 = 0, 1, 2, … which could be shown to be equal to number of
nodes of the eigenfunction 𝑢𝑢(𝑟𝑟) between 𝑟𝑟 = 0 and 𝑟𝑟 = ∞ . This means that a bound state of a
particle in a central potential can always be specified by three quantum numbers 𝑛𝑛𝑟𝑟 , 𝑙𝑙, and 𝑚𝑚𝑙𝑙 and
that eigenfunction has the form

𝑢𝑢𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟)


𝜓𝜓𝑛𝑛𝑟𝑟 ,𝑙𝑙,𝑚𝑚𝑙𝑙 (𝑟𝑟, 𝜃𝜃, 𝜙𝜙) = 𝑟𝑟
𝑌𝑌𝑙𝑙,𝑚𝑚𝑙𝑙 (𝜃𝜃, 𝜙𝜙). 8.21
2
By using the normalization condition ∫�𝑌𝑌𝑙𝑙,𝑚𝑚𝑙𝑙 � 𝑑𝑑Ω = 1, with 𝑑𝑑Ω = sin 𝜃𝜃 𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑 being the solid angle
of spherical harmonics it can be shown that the eigenfunction 𝜓𝜓𝑛𝑛𝑟𝑟,𝑙𝑙,𝑚𝑚𝑙𝑙 (𝑟𝑟, 𝜃𝜃, 𝜙𝜙) is normalized
provided the radial function 𝑢𝑢(𝑟𝑟) satisfy the condition

∞ 2
∫0 �𝑢𝑢𝑛𝑛𝑟𝑟,𝑙𝑙 (𝑟𝑟)� 𝑑𝑑𝑑𝑑 = 1. 8.22

The energy of these bound states will be denoted by 𝐸𝐸𝑛𝑛𝑟𝑟,𝑙𝑙 . 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 𝑛𝑛𝑟𝑟 and 𝑙𝑙 have higher energies.
The energy 𝐸𝐸𝑛𝑛𝑟𝑟 ,𝑙𝑙 increases with increasing 𝑙𝑙 because the average transverse kinetic energy is given
by

∞ 𝑙𝑙(𝑙𝑙+1)ℏ2
∫0 𝑢𝑢𝑛𝑛∗ 𝑟𝑟 ,𝑙𝑙 (𝑟𝑟) � 2𝑚𝑚𝑟𝑟 2
� 𝑢𝑢𝑛𝑛𝑟𝑟,𝑙𝑙 (𝑟𝑟)𝑑𝑑𝑑𝑑; 8.23

and the energy 𝐸𝐸𝑛𝑛𝑟𝑟 ,𝑙𝑙 also increases with increasing 𝑛𝑛𝑟𝑟 because the radial kinetic energy is given by

∞ ℏ2 𝑑𝑑 2
∫0 𝑢𝑢𝑛𝑛∗ 𝑟𝑟,𝑙𝑙 (𝑟𝑟) �− 2𝑚𝑚 𝑑𝑑𝑑𝑑 2� 𝑢𝑢𝑛𝑛𝑟𝑟,𝑙𝑙 (𝑟𝑟)𝑑𝑑𝑑𝑑, 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 𝐿𝐿 = �𝑙𝑙(𝑙𝑙 + 1)ℏ and 𝐿𝐿𝑧𝑧 = 𝑚𝑚𝑙𝑙 ℏ. These states have wave
function of the form

𝜓𝜓𝑛𝑛𝑟𝑟,𝑙𝑙,𝑚𝑚𝑙𝑙 (𝑟𝑟, 𝜃𝜃, 𝜙𝜙, 𝑡𝑡) = 𝜓𝜓𝑛𝑛𝑟𝑟 ,𝑙𝑙,𝑚𝑚𝑙𝑙 (𝑟𝑟, 𝜃𝜃, 𝜙𝜙)𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖/ℏ 8.25

with

𝑢𝑢𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟)


𝜓𝜓𝑛𝑛𝑟𝑟,𝑙𝑙,𝑚𝑚𝑙𝑙 (𝑟𝑟, 𝜃𝜃, 𝜙𝜙) = 𝑟𝑟
𝑌𝑌𝑙𝑙,𝑚𝑚𝑙𝑙 (𝜃𝜃, 𝜙𝜙), 8.26

where 𝑢𝑢𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟) is the eigenfunction for the radial Schrӧdinger equation for an electron in a Coulomb
potential

𝑒𝑒 2
𝑉𝑉(𝑟𝑟) = − 4𝜋𝜋𝜖𝜖 .
0 𝑟𝑟

Specifically the radial function 𝑢𝑢𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟) is a solution to a differential equation

ℏ2 𝑑𝑑 2 𝑢𝑢𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟) 𝑙𝑙(𝑙𝑙+1)ℏ2 𝑒𝑒 2


2𝑚𝑚𝑒𝑒 𝑑𝑑𝑑𝑑 2
+� 2𝑚𝑚𝑒𝑒𝑟𝑟 2
− 4𝜋𝜋𝜖𝜖 𝑟𝑟� 𝑢𝑢𝑛𝑛𝑟𝑟,𝑙𝑙 (𝑟𝑟) = 𝐸𝐸𝑛𝑛𝑟𝑟 ,𝑙𝑙 𝑢𝑢𝑛𝑛𝑟𝑟,𝑙𝑙 (𝑟𝑟) 8.27
0

which satisfy the boundary conditions:


𝑢𝑢𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟) = 0 at 𝑟𝑟 = 0 and at 𝑟𝑟 = ∞. 8.28

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

𝑙𝑙(𝑙𝑙+1)ℏ2 𝑒𝑒 2
𝑉𝑉𝑒𝑒𝑒𝑒𝑒𝑒 (𝑟𝑟) = 2𝑚𝑚𝑒𝑒𝑟𝑟 2
− 4𝜋𝜋𝜖𝜖 𝑟𝑟. 8.29
0

The shape of this potential for an electron with different values of orbital angular momentum 𝑙𝑙 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 𝑙𝑙, the effective potential is attractive for large values of 𝑟𝑟 and repulsive for
𝑑𝑑𝑉𝑉
small values of 𝑟𝑟. By setting 𝑒𝑒𝑒𝑒𝑒𝑒�𝑑𝑑𝑑𝑑 equal to zero, it can be shown that 𝑉𝑉𝑒𝑒𝑒𝑒𝑒𝑒 (𝑟𝑟) has a minimum
value of
𝐸𝐸𝑅𝑅
𝑉𝑉𝑒𝑒𝑒𝑒𝑒𝑒 (𝑟𝑟) = − at 𝑟𝑟 = 𝑙𝑙(𝑙𝑙 + 1)𝑎𝑎0 , 8.30
𝑙𝑙(𝑙𝑙+1)

where 𝑎𝑎0 and 𝐸𝐸𝑅𝑅 are the natural units of length and energy in atomic physics. These units are
defined as follows

4𝜋𝜋𝜖𝜖0 ℏ2
Bohr radius 𝑎𝑎0 = � �
𝑒𝑒 2 𝑚𝑚𝑒𝑒
= 0.529 × 10−10 m.
𝑒𝑒 2
Rydbergy energy 𝐸𝐸𝑅𝑅 = 8𝜋𝜋𝜖𝜖 = 13.6 eV.
0 𝑎𝑎0

𝐸𝐸
The minimum energy of 𝑉𝑉𝑒𝑒𝑒𝑒𝑒𝑒 (𝑟𝑟) = − 𝑙𝑙(𝑙𝑙+1)
𝑅𝑅
implies that the bound states with angular momentum
𝐸𝐸𝑅𝑅
𝐿𝐿 = �𝑙𝑙(𝑙𝑙 + 1)ℏ have energies between 𝐸𝐸 = − �𝑙𝑙(𝑙𝑙 + 1) and 𝐸𝐸 = 0. It also implies that the
spatial extent of the bound states of eigenfunction with low angular momentum is in the order of 𝑎𝑎0
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 𝑟𝑟, has uncertain
momentum of the order of ℏ/𝑟𝑟, 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
2𝑚𝑚𝑒𝑒 𝑟𝑟 2
region of size 𝑟𝑟 can roughly be

ℏ2 𝑒𝑒 2
𝐸𝐸 ≈ − . 8.31
2𝑚𝑚𝑒𝑒𝑟𝑟 2 4𝜋𝜋𝜖𝜖0 𝑟𝑟

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

𝐸𝐸 ≈ −𝐸𝐸𝑅𝑅 at 𝑟𝑟 ≈ 𝑎𝑎0 . 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 𝐸𝐸 ≈ −𝐸𝐸𝑅𝑅 and 𝐸𝐸 = 0.

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

ℏ2 𝑑𝑑 2 𝑢𝑢𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟) 𝑙𝑙(𝑙𝑙+1)ℏ2 𝑒𝑒 2


2𝑚𝑚𝑒𝑒 𝑑𝑑𝑑𝑑 2
+� 2𝑚𝑚𝑒𝑒 𝑟𝑟 2
− 4𝜋𝜋𝜖𝜖 𝑟𝑟� 𝑢𝑢𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟) = 𝐸𝐸𝑛𝑛𝑟𝑟,𝑙𝑙 𝑢𝑢𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟), 8.33
0

satisfying the boundary conditions 𝑢𝑢𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟) = 0 at 𝑟𝑟 = 0 and at 𝑟𝑟 = ∞. 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
𝐸𝐸𝑅𝑅
𝐸𝐸 = − (𝑛𝑛 2 with 𝑛𝑛𝑟𝑟 = 0, 1, 2, 3, … 8.34
𝑟𝑟 +𝑙𝑙+1)

The quantum number 𝑛𝑛𝑟𝑟 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 𝑛𝑛𝑟𝑟 + 𝑙𝑙, have the same energy. Due to this degeneracy, energy levels of the hydrogen atom
are given as
𝐸𝐸
𝐸𝐸 = − 𝑛𝑛𝑅𝑅2 , 8.35

where 𝐸𝐸𝑅𝑅 is the Rydberg energy and 𝑛𝑛 is the quantum number given by 𝑛𝑛 = 𝑛𝑛𝑟𝑟 + 𝑙𝑙 + 1.

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

It can also be shown that the radial eigenfunction 𝑢𝑢𝑛𝑛𝑟𝑟,𝑙𝑙 (𝑟𝑟) belonging to the eigenvalue 𝐸𝐸𝑛𝑛𝑟𝑟 ,𝑙𝑙 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 𝐸𝐸 = ℏ 𝛼𝛼 �2𝑚𝑚 falls off exponentially like 𝑒𝑒 −𝛼𝛼𝛼𝛼 . In a similar way the
radial eigenfunction of an electron in a Coulomb potential with binding energy
𝐸𝐸𝑅𝑅 ℏ2 1
𝐸𝐸 = = 2𝑚𝑚 , falls off exponentially at large 𝑟𝑟 like 𝑢𝑢𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟) ∝ 𝑒𝑒 −𝑟𝑟/𝑛𝑛𝑎𝑎0 .
𝑛𝑛2 𝑒𝑒 𝑛𝑛2 𝑎𝑎02
(2) Because of the singular nature at 𝑟𝑟 = 0 of the centrifugal potential energy
𝑙𝑙(𝑙𝑙 + 1)ℏ2�
, the behaviour of the eigenfunction at small 𝑟𝑟 governed by the orbital
2𝑚𝑚𝑒𝑒 𝑟𝑟 2
angular momentum and is given by 𝑢𝑢𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟) ∝ 𝑟𝑟 𝑙𝑙+1.
(3) The radial quantum number 𝑛𝑛𝑟𝑟 denotes the number of nodes between 𝑟𝑟 = 0 and 𝑟𝑟 = ∞,
the eigenfunction 𝑢𝑢𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟) is proportional to a polynomial with 𝑛𝑛𝑟𝑟 zeros. If this polynomial
is denoted by 𝑝𝑝𝑛𝑛𝑟𝑟,𝑙𝑙 (𝑟𝑟), we have
𝑢𝑢𝑛𝑛𝑟𝑟,𝑙𝑙 (𝑟𝑟) ∝ 𝑝𝑝𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟).

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

∞ 2
𝑢𝑢𝑛𝑛𝑟𝑟,𝑙𝑙 (𝑟𝑟) = 𝑁𝑁𝑝𝑝𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟)𝑟𝑟 𝑙𝑙+1 𝑒𝑒 −𝑟𝑟/𝑛𝑛𝑎𝑎0 , where 𝑁𝑁 is the constant ensuring that the ∫0 �𝑢𝑢𝑛𝑛𝑟𝑟 ,𝑙𝑙 (𝑟𝑟)� 𝑑𝑑𝑑𝑑 = 1
normalization condition is satisfied.

Expressions of the radial eigenfunctions with low values of the angular momentum quantum number
𝑙𝑙 and low values of the radial quantum number 𝑛𝑛𝑟𝑟 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 𝒖𝒖𝒏𝒏,𝒓𝒓 (𝒓𝒓) for an electron in a
hydrogen atom with radial quantum numbers 𝒏𝒏𝒓𝒓 = 𝟎𝟎, 𝟏𝟏, 𝐚𝐚𝐚𝐚𝐚𝐚 𝟐𝟐 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 𝑛𝑛 = 𝑛𝑛𝑟𝑟 + 𝑙𝑙 + 1 and a
letter to designate 𝑙𝑙; the letter 𝑠𝑠 is used for 𝑙𝑙 = 0, 𝑝𝑝 for 𝑙𝑙 = 1, 𝑑𝑑 for 𝑙𝑙 = 2, and 𝑓𝑓 for 𝑙𝑙 = 3.

You might also like