Sie sind auf Seite 1von 95

LECTURE NOTES ON

Physics 18324410
QUANTUM PHYSICS
Tjipto Prastowo, Ph.D
Endah Rahmawati, M.Si
Department of Physics
Faculty of Mathematics and Natural Sciences
The State University of Surabaya
December 2012
TO THE STUDENT WE LOVE
Lecture Notes on Quantum Physics contain topics given to the third year students of an
International Class Program in Department of Physics, Faculty of Mathematics and Natural
Sciences, The State University of Surabaya twice a week. These notes focus upon four major
discussions written as separate chapters: quantum theory revisited given in Chapters 1-5; basic
physics principles of quantum mechanics discussed in Chapter 6; Schr odinger wave-mechanics
approach and three dimensional formulation for hydrogen atom using spherical coordinates in
Chapters 7 and 8, respectively. Each chapter is accompanied with some exercises suitable for
students homework assignments. To master materials covered, you need not just knowledge
but skill. This can only be obtained through continual practices. Perhaps, you may obtain
a supercial knowledge by listening to lectures, but you cannot reach the skill expected by
that way. It is common to come across conversation between students like this I understand it
but I cant do the problem ! This student feels uncomfortable with some problems although
they look so easy to do.
The above example shows lack of practice and hence lack of skill required in this course.
Our dearest students, please always study with pencil and paper at hand. You will nd that
the more able you are to choose eective methods of solving problems the easier it will be
for you to master new materials. This costs you nothing but practice, practice and again
practice. Please do remember that the best way to learn to solve problems is to solve them.
We eventually welcome good comments on the content of this course from all readers for
further improvement of these notes as the availability of the notes is important to improve
the quality of learning and teaching processes, particularly in the course of Quantum Physics.
Hope these notes are useful for all users in the department.
All the best,
Kampus Unesa Ketintang, 31 December 2012
T jipto Trastowo
ii
iii
General Guidance
PHYSICS 18324410: QUANTUM PHYSICS
Pre-requisites: Fundamental Physics (I and II)
Modern Physics
Mathematical Physics (I and II)
Lecturers: Tjipto Prastowo, Ph.D and Endah Rahmawati, M.Si
References: Libo, 1980; Beiser, 1988; Gasiorowicz, 1996;
Serway et al., 2005; McMahon, 2006; Harris, 2007
Time and Place: Monday 7-8.30 am, C12 and Friday, 9-10.30 am, D4
Marking Scheme: NA = 20%P + 20%UTS + 30%T + 30%UAS
NA=Final Mark, P=Presence, UTS=Mid-Exam, T=Homework, UAS=Final Exam
Notes:
1. Students are not allowed to join the class for being late (a maximum of 15 minutes from
the starting time is permitted), except for reasonable arguments.
2. Each lecturer contributes an equal proportion of mark to the nal mark.
3. P is possibly reduced to a minimum.
4. UTS = 100% taken from Quiz
5. T = 100% taken from Homework
6. UAS normally contains 4 but possibly 5 problems.
7. Homework will be distributed to class members and all students are required to hand
the completed assignments in within a given time. Various penalties will be given for
any delay, i.e, 25% discounted mark for a one-day delay and 50% for a two-day delay.
There will be no mark given for those who submit the assignments more than two-day
delays.
8. No additional assignments or examinations after formal exam (both Mid and Final),
except for specied reasons with very limited permission given or medical examination
required.
9. Students are allowed to work with their notes and books in both Mid and Final Exams.
10. Other important issues, if any, will be discussed in the class. Students are strongly
encouraged to be active and well-prepared. If possible, tutorial is available for a further,
detailed description of each topics.
Course Contents
1. Chapter One: The Black-Body Radiation (Week 1)
classical black-body radiation, Stefan-Boltzmanns law, Wiens displacement law,
Rayleigh-Jeans law, Plancks radiation formula
2. Chapter Two: The Nature of Electromagnetic Radiation (Week 2)
photo-electric eect, Compton eect
3. Chapter Three: The Bohr Model of the Atom (Week 3)
early atomic model, Bohrs theory of an atom, Bohrs explanation to the theory,
hydrogen spectral lines, Bohrs correspondence principle
4. Chapter Four: The Wave Behaviour of Sub-Atomic Particles (Week 4)
de Broglie hypothesis and its implications, Davisson-Germer experiment
5. Chapter Five: The Heisenberg Uncertainty Principle (Week 5)
mathematical basis of the Heisenberg uncertainty principle, its interpretation and
consequences of the uncertainty principle
6. Quiz I : Chapters 1, 2, 3, 4, and 5 (Week 6)
7. Chapter Six: The Basic Physics Principles of Quantum Mechanics (Week 7)
observables, operators, wavefunctions, quantum measurements in microscopic world,
Born interpretation, normalisation procedure, Hilbert space in quantum mechanics
and superposition principle, expectation values
8. Chapter Seven: The Wave-Mechanics Approach (Weeks 8, 9, 10, and 11)
steady state time-independent Schr odinger equation, stationary states, eigen values
of energy, free particle, innite square well, simple-step potential, reection and
transmission coecients, harmonic oscillator
9. Chapter Eight: The Schrodinger Equation in Spherical Coordinates (Weeks 12, 13,
14, and 15)
solution to the time-independent Schrodinger equation in spherical coordinates,
theory of hydrogen atom revisited, orbital and spin angular momenta, spin-orbit
coupling and total angular momentum
10. Quiz II: Chapters 6, 7, and 8 (Week 16)
iv
Contents
1 THE BLACK-BODY RADIATION 1
1.1 Thermal Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Wiens Displacement Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Rayleigh-Jeans Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.4 Plancks Formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2 THE NATURE OF ELECTROMAGNETIC RADIATION 7
2.1 Photo-Electric Eect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Compton Eect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3 THE BOHR MODEL OF THE ATOM 15
3.1 The Early Atomic Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2 Bohrs Atomic Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3 Bohrs Explanation to the Model . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.4 Hydrogen Spectral Lines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.5 Bohrs Correspondence Principle . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4 THE WAVE BEHAVIOUR OF SUB-ATOMIC PARTICLES 27
4.1 De Broglie Hypothesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.2 Implications of the De Broglie Hypothesis . . . . . . . . . . . . . . . . . . . . 30
4.3 The Davisson-Germer Experiment . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
5 THE HEISENBERG UNCERTAINTY PRINCIPLE 37
5.1 Mathematical Basis for the Uncertainty Principle . . . . . . . . . . . . . . . . 38
5.2 Interpretations of the Uncertainty Principle . . . . . . . . . . . . . . . . . . . 41
5.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
6 THE BASIC PHYSICS PRINCIPLES OF QUANTUM MECHANICS 45
6.1 Observables, Operators, and Wavefunctions . . . . . . . . . . . . . . . . . . . 45
6.2 Born Interpretation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
6.3 Normalisation Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
6.4 Hilbert Space and Superposition Principle . . . . . . . . . . . . . . . . . . . . 49
6.5 Expectation Values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
v
vi CONTENTS
6.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
7 THE WAVE-MECHANICS APPROACH 53
7.1 Stationary States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
7.2 The Free Particle and Innite Square Well . . . . . . . . . . . . . . . . . . . . 56
7.3 Simple-Step Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
7.3.1 The Continuity Equation and Current Density . . . . . . . . . . . . . . 59
7.3.2 Reection and Transmission Coecients . . . . . . . . . . . . . . . . . 60
7.3.3 Cases where E is greater than V . . . . . . . . . . . . . . . . . . . . . . 63
7.3.4 Cases where E is less than V . . . . . . . . . . . . . . . . . . . . . . . . 64
7.4 The Harmonic Oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
7.4.1 The Algebraic Method . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
7.4.2 The Analytic Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
7.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
8 THE SCHR

ODINGER EQUATION IN SPHERICAL COORDINATES 75


8.1 The Schr odinger Equation in Spherical Coordinates . . . . . . . . . . . . . . . 76
8.1.1 Separation of Dynamic Variables . . . . . . . . . . . . . . . . . . . . . 76
8.1.2 The Angular Component . . . . . . . . . . . . . . . . . . . . . . . . . . 78
8.1.3 The Radial Component . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
8.2 The Hydrogen Atom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
8.3 Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
8.3.1 Orbital Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . 86
8.3.2 Spin Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . . 86
8.3.3 Total Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . . 86
8.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
Bibliography 89
Chapter 1
THE BLACK-BODY RADIATION
The achievement made in theoretical and experimental physics towards the end of the-19th
century had been immense. Classical mechanics and thermodynamics were well understood.
The concept of electromagnetics was also remarkably established. It was found that light is a
form of electromagnetic waves, providing a rm theoretical framework for the wave theory of
light, which could account for most phenomena in optics. Thus, most physicists at that time
believed that the combination of the three subjects could account for all physical phenomena.
In fact, it was a period of a turmoil state when there were surprising experimental results,
which based on classical theory were totally inexplicable. One dilemma lay in the observations
of thermal radiation. Existing classical theory was unable to explain the observed frequency
(or wavelength) of radiant energy. The incredibly greatest minds of physics in the period
were about to begin. Out of the turmoil state came a new philosophy of science a new
way of thinking the so-called Quantum Physics. The way we think about natural laws
particularly in microscopic world has totally changed since then. In this new paradigm, light
is considered not only as a wave but also as a series of bundles of energy called quanta.
Unlike Newtonian mechanics where dynamic variables, such as position and momentum can
be accurately determined with high precision, quantum world makes these variables uncertain
in that all measurements made are undeterministic.
This chapter is aimed to examine the birth of quantum physics that has shaped the world
of physics. We rst begin with experimental observations of the spectrum of thermal radiation,
what comes later to be known as black-body radiation, that put physicists at that time
into a Pandora Box. The second issue to discuss is the theoretical work of Wien in 1893, who
provided a primitive formula for the distribution of radiant energy. In line with this, we then
discuss the contributions of Rayleigh-Jeans in 1900 to the problem in question. Finally, we
revisit the notion of the photon concept of electromagnetic radiation in the light of Plancks
ideas proposed in 1901. This great work of Planck, along with relativity theory suggested by
Einstein, serves as a connecting bridge between classical physics and modern physics.
1
2 1. THE BLACK-BODY RADIATION
i
n
t
e
n
s
i
t
y

i
n

w
a
t
t
/
m
2
(a) (b)
Figure 1.1: Plots of radiant energy c, showing the characteristics of thermal radiation
of a black-body as a function of wavelength : (a) at various temperatures T and (b) at a
given temperature with theoretical predictions suggested by Rayleigh-Jeans and Planck
(taken from Ch.1, Quantum Physics, Stephen Gasiorowicz, 1996).
1.1 Thermal Radiation
In 1859, Kircho proposed a theorem regarding with the relation between the coecients
of emission and absorption of electromagnetic radiation; how these could be related to the
spectrum of thermal equilibrium radiation, the so-called black-body radiation phenomenon.
He challenged the community to work this out. A black-body is an object that absorbs all
the radiant energy falling on it and hence the black-body reects no light, for which it would
appear as black. A black-body is thus a perfect absorber, as well as a perfect emitter. Kircho
suggested a functional dependence of the radiant energy c on temperature T and frequency
(or wavelength ),
c = c (T, ) = c (T, ) (1.1)
It has since then been of fundamental interest among physicists during the last two decades of
the-19th century to nd out what the explicit form of the function is. This has been examined
through an approach of both experimental and theoretical considerations.
In 1879, Stefan based on experimental results argued that the radiant energy emitted by a
hot body is proportional to the fourth power of its absolute temperature. The same conclusion
based on classical thermodynamics was derived by Boltzmann in 1884. The ndings led to
the Stefan-Boltzmanns law. Here, we are not going to provide the derivation of the law.
1.2. Wiens Displacement Law 3
Rather, we pick up the result as follows,
c T
4
(1.2)
where c denotes the energy emitted per unit time per unit area from the surface of a black-body
at a given temperature T, and is the Stefan-Boltzmann constant (5.67 10
8
Wm
2
T
4
).
1.2 Wiens Displacement Law
In 1893, Wien proposed a formula to answer to the question previously posed by Kircho.
Wiens displacement law was derived using a combination of classical electromagnetics
and thermodynamics, as well as dimensional analyses. From the two basic concepts, it can be
shown that
T
1
(1.3)
The above relation says that the wavelength of a set of radiant waves changes inversely with
absolute temperature T, as seen in Figure 1.1(a).
Wien then went further by combining his own law with the Stefan-Boltzmann law. If we
dene c() as the energy density radiated, then we can write
c() T
5

5
(1.4)
for which we can derive
c =
5
f(T) =
3
g(/T) (1.5)
where both f = f(T) and g = g(/T) are constants. The resulting prediction (1.5) matches
experimental observations, particularly for a region of small wavelengths (or high frequencies),
but it was brokendown for a region of large wavelengths (or low frequencies). Equation (1.5)
is the complete form of Wiens law and sets constrains on the black-body radiation spectrum.
But it is actually incomplete because f = f(T) and g = g(/T) are both undetermined.
The complete form of the energy density of spectral distribution cannot be determined from
classical physics.
1.3 Rayleigh-Jeans Law
In 1900, Rayleigh analysed experimental data of black-body radiation and its corresponding
theory. His interest was stimulated by the inadequacies of Wiens law in (1.5) in accounting
for the large-wavelength (or low-frequency) behaviour of black-body radiation as a function
4 1. THE BLACK-BODY RADIATION
of temperature. The reason for the inclusion of Jeans name in what became known as the
Rayleigh-Jeans law is that there was an error in Rayleighs analysis, which was corrected
by Jeans in 1906. Based on theoretical considerations, Rayleigh-Jeans found that the spectral
energy density radiated is given by
c =
8
2
c
3
E (1.6)
where E is something that must be in a unit of energy. Here Rayleigh put E = kT based on
the principle of the equipartition of energy from classical kinetic theory into (1.6). Hence, we
have
c =
8
2
c
3
kT (1.7)
for the energy spectral distribution at a given temperature. This formula is only valid at low
frequencies (or large wavelengths). It does not hold for high frequencies (or small wavelengths)
because the spectrum of black-body radiation does not increase as
2
to innite frequency, as
shown in Figure 1.1(b).
Thus, both Wiens law and Rayleigh-Jeans law are counter-part in that they complete one
to another. However, what people want is a comprehensive theory that can be used to explain
the black-body radiation for all ranges of frequency or wavelength. It seems that the existing
theory based on classical view of thermal radiation of a black-body where such radiation is
considered to propagate in space as waves is no longer relevant.
1.4 Plancks Formula
Here we do not intend to derive explicitly what Planck did in his best times. Rather, we try to
connect (1.6) previously derived by Rayleigh to the correct formula. From classical statistics,
E is related to the mean energy of a harmonic oscillator having two degrees of freedom, which
should be
1
2
kT +
1
2
kT = kT. This is what exactly Rayleigh did in deriving his law, which
turns out to be the correct expression for the black-body radiation law at low frequencies (or
large wavelengths), as shown in Figure 1.1(b).
An interesting question is that why did Planck not make such substitution for E ? Firstly,
the equipartition theorem of Maxwell-Boltzmann is a result of statistical thermodynamics,
which was a point of view that he had rejected. Secondly, the Maxwells kinetic theory could
not account for the problem of specic heats of diatomic gases at high temperatures. Thirdly,
Planck did not agree fully with the Boltzmanns statistical approach.
In 1901, Planck worked out the problem of fundamental interest the energy spectral
distribution of the black-body radiation. Planck has introduced the concept of quantisation
by assuming that radiation consists of a bundle of quanta each having an energy which is
proportional to the radiant frequency . The energy density distribution c of the black-body
1.4. Plancks Formula 5
radiation can then be written as a function of radiant frequency as follows,
c() =
8 h
3
c
3
1
e
h/kT
1
(1.8)
The above expression satises all regions of frequency and wavelength in the spectrum, as
seen in Figure 1.1(b). The Plancks formula in (1.8) for the black-body radiation is valid
over the whole spectrum of wavelengths, or accordingly frequencies. It covers the empirically
observed ndings of the spectral distribution of electromagnetic radiation that are previously
unexplained by classical theories of both electrodynamics and thermodynamics.
It is straightforward to integrate (1.8) to nd the total energy density of radiation | in
the black-body spectrum,
| =
_

0
c() d =
8 h
c
3
_

0

3
e
h/kT
1
d (1.9)
for which the total power of radiation emitted per unit area can be calculated from classical
theory of radiation, J = c |/4, such that we can write
J =
2k
4
T
4
h
3
c
2
_

0
(h/kT)
3
e
h/kT
1
d
_
h
kT
_
=
2k
4
T
4
h
3
c
2
_

0
x
3
e
x
1
dx (1.10)
where J denotes the total energy radiated per unit time per unit area and x h/kT. The
integral can be evaluated with the help of some steps below (see p.5, Gasiorowicz, 1996),
_

0
x
3
e
x
1
dx =
_

0
x
3
dx e
x

n=0
e
nx
=

n=0
1
(n + 1)
4
_

0
y
3
e
y
dy = 6

1
1
n
4
=

4
15
Equation (1.10) then becomes
J =
_
2
5
k
4
15h
3
c
2
_
T
4
= T
4
(1.11)
where the quantity in the bracket is replaced with , and hence
=
2
5
k
4
15h
3
c
2
(1.12)
representing the Stefan-Boltmann constant, as previously shown in (1.2).
Most people at the time Planck proposed his ideas in verifying the experimental data of
black-body radiation got an impression that the Plancks concept of energy quantization was
only for the case of thermal radiation. It remained until Einstein who used the work of Planck
to show that light is quantized in the case of photo-electric eect.
6 1. THE BLACK-BODY RADIATION
1.5 Exercises
1. Show that the Rayleigh-Jeans spectral distribution of black-body radiation given in (1.7)
is of the form required by Wiens law given in (1.5).
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
2. Obtain the correct form Wiens undetermined function f(T) from Plancks formula
given in (1.8).
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
3. Use (1.8) to prove Wiens displacement law in the form of

max
T = constant
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
4. (a) Use the Stefans law to calculate the total power radiated per unit area by a tungsten
lament at a temperature of 3000 K (assume that the lament is an ideal radiator).
(b) If the tungsten lament of a lightbulb is rated at 75 W, what is the surface area of
the lament ? (assume that the main energy is lost due to radiation).
(taken from Ch.3, Modern Physics, 3rd edition, Serway et al., 2005).
5. At what wavelength does the human body emit the maximum electromagnetic radiation?
(assume that the skins temperature is about 70

F).
(taken from Ch.3, Modern Physics, 2nd edition, Randy Harris, 2007).
Chapter 2
THE NATURE OF
ELECTROMAGNETIC RADIATION
As discussed in Chapter 1, Planck proposed a remarkable formula for the spectral distribution
of black-body radiation, which shed light on the photon concept of quantisation. There soon
followed an understanding of the quantum nature of electromagnetic radiation. However, it
was not until Einstein who suggested that the concept was applied to real cases when metals
were irradiated with light. Another support for the concept came from experimental evidence
of the particle nature of photons observed by Compton. These two fascinating phenomena
contribute signicantly to the development of modern physics. One of the corner stones of
quantum physics is wave-particle duality, meaning that things may behave as waves or
particles depending upon a physical situation. In a classical situation, say the propagation
of sunlight in space for example, such electromagnetic radiation is considered as waves with a
continuous ranges of energy spread out. In this chapter, however, we study the complementary
topic electromagnetic radiation behaves as a collection of discrete particles.
2.1 Photo-Electric Eect
An important contribution to the energy quantisation of electromagnetic radiation came from
the work of Einstein who in 1905 applied the concept of the quantum nature of light to explain
unresolved, experimental ndings in the photo-electric eect originally discovered by Hertz
in 1887. Hertz found peculiar properties of metals while observing his famous experiments on
electromagnetic waves to validate the Maxwells hypothesis of the speed of light.
The experimental set up of the photo-electric eect experiments is depicted in Figure 2.1,
where a metal plate of a certain material placed as a cathode in the apparatus is irradiated with
light of a given frequency . There were a number of surprising results from the experiments.
The most remarkable feature was that the kinetic energies of the photo-electrons emitted from
7
8 2. THE NATURE OF ELECTROMAGNETIC RADIATION
Figure 2.1: Sketch of the experimental set up of photo-electric eect experiments (taken
from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
the metal surface are independent of the intensity of the incident radiation, but are linearly
dependent upon the incident frequency (or wavelength). This feature could not be explained
by classical theory of electromagnetic radiation at that time.
Einstein proposed a brilliant idea to solve the problem in question by assuming that the
radiation consists of a collecting bundle of energy called quanta of the same energy h. When
the bundle of energy is absorbed by the metal, electrons may receive sucient energy to escape
from the metal surface, against the energy that binds them. When the intensity of the incident
light is increased more electrons will be ejected from the surface, with their kinetic energies
remaining unaltered although the associated photo-electric current measured by Ammeter is
larger, as shown in Figure 2.2(a). The result for an increase in the photo-current as the light
intensity increases is actually predicted from classical view of electromagnetic radiation. But,
the result for the independence of kinetic energy on the light intensity is totally unexpected.
This is a point where the glorious story of the quantum nature of light begins.
To cope with the unexpected results above, Einstein further suggested that the maximum
kinetic energy K
max
of the photo-electrons linearly depends on both the incident frequency
(or wavelength) and the work function W of a given material, as shown in Figure 2.2(b).
The work function of a metal here is dened as the minimum amount of work necessary to
remove the electrons from the metal surface. Thus, this kinetic energy is expressed as
K
max
= h W
=
hc


hc

o
(2.1)
where the term max is used to show that W is needed to free the least strongly bond electrons.
2.2. Compton Eect 9
Figure 2.2: Some of the surprising results for experiments of photo-electric eect,
showing (a) photo-current is only aected by the intensity of incident light with a xed
stopping potential for a given material; and (b) the dependence of the maximum kinetic
energy of photo-electrons on both the incoming frequency and the material used for a
cathode-plate (taken from Ch.3, Modern Physics, Serway et al., 2005).
Determination of the magnitude of the work function for a given material involves placing
a photo-cathode in an opposing potential so that, when the potential reaches a certain value,
the ejected electrons can no longer reach a collecting plate that serves as an anode, causing
photo-electric currents to fall to zero. This situation occurs at a value of an applied voltage
called the stopping potential V
s
. In this situation, K
max
= eV
s
such that (2.1) becomes
eV
s
= h(
o
) (2.2)
where we have used W = h
o
with
o
being the threshold frequency, that is, the minimum
frequency needed to eject the electrons from the metal surface. Note that based on (2.2) eV
s
must be a linear function of the incident frequency, with the slope of the straight line being h
known as the Plancks constant experimentally found to be 6.63 10
34
Js, independent of
the nature of the material chosen.
The central point in Einsteins contribution to the photo-electric eect experiments is that
electromagnetic radiation behaves as a collection of particles, each having a discretized energy
as opposed to the classical view of continuous wave energy for the radiation. With all respects,
Einsteins explanations agree well with experimental evidence, for which the work earned him
the Nobel Prize in physics in 1921.
2.2 Compton Eect
In 1922, Compton provided direct experimental evidence for the correctness of the photon
concept of electromagnetic radiation. In the experiment known as the Compton scattering,
10 2. THE NATURE OF ELECTROMAGNETIC RADIATION
Figure 2.3: Sketch of the Compton eect, showing (a) initial and (b) nal states of
the X-ray scattering experiment (taken from Ch.2, Introductory Quantum Mechanics,
Richard Libo, 1980).
he used a beam of X-ray of initial energy h and momentum p to bombard a targeted electron
initially at rest, as shown in Figure 2.3(a). It was found that the X-ray beam was scattered in
a manner which is not consistent with classical theory of electromagnetic radiation. According
to the classical wave theory, the electron would scatter electromagnetic energy in all directions
at the same frequency as the incoming radiation frequency. But in fact, the scattered X-ray
beam has a dierent, outgoing frequency. To this end, Compton argued that the incoming
radiation should be treated as a beam of photons with individual photons scattering elastically
o individual electrons.
Based on the sketch of the Compton eect experiment (see Figure 2.3) where both energy
and momentum must be conserved, the conservation of energy is then given by
h + m
o
c
2
= h

+ m
o
c
2
(2.3)
where m
o
is the rest mass of electron, and

are the frequencies of incoming and scattered
X-ray beam, respectively, and is dened to be (1
2
)
1/2
, where = v/c. Note that
here we use the relativistic expression for energy, and later momentum, as the recoil electron
is frequently observed to move very fast after the collision. We do not need to worry about
the expression for photons as the non-relativistic expression for objects that always move at
speed c does not exist.
Unlike energy, conservation of linear momentum is separately given for each direction, one
2.2. Compton Eect 11
component in the direction of the incident radiation and the other component perpendicular
to the former. The equations denoting the momentum conservation are
h

= m
o
c cos +
h

cos (2.4)
and
0 = m
o
c sin +
h

sin (2.5)
where is an angle through which the photon is scattered and is an angle of the recoil
electron, kicked o from its original position as seen in Figure 2.3(b).
Compton pointed out that the scattered beam has a larger wavelength than the incident,
associated with less amount of energy for the scattered beam. The dierence in wavelength
between the two beams can therefore be directly derived from the energy and momentum
conservation, and is measured as an increase in wavelength of the incoming radiation in
a billiard-ball type of collision with a stationary electron. We can then write
=

=
h
m
o
c
(1 cos ) (2.6)
where and

denote the wavelengths of the incoming radiation and scattered photon,


respectively, and h/m
o
c = 0.024

A is called the Compton wavelength. Thus, the dierence
in the wavelength between the incident and scattered photons depends only on the angle
of the scatter.
Let us take a closer look at (2.6). If the targeted electron is replaced with an atom, for
example, then m
o
would be the mass of the atom, which is much more larger than the electron
mass. Consequently, the shift in the wavelength is suciently small to observe so that it
is eectively zero. This situation is similar to the case of an elastic collision between a small
projectile and a relatively massive body as a target. After a head-on collision, the projectile
is reected back from the body with the same values of both kinetic energy and momentum,
but is in the opposite direction. For this case, the increase in wavelength is maximum for
which this type of scattering is called backward scatter of the incoming photon with = 180

because the collision imparts the maximum possible energy to the recoil electron.
The measurements of the shift in the wavelength and the kinetic energy of the recoil
electron are in good agreement with the Comptons theoretical predictions. These provide
convincing justication of the correctness of the classical, billiard-ball collision interpretation,
demonstrating the particle behaviour of the photon. The results for the Compton scattering
experiments opened up the problems of classical view of electromagnetic radiation theory, for
which Compton won the Nobel Prize in 1927.
12 2. THE NATURE OF ELECTROMAGNETIC RADIATION
2.3 Exercises
1. The photo-electric threshold of tungsten is 2300

A. Determine the kinetic energy of the
electrons ejected from the metal surface by ultraviolet light of wavelength 1900

A.
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
2. The work function of zink is known to be 3.6 eV. What is the energy of the most energetic
photo-electron emitted by ultraviolet light of wavelength 2500

A ?
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
3. Photo-electrons emitted from a cesium plate that are illuminated with ultraviolet light
of wavelength 2000

A are stopped by a potential of 4.21 V. What is the work function
of cesium ?
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
4. Ultraviolet light of wavelength 3500

A falls on a potassium plate. The maximum kinetic
energy of the photo-electrons is 1.6 eV. What is the work function of potassium ?
(taken from Ch.1, Quantum Physics, Stephen Gasiorowicz, 1996).
5. Consider the following metals lithium, beryllium, and mercury, having work functions
of 2.3 eV, 3.9 eV, and 4.5 eV, respectively. If light of wavelength 300 nm is incident on
each of these metals, then determine
(a) which metal exhibit the photo-electric eect;
(b) the maximum kinetic energy for the photo-electrons in each case
(taken from Ch.3, Modern Physics, Serway et al., 2005).
6. A photon of energy h collides with a stationary electron of rest mass m
o
. Show that it
is not possible for the photon to impart all its energy to the electron.
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
7. A 100 MeV photon collides with a proton at rest. What is the maximum possible energy
loss for the photon ?
(taken from Ch.1, Quantum Physics, Stephen Gasiorowicz, 1996).
8. A 100 keV photon collides with an electron at rest. It is scattered at 90

. What is its
energy after the collision ? What is the kinetic energy in eV of the electron after the
collision, what is the direction of the recoil electron ?
(taken from Ch.1, Quantum Physics, Stephen Gasiorowicz, 1996).
2.3. Exercises 13
9. A beam of X rays is scattered by electrons at rest. What is the energy of the Xray if
the wavelength of the X-ray scattered at 60

to the beam axis is 0.035



A ?
(taken from Ch.1, Quantum Physics, Stephen Gasiorowicz, 1996).
10. Gamma rays (high-energy photons) of energy 1.02 MeV are scattered from electrons
that are initially at rest. If the scattering is symmetric, that is, if = then determine
(a) the scattering angle ;
(b) the energy of the scattered photons
(taken from Ch.3, Modern Physics, Serway et al., 2005).
14 2. THE NATURE OF ELECTROMAGNETIC RADIATION
Chapter 3
THE BOHR MODEL OF THE ATOM
3.1 The Early Atomic Model
The rst half of the 20th century was generally considered as a period of remarkable advances
in modern physics, pioneered by the triumphs of relativity and quantum theories, as well as
similar discoveries in the elds of atomic and nuclear physics. As with relativity theory and
quantum theory for electromagnetic radiation previously discussed in the rst four chapters,
this chapter discusses a new way of looking at an atom, in particular the introduction of
quantization concept to atomic scales yet in its primitive form as only a single quantum
number involved in describing the dynamics of the atom. To start with, we here give a brief
introduction to the development of atomic model.
The early modern, if it were worth to say, atomic model was the Thomson model in 1898,
in which an atom was viewed as a homogeneous solid sphere of uniformly distributed mass
having positive charge and negatively charged electrons distributed over the sphere surface to
produce a totally, electrically neutral atom. But, this model failed to explain the various lines
of atomic spectra found from direct observations. The mysterious lines and the corresponding
revised model were then suggested by Rutherford in 1912 after a series of experiments using
a beam of -particles scattered by a thin metal foil (see Figure 3.1). The important result of
these experiments was that most of atomic mass and its corresponding positive charge lie in
a central region of the atom called the nucleus, while the electrons surrounding it. However,
the idea of such a nuclear atom raised some fundamental questions as follows: (1) If there are
only Z protons within the nucleus comprising roughly a half of the total mass of the nucleus,
what composes the other half ? (2) What provides the cohesive force to keep protons within
a conned region of 10
14
m, the size of the nucleus ? (3) What drives the electrons to move
around the nucleus ? (4) How does this motion account for the observed spectral lines ?
Among these, Rutherford gave his best speculation to answer to the rst two questions. But,
he had certainly no answer to the other two until the work of Bohr came into play in 1913.
15
16 3. THE BOHR MODEL OF THE ATOM
Figure 3.1: A beam of -particles scattered by a dense positively charged nucleus
(taken from Ch.4, Modern Physics, Serway et al., 2005).
3.2 Bohrs Atomic Model
The Rutherford atomic model is based on the fact that the electrons must revolve the nucleus.
Meanwhile, classical electromagnetic theory of radiation claims that accelerated electrons
revolving the nucleus will radiate some energy that makes orbital radius of the electrons
become less and less. In this sense, the radius decreases steadily followed by an ever-increasing
radiant frequency corresponding to the energy released. The atom may therefore be unstable
and collapsing as the electrons plunge into the nucleus, as shown in Figure 3.2.
The wrong deductions above that lead to a continuous emission spectrum were well tackled
by Niels Bohr, a young Danish physicist. He argued that classical electromagnetic radiation,
theoretically proposed by Maxwell and experimentally conrmed by Hertz long before Bohrs
new model of the atom shook the world of physics to its foundation, is false when applied to
systems at a microscopic scale. Based on the great work of Planck, he applies the concept of
quantized levels of energy and angular momentum to orbital electrons that occupy stable
stationary states in which no energy is lost. At the same time, based on the corresponding
work of Einstein, he applies the concept of photon emitted when there is a quantum jump
of an electron from a particular stationary state associated with certain values of energy and
angular momentum to another. In this way, Bohr developed a new theory of an atom in 1913
by combining basic physic principles of classical mechanics with a simple quantum theory of
light emission. The Bohr model is then able to explain the existence of atomic spectral lines
of the hydrogen atom, in good agreement with experimental evidence and thereby resolving
the shortcomings of the previous, classical atomic models.
3.2. Bohrs Atomic Model 17
Figure 3.2: Classical radiation theory of an orbital electron having less and less energy
as it approaches the nucleus (taken from Ch.4, Modern Physics, Serway et al., 2005).
In this context, let us now examine in detail the semi-classical, Bohr theory for an atom.
As it originally applies to a hydrogen atom with a single proton and a single electron only,
the basic idea of the Bohrs theory is that the dynamics of the hydrogen is determined by
the Coulomb attraction of a nucleus (the proton) and the electron that drives the electron to
orbit the nucleus. The orbiting electron has an energy derived from a classical expression for
the total energy. This concept is then combined with two quantum postulates as follows :
The hydrogens electron is assumed to orbit the nucleus in concentric circular paths
associated with stationary orbits, each having its own angular momentum L equal to
an integer multiplied by h/2, where h is the Plancks constant. These orbits are then
referred to atomic shells. This can be mathematically expressed as
_
Ld = nh
L = n
(3.1)
where n = 1, 2, 3 .... is an integer, the so-called principal quantum number associated
with the nth stationary orbit at which the electron may be located. The stationary orbits
are also related to discrete values of energy representing quantum states, each having its
own quantum number n, for the orbiting electron. In such stable orbits, this electron
radiates no energy.
When an electron from a quantum state at a particular level of energy jumps into another
quantum state at a lower level of energy, then photon with a certain value of frequency
is emitted (see Figure 3.3). In this case, the radiant frequency is then proportional to
the dierence in the level of energy E between the two states, and is written as
=
E
h
=
E
i
E
f
h
(3.2)
18 3. THE BOHR MODEL OF THE ATOM
Figure 3.3: A photon of frequency is emitted as a result of quantum transition from
an initial state with a particular level of energy E
i
to a nal state with a lower level of
energy E
f
(taken from http://www.kottan-labs.bgsu.edu).
where E
i
and E
f
are levels of energy for the initial and nal quantum states, respectively.
Note that the other way around of atomic transition may occur when an electron absorbs
some amount of energy to move from an initial state of one energy level to a nal state
of a higher energy level. For this case, (3.2) still holds with is the absorbed frequency.
3.3 Bohrs Explanation to the Model
This section describes the uniqueness of the Bohrs model in that it combines well classical and
quantum concepts. The combined ideas then lead to discretized orbit radii and levels of the
electrons energy, and are at the same time prove the Bohrs postulates to be inter-connected.
Let us start with the hydrogen having a single electron only. From the classical point of view,
the total energy E of the hydrogens electron is given by
E =
1
2
mv
2
k
e
2
r
=
L
2
2mr
2
k
e
2
r
(3.3)
where m = 9.11 10
31
kg denotes the mass of the hydrogens electron, e = 1.6 10
19
C is
the electron charge, and k = 9 10
9
Nm
2
C
2
is the Coulomb constant for vacuum.
Next, the dynamics of such an electron can be written in the form of Newtons second law
of motion as follows,
mv
2
r
=
L
2
mr
3
= k
e
2
r
2
(3.4)
Inserting (3.1) to (3.4) yields the hydrogens electron orbit radii r
n
as follows,
r
n
=
n
2

2
mke
2
= n
2
a
o
for n = 1, 2, 3, ... (3.5)
3.3. Bohrs Explanation to the Model 19
Figure 3.4: The rst three Bohr atomic orbits for the hydrogen atom (taken from Ch.4,
Modern Physics, Serway et al., 2005).
where = h/2 and a
o
=
2
/mke
2
= 0.529

A is here called the Bohr radius. It is then clear
from (3.5) that only orbits illustrated in Figure 3.4 with certain values of quantum number n
are allowed to occupy. These discrete orbits are due to the non-classical requirement for the
electrons angular momentum to be an integral multiple of . Such orbits form atomic shells
surrounding the nucleus, where the smallest orbit is called / shell having radius r
1
= a
o
and
is associated with n = 1. For larger orbits with n > 1, the shells are called /, /, A ........
Later on, according to quantum mechanics these orbits are divided into sub-shells.
The above quantization of the electrons orbit radii for the hydrogen atom immediately
leads to the discretized levels of the electrons total energy, which will be derived below.
Substituting (3.4) into (3.3) results in
E =
L
2
2mr
2

L
2
mr
2
=
L
2
2mr
2
(3.6)
With the help of (3.1) and (3.5) to be inserted into (3.6), we can write the allowed levels E
n
of the total energy for the hydrogens electron as
E
n
=
mk
2
e
4
2
2
n
2
=
1
n
2
for n = 1, 2, 3, ... (3.7)
where 1 = mk
2
e
4
/2
2
= 13.6 eV is called the Rydberg constant for the electrons energy.
In other literatures, the Rydberg constant is dierently valued as it is meant for a dierent
quantity (see further 3.4). The negative sign for the electrons total energy in (3.7) suggests
that the electron is bound to the nucleus and thereby requiring some amount of energy to
20 3. THE BOHR MODEL OF THE ATOM
Figure 3.5: A diagram showing the energy-discretized states of the hydrogen atom
and some spectral lines following atomic transitions (taken from Ch.4, Modern Physics,
Serway et al., 2005).
liberate it from the Coulomb attraction. The lowest possible stationary state corresponding
to quantum number n = 1 is generally called the ground state having the smallest energy of
E
1
= 13.6 eV. The next state is the rst excited state, corresponding to quantum number
n = 2 and hence having an energy of E
2
= 3.4 eV.
An energy-level diagram for the hydrogen atom showing the energy-discretized states with
the corresponding quantum numbers is depicted in Figure 3.5. The uppermost level of energy
corresponds to n = and hence E
n
= 0, representing a state in which an electron is free
from Coulomb nuclear eld. When this electron experiences an external force it has a kinetic
energy only for which it is usually termed as a free electron. The minimum energy required to
remove the bound electron from its ground state to a large distance from the nuclear inuence
is called the ionization energy. Thus for the hydrogen, the ionization energy is 13.6 eV.
3.4 Hydrogen Spectral Lines
The great success of the Bohr atomic model for the hydrogen is achieved when it is used
to describe the existence of the hydrogen spectral lines. Instead of a continuous emission
spectrum argued by Rutherfords model, the Bohrs theory predicts that atomic transitions
between levels of energy produce a spectral series of separated lines (see Figure 3.5), each
having its own characteristics; all the observed lines in Lymann series end at a state of n = 1,
lines of Balmer series are down to n = 2, and Paschen lines towards n = 3.
3.4. Hydrogen Spectral Lines 21
Here we derive a mathematical expression for the observed hydrogen spectral lines using
(3.2) and (3.7). Let us suppose a quantum transition from an outer orbit at an initial state of
quantum number n
i
having energy E
i
= 1/n
2
i
to an inner orbit at a nal state of quantum
number n
f
having energy E
f
= 1/n
2
f
. Such a transition emits a photon of frequency
given by
=
1
h
_
1
n
2
f

1
n
2
i
_
(3.8)
The above equation is equivalent to
1

=
1
hc
_
1
n
2
f

1
n
2
i
_
= R
_
1
n
2
f

1
n
2
i
_
(3.9)
where represents the photon wavelength corresponding to the emitted frequency in (3.8)
and R is exactly the same as 1/hc = 1.097 10
7
m
1
, dened to be the Rydberg constant
for the emitted wavelengths of the hydrogen spectral series depicted in Figure 3.5.
When Bohr proposed his theory in 1913, the spectral lines of Balmer series (n
f
= 2 and
n
i
> 2) and Paschen series (n
f
= 3 and n
i
> 3) had already been found two years before.
The Rydberg constant R in (3.9) theoretically derived from the Bohrs quantum theory for
the hydrogen is found to be in good agreement with an empirical value for a constant used to
describe Balmer and Paschen lines. This was followed by the Lymann series (1916) for which
n
f
= 1 and n
i
> 1. In the years to come since the rst three spectral series, Brackett (1922)
and Pfund (1924) observed the spectral lines for n
f
= 4 and n
f
= 5, respectively. Among all
these series, only Balmer lines constitute a range of wavelengths in a visible light spectrum.
Impressed by his successful explanation to the experimental observations of the hydrogen
spectral lines, Bohr then immediately extended his theory to hydrogen-like atoms, such
as He
+
and Li
2+
, in which all but one electron had been removed from a nucleus of positive
charge Ze, where Z denotes the number of proton. For a single electron of negative charge e
orbiting the nucleus of a hydrogen-like atom, the quantized orbit radii in (3.5) becomes
r
n
=
n
2

2
mkZe
2
= n
2
a
o
Z
for n = 1, 2, 3, ... (3.10)
and the corresponding discrete energies in (3.7) becomes
E
n
=
mk
2
Z
2
e
4
2
2
n
2
= Z
2
1
n
2
for n = 1, 2, 3, ... (3.11)
The extended theory was used to explain mysterious lines observed in hot stellar atmospheres.
In the rst paper of his great trilogy published in late 1913, Bohr noted that a formula similar
to (3.9) could account for the Pickering series (1896), a series of lines describing the spectra
22 3. THE BOHR MODEL OF THE ATOM
of stars. Bohr argued that singly ionized helium atoms would have exactly the same spectrum
as that of the hydrogen, but the corresponding wavelengths would be four times shorter.
In his eort to explain such spectral lines, Bohr even went further to take into account the
contributions of the motion of both the electron and the nucleus about their centre of mass
to derive the so-called reduced mass. Using the reduced mass of the hydrogen and He
+
, Bohr
found that the ratio of the Rydberg constants for ionized helium and hydrogen is 4.00160,
compared to the value of 4.00163 obtained from laboratory experiments of Fowler in 1912. He
then predicted the lines to be due to the presence of singly ionized helium atoms, instead of
the hydrogen.
Although Bohrs remarkable achievement in explaining both emission and absorption lines
of the hydrogen and hydrogen-like atoms and in describing the shell structure of an atom
is of paramount importance in the development of atomic model using an approach of an
uneasy mixture of classical and quantum ideas, the Bohr model of the atom is fundamentally
incomplete. The model has, however, some diculties regarding with discrete circular orbits.
These orbits give a classical, deterministic property to the position of an orbiting electron
in that the electrons position at any time can thus be determined with a high accuracy. This
property is one in which it is not acceptable in the context of quantum mechanical model of
an atom, where the electrons exact position cannot be precisely determined but only with
some condence by a probabilistic value.
The above argument is in line with Heisenbergs uncertainty principle (which will be
discussed in Chapter 5), where it is argued that the electrons of a multi-electron atom occupy
a region like a cloud around the nucleus with the cloud density being the probability to nd
an electron in a particular region. As more electrons are introduced to a stable multi-electron
atom, the paths through which the electrons move around the nucleus are complicated, and are
likely to be overlapping in elliptical orbits, making the circular orbits are no longer applicable.
This implies that additional quantum numbers other than the principal quantum number n
are needed to describe the complex structure of a multi-electron atom. Detailed calculations
of quantized energy levels of such an atom reveal the presence of atomic sub-shells, associated
with orbital quantum numbers. The complete description of all quantum numbers required to
describe the dynamics of a system at atomic scales is given in the course of Quantum Physics.
3.5 Bohrs Correspondence Principle
Bohrs correspondence principle is a simple principle that provides a smooth and gradual
change of the new but in some sense yet primitive quantum theory into classical theory
in the limit of a very large quantum number previously dened. This principle naturally comes
from the basic idea that a new theory should be able to capture all the essence of the old law
3.5. Bohrs Correspondence Principle 23
Figure 3.6: A sketch showing the classical limit of the Bohr theory as the principal
quantum number approaches innity, associated with very large orbit radii (taken from
Ch.4, Modern Physics, Serway et al., 2005).
with no exception, and at the same time to demonstrate when the new theory approaches the
old one. If this requirement is satised, then the new theory is said to be rmly established
with no doubt. A similar situation can also be found in the context of special relativity theory.
This theory shows its strength not only by combining spatial and temporal components of
the four-dimensional space-time coordinates with the concept of the four-vector thoroughly
discussed in Lecture Notes on Modern Physics, but also by demonstrating that calculations of
the physical quantities based on the special relativity gradually change into classical results
of Newtonian mechanics when the speed is relatively small compared to the speed of light.
In the hands of Bohr, the correspondence principle becomes a master tool to demonstrate
where the quantization of the electrons orbital angular momentum comes from. Once derived,
the validity of the quantization concept depends upon the agreement with measurements of
lines of atomic spectra. Here, we simply use Figure 3.6 to derive the rst Bohr postulate
earlier mentioned in 3.2 from which the quantized electrons orbit radii shown in (3.5) and
the discretized electrons total energy in (3.7) are derived. Let us assume that r
1
and r
2
are
the radii of the two adjacent quantum orbits, each having its own orbital angular frequencies

1
and
2
, respectively. We go further by assuming that r
1
r
2
r and hereby
1

2
.
We also assume that

is the frequency of a photon emitted as a result of a quantum transition


from the initial state of radius r
1
to the nal state of radius r
2
.
Using the fact that the orbiting electron is under the inuence of the Coulomb force, we
can show that the total energy for the electron is given by
E =
mk
2
e
4
2L
2
(3.12)
24 3. THE BOHR MODEL OF THE ATOM
Note that we obtain the close relationship of the electrons energy and the angular momentum
for the Bohr model. Taking the derivative of (3.12), we have
dE
dL
=
mk
2
e
4
L
3
=
mk
2
e
4
(mk
2
e
4
)/
= (3.13)
or it can be simply written as
dE = dL (3.14)
where dE is the energy of a photon emitted when a quantum transition from r
1
to r
2
occurs
and dL is the corresponding change in the electrons angular momentum. Thus based on
Figure 3.6 and the denitions of dE and

, we have dE =

and so (3.14) becomes


= dL (3.15)
As earlier noted, the correspondence principle must be able to predict the same frequency
occurs for the emitted photon following the transition, as suggested by the quantum theory of
light emission, and for the electrically charged electron moving around the nucleus, based on
classical electromagnetic theory of radiation. In this way, it is true that

= , and therefore
we can write
dL = (3.16)
It is understood from (3.16) that the electrons orbital angular momentum in a specic orbit
is an integral multiple of , from which the rst postulate of Bohr L = n given in (3.1) is
proposed. What an amazing result ! Bohr realized that although (3.16) originally was derived
for the special case of orbits with large radii, it was a universal quantum principle that it was
relevant to a wider range of applicabilities than the classical Maxwells theory of radiation.
3.6 Exercises
1. If an electron moves from an inner orbit to an outer orbit, does its total energy increase
or decrease ? Does its kinetic energy increase or decrease ?
(taken from Ch.37, Physics for scientists and engineers, Tipler, 1999).
2. Show that the speed of an electron in the nth Bohr orbit of the hydrogen atom is given
by v
n
= e
2
/2
o
nh.
(taken from Ch.37, Physics for scientists and engineers, Tipler, 1999).
3. The binding energy of an electron is the minimum energy required to remove the electron
from its ground state to a large distance from the nucleus.
PR, UTS
3.6. Exercises 25
(a) What is the binding energy for the hydrogen atom ?
(b) What is the binding energy for He
+
?
(c) What is the binding energy for Li
2+
?
Note that a singly ionized helium He
+
and a doubly ionized lithium Li
2+
behave as
hydrogen-like atoms in that such ionized elements consist of a positively charged nucleus
and a single bound electron.
(taken from Ch.37, Physics for scientists and engineers, Tipler, 1999).
4. What is the radius of the rst Bohr orbit in He
+
and Li
2+
?
(taken from Ch.4, Modern Physics, Serway et al., 2005).
5. A hydrogen atom is in its ground state. Using the Bohr theory of the atom, calculate
(a) the radius of the orbit
(b) the linear momentum
(c) the angular momentum
(d) the kinetic energy
(e) the potential energy
(f) the total energy of the electron
(taken from Ch.4, Modern Physics, Serway et al., 2005).
6. A photon is emitted when a hydrogen atom undergoes a jump of electronic transition
from the initial quantum state of n = 3 to the nal quantum state of n = 2. Determine
(a) the energy
(b) the wavelength
(c) the frequency of the emitted photon
(taken from Ch.4, Modern Physics, Serway et al., 2005).
7. A hydrogen atom initially at rest in the n = 3 quantum state decays to the ground state
with the emission of a photon.
(a) Calculate the wavelength of the emitted photon
(b) Estimate the recoil momentum of the atom
(c) Estimate the kinetic energy of the recoiling atom
(d) Where does this energy come from ?
(taken from Ch.4, Modern Physics, Serway et al., 2005).
26 3. THE BOHR MODEL OF THE ATOM
8. If one assumes that in a stationary state of the hydrogen atom the electron ts into a
circular orbit with an integral number of wavelengths, one can reproduce the results of
the Bohr theory. Work this out.
(taken from Ch.1, Quantum Physics, Stephen Gasiorowicz, 1996).
9. Use the Bohr quantisation rules to calculate the energy levels for a harmonic oscillator,
for which the energy is p
2
/2m + m
2
r
2
/2 directly given by the driving force m
2
r.
Restrict yourself to circular orbits. What is the analog of the Rydberg formula ? Show
that the corresponding principle is satised for all values of the principal quantum
number n used in quantizing the angular momentum.
(taken from Ch.1, Quantum Physics, Stephen Gasiorowicz, 1996).
10. A muon is a particle with a charge equal to that of electron and a mass equal to 207
times the mass of an electron. Muonic lead is formed when Pb
208
captures a muon to
replace an electron. Assume that the muon moves in such a small orbit that it sees a
nuclear charge of Z = 82. According to the Bohr theory, what are the radius and energy
of the ground state of muonic lead ? Use the concept of reduced mass here.
(taken from Ch.4, Modern Physics, Serway et al., 2005).
Chapter 4
THE WAVE BEHAVIOUR OF
SUB-ATOMIC PARTICLES
As widely known, classical wave theory suggests that the propagation of light is considered
as a natural wave phenomenon. Examples of this can be found in many physical situations.
In particular, the laws of geometric optics are empirically proved to be held when light is
incident on a surface (reection) or arrives at a boundary between two media (refraction).
In other optical phenomena, such as interference and diraction of light, the laws of physical
optics are observed to occur when a beam of light rays passes through a very small aperture.
In Chapter 2, however, it was shown that electromagnetic radiation in the form of photons
may behave as particles when interacting with matter, such as those in the photo-electric and
Compton eects. A somewhat bizarre, fundamental question based on the reverse mechanism
is raised: can a classical particle, say an electron, with its own character behave as a wave
with a totally dierent character ? If this is the case, then the revolutionary idea about
matter waves that leads to the new concept of the way we are looking at the dynamics of
moving particles at microscopic scales has to be proven by both theoretical considerations and
experimental measurements. These topics are the primary issues we discuss in this chapter.
4.1 De Broglie Hypothesis
In classical physics, the concept of particle distinguishes both qualitatively and quantitatively
from that of wave. In the context of a classical particle, it is common to characterize particle
with its inertial mass; a particle is spatially localized at a certain time. Whereas a wave is
characterized by its spatial periodicity, i.e., wavelength, or temporal periodicity i.e., frequency
for which it is also dened as an energy propagated from one point to another in space-time
coordinates; the wave spreads and hence its energy distributes over space. It is clear from this
that none can demonstrate itself to be a particle and at the same time behaves as a wave.
27
28 4. THE WAVE BEHAVIOUR OF SUB-ATOMIC PARTICLES
Figure 4.1: The equivalence between (a) a moving particle of mass m and speed v
o
with (b) a wave packet of wavelength and speed v
g
(taken from Ch.5, Modern Physics,
Serway et al., 2005).
In 1925, de Broglie postulated a radical idea with no experimental supports at the time
it was proposed. His postulate was nally known as the de Broglie hypothesis derived
from theoretical considerations of classical wave, relativity theory, and quantum concept for
a photon with zero rest mass. The mathematical expression of his postulate is as follows,
=
h
p
(4.1)
where h is the Plancks constant, and p are the wavelength and momentum of the photon,
respectively. At this stage, de Broglie argued that the above expression holds also for any
moving object of mass m. The following paragraphs show logical reasons for this.
De Broglie generalized (4.1) to introduce the concept of a matter wave by assuming that
a moving particle can be in principle viewed as a wave packet or a wave group. A wave
of this type must reect the fact that such a particle has a large probability of being found
within a small, conned region of space at a limited time (see Figure 4.1). It follows that a
single traveling sinusoidal wave with constant amplitude and innite extent is not relevant to
model the particle. Instead, a group of waves of limited spatial extent consisting of individual
waves with dierent wavelengths can then represent the particle. In this case, the resulting
wave group travels at a speed v
g
dened to be the group velocity of the wave, which is
identical to the observed speed v
o
of the corresponding particle.
Let us then apply the de Broglie hypothesis given in (4.1) to a photon of energy E =
and momentum p = k. These quantities are attributed to a wave character through two
related parameters, namely the angular frequency = 2 and the wave number k = 2/.
4.1. De Broglie Hypothesis 29
From these quantities, or parameters, we have = ck for the photon. If the group velocity of
de Broglie wave corresponding to this photon is dened as
v
g
=
d
dk
=
dE
dp
(4.2)
then we have v
g
= c for the speed of the wave group, as it should be for photons with zero
rest mass, m
o
= 0.
As noted, we have already made (4.1) true for photons. An interesting question is that
whether such an equation also holds for a classical particle of mass m moving at speed v.
Here, we provide a somewhat crude argument for this. During its motion, the particle has
momentum p = mv and kinetic energy E = p
2
/2m. If the motion is considered as a wave
group, then the group velocity can be calculated by assuming d/dk to be equivalent to
dE/dp and by inserting this equivalence into the denition of v
g
to get the result for v
g
= v,
consistent with the particles speed. The result also implies that, for photons, the wave group
travels at speed c. Based on this simple calculation, it is shown that (4.1), originally derived
for photons, is relevant to any moving particle with non zero rest mass.
A more detailed use of (4.1) for any moving particle with non zero rest mass is given here,
as we are now in a position to apply the de Broglie hypothesis to sub-atomic particles, such
as a proton, a neutron, and an electron. In doing so, we need to dene what is called the
phase speed, that is, the speed of a point of constant phase on a wave. The phase speed of
the wave is given by
v
p
=

k
=
E
p
(4.3)
where E dan p represent the relativistic expressions for the total energy and momentum of the
particle, respectively. We can also write the phase speed in (4.3) as a function of k only with
the help of E
2
= E
2
o
+ (pc)
2
, where E
o
is the particles rest energy. Inserting this relativistic
relation into (4.3) results in
v
p
= c

1 +
_
mc
k
_
2
(4.4)
where m is the particles rest mass. As k varies with wavelength, or equivalently frequency,
then (4.4) is meant for individual waves of the wave packet. Note that the expression in (4.4)
is also called a dispersion relation for the phase speed of each component of the wave group.
The group velocity of the wave group can then be calculated from
v
g
= v
p
+ k
dv
p
dk
(4.5)
to be evaluated at k
o
, the central wave number of a continuous distribution of wavelengths
constituting the wave group.
30 4. THE WAVE BEHAVIOUR OF SUB-ATOMIC PARTICLES
Substituting (4.4) to (4.5) and after some simple algebra, we can write the group velocity
of the de Broglie wave as
v
g
=
c
_
1 + (mc/k)
2
=
c
2
v
p
(4.6)
Let us go back for a moment to the phase speed given in (4.3). Solving for this phase speed
yields
v
p
=
E
p
=
mc
2
mv
=
c
2
v
(4.7)
where is dened as
=
1
_
1 v
2
/c
2
(4.8)
and v is the particles speed. We eventually get the right expression for the group velocity
by inserting (4.7) into (4.6) to have v
g
= v. The nal result obtained for v
g
is convincing in
that the group velocity of the wave packet is the same as the corresponding particles speed,
as expected.
4.2 Implications of the De Broglie Hypothesis
Although it looks so simple, the de Broglie hypothesis dened in (4.1) is actually powerful.
The hypothesis combines well fundamental aspects of both classical wave of wavelength
and classical particle of momentum p. The Plancks constant h serves as a connecting bridge
between the two properties, implying that the application of the hypothesis is limited only
to the case of moving particles at a microscopic scale. In this context, the hypothesis is then
used to validate the particle property of a photon and the quantization of the orbital angular
momentum of the electron for the Bohr model of the hydrogen atom. These two implications
will be discussed in the following paragraphs.
It has been known that photons have zero mass, will always be moving at speed c, and may
behave as particles. This particle property, however, would have made photons to have mass.
This seems to be contradictory with the fact that photons are massless. To examine this, let
us dene what is called the eective inertial mass of a photon, a quantity describing how
a photon responds to an applied force acting on it. The photons eective inertial mass m
e
may reasonably be taken to be proportional to its total relativistic energy E as follows,
m
e
=
E
c
2
=
h
c
2
=
h
c
(4.9)
for which (4.9) can be rearranged to become
=
h
m
e
c
=
h
p
e
(4.10)
4.2. Implications of the De Broglie Hypothesis 31
Figure 4.2: De Broglie standing waves in Bohrs stationary orbits (taken from Ch.4,
Konsep Fisika Modern, The Houw Liong, 1987, adapted from the work of Arthur Besier).
which is identical to (4.1) in the sense that the associated wavelength of a moving particle
is inversely proportional to its momentum, where m
e
c is the photons eective momentum.
By (4.9) and (4.10) the de Broglie hypothesis demonstrates its-self to be self consistent with
both relativity and quantum theories.
Another important aspect of the hypothesis is that it provides a physical feature of the
Bohrs atom theory. Although the Bohrs model is useful to describe the dynamics of the
atoms, it has also some shortcomings regarding with, say for example, only certain values of
electronic energy are allowed to occupy in the model. De Broglie recognized this problem and
made it clear by visualizing the orbiting electrons around the nucleus as standing waves bent
into circles of discrete radii (see Figure 4.2). This point of view actually comes from the simple
but brilliant idea that an atom connes its orbiting electrons to a very small atomic dimension,
in which case the wave nature of the electrons should predominate over their particle property.
From classical wave theory we know that when a wave is restricted to a small area, then only
a discrete set of standing waves are possible to occur within that area. The Bohrs stationary
orbits are thus viewed as a result of constructive interference of these waves for a range of
32 4. THE WAVE BEHAVIOUR OF SUB-ATOMIC PARTICLES
wavelengths. This constructive interference ts into the circumference of circular orbits and
corresponds to an integral number of the de Broglie waves, each having wavelength . We can
then express this de Broglies argument by rewriting the rst Bohrs postulate given in (3.1)
for the quantized orbital angular momentum L = n of the hydrogens electron to start from.
After a simple algebra we have
2r
n
= n for n = 1, 2, 3, ... (4.11)
where r
n
is the Bohrs orbit radii dened in (3.5). The fact that the de Broglie hypothesis
originating from the wave nature of the electrons revolving the nucleus matches with the Bohrs
theory is a key to understanding the nature of microscopic world. Following this, scientists
have started realizing that electrons and hence other sub-stomic particles have a dual property,
the so called particle-wave duality. It can then be inferred from many cases considered that
it is not possible to observe both the particle and wave properties simultaneously. Rather, one
property completes the other for which Bohr called this as a complementary principle.
4.3 The Davisson-Germer Experiment
Following his successful doctoral degree, de Broglie as a consequence of his postulate suggested
that a stream of electrons passing through a very small aperture would produce diraction
pattern, as it would for a beam of light. In this context, a series of laboratory experiments
using Davisson-Germer experimental apparatus (see Figure 4.3) were completed in 1927 to
test the de Broglie hypothesis of a matter wave. The primary result of these experiments
was that electrons experience diraction, providing convincing support for the wave nature
of an electron. The apparatus allows for the variations of three experimental parameters,
namely the energy of an electron, the orientation of a nickel target, and the angle of an
elastic scattering. Initially, the progress of these experiments seemed to give nothing. After
an unexpected but fortunate experimental accident occurred, however, Davisson and Germer
realized that further analyses on the experimental procedure give dierent results that exhibit
a diraction pattern. Thus, they took this peculiar chance to calculate the wavelength of an
electron using both the de Broglie hypothesis given in (4.1) and a simple diraction formula,
corresponding to various values of the three parameters. In their theoretical calculation on
the basis of conservation of energy, Davisson and Germer used the non-relativistic expression
for the speed v of the electron as follows,
1
2
mv
2
= P (4.12)
where m = 9.11 10
31
kg is the electrons rest mass and P is the potential energy supply.
4.4. Exercises 33
Figure 4.3: The Davisson-Germer experimental apparatus, used for examining the
wave nature of an electron by elastically scattering a beam of low-speed electrons from
a polycrystalline nickel target (taken from Ch.5, Modern Physics, Serway et al., 2005).
The remaining step is straight forward, solving for v from (4.12) and inserting the result
into (4.1) yields
=
h

2meV
(4.13)
where e = 1.6 10
19
C is the electron charge and V is the experimental voltage supply. For
the special case of corresponding to the diraction maximum where V = 54 volt, we have
= 1.67 10
10
m, in good agreement with = 1.65 10
10
m obtained from formula for
constructive interference,
d sin = n (4.14)
where d = 2.15 10
10
m obtained from previous measurements of X-ray diraction, = 50

and n = 1 corresponds to the diraction maximum pattern.


4.4 Exercises
1. Is light a wave or a particle ? Is an electron a particle or a wave ? Support your answer
by citing specic experimental evidence.
(taken from Ch.5, Modern Physics, Serway et al., 2005).
34 4. THE WAVE BEHAVIOUR OF SUB-ATOMIC PARTICLES
2. If matter has a wave nature, why is this wave-like character not observable in our daily
experiences ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
3. Show that the de Broglie wavelength of an electron of kinetic energy E, measured in eV,
is given by

e
=
12.3 10
10
E
1/2
m
and that of a proton is given by

p
=
0.29 10
10
E
1/2
m
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
4. Show that in order to associate a de Broglie wavelength with the propagation of photons
(electromagnetic radiation), photons must travel with the speed of light c and their rest
mass must be zero. Do this relativistically.
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
5. Calculate the de Broglie wavelength for
(a) an electron of kinetic energy 250 eV (the rest energy of the electron is 0.511 MeV)
(b) a neutron of kinetic energy 0.02 eV (the rest energy of the neutron is 940 MeV)
(c) a proton of kinetic energy 2 MeV (the rest energy of the proton is 938 MeV)
(taken from Ch.17, Physics for scientists and engineers, Tipler, 1999).
6. To observe small objects, one measures the diraction of particles whose de Broglie
wavelength is approximately equal to the objects size. Find the kinetic energy, measured
in eV, required for electrons to resolve
(a) a large organic molecule of size 10 nm
(b) atomic features of size 0.10 nm
(c) a nucleus of size 10 fm
Repeat these calculations using alpha particles in place of electrons.
(taken from Ch.5, Modern Physics, Serway et al., 2005).
7. Find the de Broglie wavelength of a ball of mass 0.2 kg just before it strikes the Earth
after being dropped from a building 50 m tall.
(taken from Ch.5, Modern Physics, Serway et al., 2005).
4.4. Exercises 35
8. An electron has a de Broglie wavelength equal to the diameter of the hydrogen atom.
What is the kinetic energy of the electron ? How does this kinetic energy compare with
the ground-state energy of the hydrogen atom ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
9. For an electron to be conned to a nucleus, its de Broglie wavelength would have to be
less than 10
14
m.
(a) What would be the kinetic energy of an electron conned to this region ?
(b) On the basis of this result, would you expect to nd an electron in a nucleus ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
10. The dispersion relation for free electron waves is
(k) =
_
c
2
k
2
+ (mc
2
/)
2
From the above equation, obtain expressions for the phase velocity v
p
and the group
velocity v
g
of these waves and show that their product is constant, independent of k.
From your results, what can you learn about v
g
when v
p
> c ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
36 4. THE WAVE BEHAVIOUR OF SUB-ATOMIC PARTICLES
Chapter 5
THE HEISENBERG UNCERTAINTY
PRINCIPLE
One of the most basic and far-reaching concepts of modern physics is Uncertainty Principle,
proposed by Werner Heisenberg in 1927. This beautiful principle is fundamentally concerned
with the limit of our ability in simultaneously determining with high accuracy and precision
two independent but conjugate variables in physics, which can be in the form of a pair of two
quantities, such as position and momentum, energy and time, angle and angular momentum.
The limit itself has nothing to do with imperfections in practical measuring instruments, or
equivalently with careless measurements. Rather, it arises automatically from the nature of
microscopic systems. Although this principle in its simple form is later known to be one of
cornerstones in modern physics, Heisenberg is also famous for his contributions to develop a
complete theory of quantum mechanics. Indeed, modern quantum theory was pioneered
by the works of Schrodinger and Heisenberg, and some other physicists in the late of 1920s.
Apparently, there were two dierent quantum theories, namely wave mechanics proposed
by Schr odinger and matrix mechanics suggested by Heisenberg. Although the latter is quite
elegant, it only paid little attention to physicists at the time it was launched for some reasons.
The matrix mechanics involved an unfamiliar way of complicated mathematics to describe the
dynamics of sub-atomic particles, and was arguably based on rather vague physical concepts.
However, these two formulations were later shown to be completely equivalent, in the sense
that both theoretical approaches conrm that the principal quantum number is not adequate
to describe the dynamics of microscopic particles. Instead, additional parameters associated
with orbital, magnetic, and intrinsic quantum numbers are needed to complete an atomic
description. A detailed derivation of these quantum numbers will be given in the course of
Quantum Physics. Instead, this chapter focuses on the Heisenberg uncertainty principle,
providing a mathematical basis for the principle and exploring its consequences.
37
38 5. THE HEISENBERG UNCERTAINTY PRINCIPLE
x
f(x)
-L
0
Eo
L
A(k)
0
(a)
(b)
kp -kp
k
kp+/L kp-/L
EoL
Figure 5.1: The idealised prole of (a) a cosine pulse of width 2L and amplitude E
o
,
and (b) its Fourier transform with a carrier spatial frequency k
p
in the space domain.
5.1 Mathematical Basis for the Uncertainty Principle
In the previous chapter, it was stated that what was meant by a wave packet or a wave group
associated with a moving particle was a group of waves of limited spatial extent consisting of
individual waves of dierent frequencies. Wave of this type is commonly found in the form
of a pulse propagated in space. The term space here is used in a general sense, for which it
could be either x-axis or t-axis depending upon an explicit mathematical function of the pulse.
To examine how a pulse is transformed from one form to another in a particular domain, here
we consider an idealised harmonic pulse in the form of a cosine periodical function, that is,
the cosine wavetrain in the space domain (see Figure 5.1a) given by
f(x) = E
o
cos k
p
x for L x L
= 0 otherwise
(5.1)
The choice to work in the space domain is optional in that the time-dependent disturbance
is equivalently applicable (see also Figure 5.2). In the time domain, the corresponding cosine
wavetrain is written as
g(t) = E
o
cos
p
t for T x T
= 0 otherwise
(5.2)
Here we solve the problem using spatial variables in (5.1) than using temporal variables
in (5.2), as it is easier to examine the wave prole at t = 0 rather than the prole at x = 0.
5.1. Mathematical Basis for the Uncertainty Principle 39
t
g(t)
-T
0
Eo
T
a()
0
(a)
(b)
p -p

p+/T p-/T
EoL
Figure 5.2: The idealised prole of (a) a cosine pulse of width 2T and amplitude E
o
,
and (b) its Fourier transform with a carrier temporal frequency
p
in the time domain.
As f(x) given in (5.1) is an even function, the corresponding cosine Fourier transform
A(k) in the space domain is calculated from
A(k) =
_

f(x) cos kx dx (5.3)


The above integral tells us about how dynamic variables k and x are interchangeable, meaning
that a function of a particular variable in a given domain can be transformed into another
function of its counter-part variable in the same domain. Thus based on Figure 5.1 and by
substituting (5.1) into (5.3), we have
A(k) =
_
L
L
E
o
cos k
p
x cos kx dx
=
_
L
L
E
o
2
_
cos(k
p
+ k)x + cos(k
p
k)x
_
dx
= E
o
L
_
sin(k
p
+ k)L
(k
p
+ k)L
+
sin(k
p
k)L
(k
p
k)L
_
= E
o
L
_
sinc (k
p
+ k)L + sinc (k
p
k)L

(5.4)
Several interesting possibilities are here discussed regarding with (5.4). When there are
many waves in the train, that is,
p
L, then the left and right peaks are both narrow
and widely spaced. When k = k
p
, the second sinc function reaches a maximum value of one,
meaning that A(k) has a maximum value of E
o
L (Figure 5.1b). In particular, the rst sinc
40 5. THE HEISENBERG UNCERTAINTY PRINCIPLE
function in the bracket becomes unimportant hence negligible compared with the second one.
Thus, the Fourier transform of the train becomes
A(k) = E
0
Lsinc (k
p
k)L (5.5)
which is the right-hand side curve of Figure 5.1b. In the case of the time-dependent cosine
wavetrain, the corresponding Fourier transform is given by
a() = E
0
T sinc (
p
)T (5.6)
which is the right-hand side curve of Figure 5.2b. Here, k and are correlated to one another
by the phase velocity. As the train becomes innitely long, that is, L or accordingly
T , the curve closes down to a single tall spike at k
p
or again accordingly
p
.
Notice that the square of A(k) or a() represents the energy density of the wave packet.
Most of this energy is carried away by the train, associated with part of the curve with a
spatial frequency k ranging from k
p
/L to k
p
+/L (see Figure 5.1b). A longer train leads
to the energy to be concentrated in a narrower range of k about k
p
. If the train were innitely
long, there would be singularity in the Fourier transform, in sense that all the energy would
be focused on a single value of frequency at k
p
or correspondingly
p
. This is the perfectly
limiting case of the idealised monochromatic wave.
In real cases, however, there have been no innitely long wavetrain such that no singularity,
hence no single value of frequency in the Fourier transform. Thus, it is sensible to calculate the
width of either spatial frequency k or temporal frequency , in which range the energy of
a given pulse is transferred. This is clearly illustrated by the right curves in Figures 5.1b and
5.2b. For the space-dependent pulse, the spatial frequency bandwidth k is calculated
from
k = k
2
k
1
=
_
k
p
+

L
_

_
k
p


L
_
=
2
L
(5.7)
and equivalently for the time-dependent pulse, the temporal frequency bandwidth is
given by
=
2

1
=
_

p
+

T
_

_


T
_
=
2
T
(5.8)
It is also clear from Figures 5.1a and 5.2a that the spatial extent x or temporal extent t
of the pulse is xed at a value of x = 2L or t = 2T, respectively. The product of these
frequency widths and their extents of the associated pulses are therefore given by
k x =
2
L
2L = 4
t =
2
T
2T = 4
(5.9)
5.2. Interpretations of the Uncertainty Principle 41
It follows that if the wave packet has a narrow frequency, spatial or temporal, bandwidth then
the packet will spread out in a large region of space or time. In quantum mechanics, a wave
packet or group represents a moving microscopic particle having nite energy and momentum.
If the narrow frequency bandwidth is associated with a denite value of energy or momentum,
then the exact location of the particle in space-time coordinates becomes highly uncertain.
Note that the precise value of the number on the RHS of (5.9) relies upon the functional
form of f(x) of the pulse, the geometrical shape chosen for A(k), and on the specic denitions
of k and x. A dierent choice of f(x) or A(k) and/or a dierent rule for dening k and
x will give a slightly dierent number. For the case of a Gaussian wave packet and use of
standard deviations for the denitions of both k and x, or correspondingly and t,
we have
k x
1
2
and t
1
2
(5.10)
for which it can be easily shown that
p x

2
and E t

2
(5.11)
In words, the uncertainty relations above state that the more precisely the momentum (energy)
is determined the less precisely the position (time) is known, and conversely. Here at this stage,
two equivalent questions are addressed : what do the uncertainty relations tell us, anyway ?
and how do we interpret them ? These issues will be discussed in 5.2.
But before doing so, it is better here to provide a three-dimensional case for the rst
formulation of the uncertainty principle in (5.11) by rewriting it as
p
x
x

2
p
y
y

2
p
z
z

2
(5.12)
While the momentum-position uncertainty relation given in (5.12) is relatively clear,
the energy-time uncertainty relation in the second formulation of (5.11) conrms that
the precision with which we are able to determine accurately the energy of a given system is
limited by the time taken for measuring the energy of such a system.
5.2 Interpretations of the Uncertainty Principle
As earlier mentioned, the term uncertainty may refer to either the uncertainty resulting from
a series of experimental measurements, meaning that it is an error in the measurements, or
the uncertainty derived from conceptual framework. The dierence in the meaning between
the two is that the uncertainty in the latter case arises from the product of uncertainties of
dierent variables (not from the uncertainty in any single quantity). Let us suppose here
42 5. THE HEISENBERG UNCERTAINTY PRINCIPLE
that the probability distributions for two physical quantities are not independent; they are
constrained such that both distributions cannot be made arbitrarily narrow. For example, it
could be the case that when one quantum distribution is made narrow, then the other must
be relatively wide so that the product of the two uncertainties satises some lower bound.
This is the truly essence of the quantum-mechanical uncertainty relations.
Let us here discuss further consequences of the quantum-mechanical uncertainty relations.
Take the rst formulation of (5.11) for a classical moving particle, where its momentum and
position are both independent and well dened in space-time coordinates. There is no doubt to
say that the particles position does not correspond to its momentum, and vice versa, and that
the two quantities can be specied to arbitrary precision. Therefore, it is possible to determine
precisely these quantities at any time in space once the velocity of the particle is known. This
behaviour is considered as a basic principle of determinism in classical mechanics. However,
the quantum-mechanical uncertainty relations state that quantum probability distributions
cannot be made independent. It is, for example, impossible for a microscopic particle that is
represented by a quantum state or a wavefunction to have a nite probability distribution
in position and at the same time no spread in its momentum, or conversely. What we can
only say is that the probability to nd such a particle in its particular state is given by
2
integrated over the whole space. According to quantum mechanics, it is thus meaningless to
get one dynamic variable precisely determined, as its corresponding conjugate variable must
be highly uncertain at the same time. This fundamental consequence makes the nature of
quantum measurements for microscopic world to be undeterministic, as opposed to classical
mechanics. Based on this basic physics principle, all measurements of physical quantity can
only be determined with some condence. This has further implications for the Bohrs theory,
which will be discussed below.
As previously discussed in Chapter 3, the Bohr model of the hydrogen atom leads to
discretized energy levels that agree well with spectroscopic data. Despite its usefulness, this
model has also some shortcomings for the following reasons. In fact, there is no basis theory for
justication of the well-dened stationary orbits. The nature of orbiting electrons is dicult
to predict as the exact location of such electrons cannot be precisely determined, as suggested
by the uncertainty principle. Rather, the electrons occupy a region like cloud surrounding
the nucleus, with the cloud density being the probability that these electrons could be found
in that region. As a consequence, the orbiting electrons may have overlapping states, making
the Bohrs circular orbits no longer applicable.
For concluding remarks, it is argued that the Heisenberg uncertainty relations in (5.11)
in their original forms are therefore not just about mathematical results, as they put forward
profound physical and philosophical implications. Indeed, the uncertainty relations remain an
important element in the structure of modern quantum mechanics to date.
5.3. Exercises 43
5.3 Exercises
1. Atomic nuclei, typically of size 10
14
m, frequently emit electrons, with energies of
about 1 MeV or less. Use the uncertainty principle to show that hypothetical electrons
trapped electrons within a conned region having an energy of order 20 MeV could
not be contained in the nucleus before the decay.
(taken from Ch.2, Quantum Physics, Stephen Gasiorowicz, 1996).
2. Use the Heisenberg uncertainty principle to estimate the ground state energy of a one
dimensional harmonic oscillator. The total energy for the oscillator is given by
E =
p
2
2m
+
1
2
m
2
x
2
(taken from Ch.2, Quantum Physics, Stephen Gasiorowicz, 1996).
3. Monochromatic light with wavelength of 6000

A passes through a fast shutter that
opens for only 10
9
second. What will be the spread in wavelengths in the no longer
monochromatic light ?
(taken from Ch.2, Quantum Physics, Stephen Gasiorowicz, 1996).
4. The size of an atom is approximately 10
8
cm. To locate an electron within the atom,
one should use electromagnetic radiation of wavelength not longer than, say, 10
9
cm.
(a) What is the energy of a photon with such a wavelength (in eV) ?
(b) What is the uncertainty in the electrons momentum if we are uncertain about its
position by 10
9
cm ?
(taken from Ch.2, Introductory Quantum Mechanics, Richard Libo, 1980).
5. A proton has a kinetic energy of 1 MeV. If its momentum is measured with an uncertainty
of 5%, what is the minimum uncertainty in its position ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
6. We wish to measure simultaneously the wavelength and position of a photon. Assume
that the measurement gives = 6000

A with an accuracy of one part in a million, that


is, / = 10
6
. What is the minimum uncertainty in the position of the photon ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
7. A beam of electrons is incident on a slit of variable width. If it is possible to resolve a 1%
dierence in momentum, what slit width would be necessary to resolve the interference
pattern of the electrons if their kinetic energy is
44 5. THE HEISENBERG UNCERTAINTY PRINCIPLE
(a) 0.01 MeV ?
(b) 1.0 MeV ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
8. An excited nucleus with a lifetime of 0.1 ns emits a ray of energy 2 MeV. Can the
energy width (uncertainty in energy is E) of this 2 MeV emission line be directly
measured if the best gamma detectors can measure energies to 5 eV ?
(taken from Ch.5, Modern Physics, Serway et al., 2005).
9. A
o
meson is an unstable particle produced in high energy particle collisions. It has a
mass-energy equivalent of about 135 MeV, and it exists for an average lifetime of only
8.7 10
17
second before decaying into two rays. Using the uncertainty principle,
estimate the fractional uncertainty m/m in its mass determination.
(taken from Ch.5, Modern Physics, Serway et al., 2005).
10. One of the cornerstones of quantum mechanics is that bound particles cannot be in
stationary states even at zero absolute temperature. A bound particle is one that
is conned in some nite region of space, as is an atom in a solid. There is a non-zero
lower limit on the kinetic energy of such a particle. Suppose a particle is conned in
one dimension to a region of width L. Obtain an approximate formula for its minimum
kinetic energy.
(taken from Ch.4, Modern Physics, Randy Harris, 2007).
Chapter 6
THE BASIC PHYSICS PRINCIPLES
OF QUANTUM MECHANICS
The preceding chapters dealed with the nature of a photon as both a wave and a particle.
Chapter 2 provided the experimental ndings that support the particle nature of photons.
However, the concept of quantisation had not been introduced till the great work of Bohr in
accounting for the discrete emission spectra of hydrogen atom was introduced. In this context,
Chapter 3 discussed the Bohrs model of hydrogen atom what then becomes the primitive
theory of quantum mechanics. In Chapter 4, we learned from the de Broglie hypothesis
that particles could behave as waves and that consequently, a brilliant insight of matter-wave
duality was born. Following this idea, the Heisenberg uncertainty principle was then discussed
in Chapter 5 as a cornerstone of natural law that distinguishes the probabilistic principle of
quantum mechanics from the deterministic principle of classical physics.
In this chapter, we examine formal formulation of quantum mechanics using mathematical
approaches developed for describing the dynamics of extremely small objects usually found
in microscopic world. The formulation is then grouped into several postulates of quantum
mechanics. These postulates serve as the basic physics principles of quantum mechanics
to which all dynamical problems of microscopic particles are referred. The aim of this chapter
is thus to introduce the fundamental postulates and to examine their possible applications to
simple problems.
6.1 Observables, Operators, and Wavefunctions
A wavefunction consists of all information about the state of a particle, hence it is also called
as a statefunction. To be valid, a particular wavefunction (r, t) must be well-behaved,
meaning that (r, t) must be single valued, continuous, dierentiable, and square-integrable
45
46 6. THE BASIC PHYSICS PRINCIPLES OF QUANTUM MECHANICS
as follows,
_

[ (r, t) [
2
dV < (6.1)
It means that the integral of the squared wavefunction over volume space above must be nite.
In general, (r, t) represents a physical state of a given system, where a well-dened physical
observable A, such as energy, momentum and position of the system will be then determined.
The wavefunction can be a complex function with

(r, t) is the complex conjugate of (r, t).


The rst postulate of quantum mechanics states that for each physical observable A there
corresponds a dynamic operator

A such that a quantum measurement of A yields an eigenvalue
a associated with

A. Thus, this postulate holds observables, operators, eigenvalues, and
wavefunctions together in a compact form, namely an eigenvalue equation as follows,

A = a (6.2)
Here, observables are measurable quantities (both scalars and vectors) and operators are
in the form of mathematical operators, such as dierential operators that correspond to the
observables. An operator is actually an instruction for us to do something to the wavefunction
that follows. Eigenvalues associated with a particular operator must be real in order to have a
physical meaning. As seen in (6.2) above, the eigenvalue equation requires an operator. Two
dynamic operators of great interest are linear momentum and energy operators, constituting
the linear momentum operator p = i and the energy operator

E = i /t. It is necessary
to mention here that the linear momentum operator can be separated into its components,
that is, p
x
= i /x, p
y
= i /y, and p
z
= i /z.
Consider now a free particle of mass m moving in the x-direction with its associated
wavefunction denoted by (x, t) = e
i(kxt)
. This wavefunction is such that [ (x, t) [
2
dx is
the probability of nding the particle with its linear momentum p
x
in the interval between x
and x + dx. The associated eigenvalue equations for both particles momentum and energy
are as follows,
i

x
e
i(kxt)
= k e
i(kxt)
=
h

e
i(kxt)
p
x
= p
x

(6.3)
and
i

t
e
i(kxt)
= e
i(kxt)
= h e
i(kxt)

E = E
(6.4)
It is clear that both equations satisfy (6.2); quantum operators working on a wavefunction
6.2. Born Interpretation 47
result in associated eigen values and the wavefunction itself. It is also necessary to say that
since a free particle has only kinetic energy, (6.4) can be re-written as
p
2
x
2m
=

2
k
2
2m

(6.5)
It follows that if the free particle is in the state denoted by the statefunction
k
= e
i(kxt)
,
quantum measurements of momentum and energy of that particle will denitely yield k and

2
k
2
/2m, respectively.
6.2 Born Interpretation
The basic principle of the Heisenberg uncertainty relations introduced to the dynamics of
microscopic objects limit the way we determine the exact location of a particular particle if the
particles momentum is known with a high accuracy. Conversely, when the particles position
can be precisely localised, then its momentum is highly undetermined. If (r, t) denotes
a physical state, such as a free particle moving in a three-dimensional space of Cartesian
coordinates with a position vector in the space given by r = x x +y y +z z, then [ (r, t) [
2
is
a probability density and accordingly, an integral of [ (r, t) [
2
dx dy dz is the probability
for nding the particle in the interval between x = and x = +, y = and y = +,
and z = and z = +.
Born suggested that the total probability of nding a particle in its statefunction given by
(r, t) in a given volume V must be equal to unity. This can be expressed as
P =
_

[ (r, t) [
2
dV = 1 (6.6)
This condition means that the particle is located somewhere in space with certainty, or in
other words it exists in the given volume. Note that (r, t) is in general a complex function,
hence the absolute value of it in the integrand of (6.6) is necessary. In a one-dimensional
notation, (6.6) can then be re-written as
P =
_

[ (x, t) [
2
dx = 1 (6.7)
However, it is possible to nd that the probability of nding the particle in its state given
by (x, t) in a location bounded by x = a and x = b (in which both a and b are not the
endwalls of the boundary) is less than or equal to one,
P =
_
b
a
[ (x, t) [
2
dx 1 for a x b (6.8)
48 6. THE BASIC PHYSICS PRINCIPLES OF QUANTUM MECHANICS
6.3 Normalisation Procedure
In general, a wavefunction that describes a certain physical state contains an unkown constant
the so called a normalisation constant, where its value is determined by normalising the
wavefunction through normalisation procedure. This procedure simply applies either (6.6)
or (6.7) to a given statefunction (x, t) = Bf(x, t), where B is the normalisation constant
and f(x, t) is a particular function such that satises an eigenvalue equation of (6.2). Thus,
the normalisation procedure for this wavefunction is
1 = B
2
_

[ f(x, t) ]
2
dx (6.9)
The second postulate of quantum mechanics states that a measurement of an observable A
through operating

A on a statefunction yields an eigenvalue a that leaves the system in the
state
n
, where n is a quantum number. This can be mathematically written as

A
n
= a
n
(6.10)
where
n
is called as the eigenfunction, with B is the undetermined normalisation constant.
When B is already specied through the normalisation procedure (6.9), then
n
is called as
the normalised eigenfunction. Once the wavefunction is normalised, it stays normalised
for all future time, and hence the integral of (6.7) is independent of time,
dP
dt
=
d
dt
_

[ (x, t) [
2
dx = 0 (6.11)
Here, we provide a simple example of how to calculate a normalisation constant B and to
nd the associated normalised wavefunction. Consider a wavefunction of a particular state
given by (x) = Bsin(x/b), where 0 x b. Applying (6.9) results in
1 = B
2
_
_
0

sin
2
(x/b) dx +
_
b
0
sin
2
(x/b) dx +
_
+
b
sin
2
(x/b) dx
_
(6.12)
For this boundary condition, it is understood that the wavefunction vanishes outside of the
given interval. Therefore, we only need to calculate the second term on the right-hand side of
(6.12), which can be expanded to be
1 =
B
2
2
_
_
b
0
dx
_
b
0
cos(2x/b) dx
_
(6.13)
The second term on the right-hand side of (6.13) is zero, for which the normalisation constant
6.4. Hilbert Space and Superposition Principle 49
B is derived to be
B =
_
2/b (6.14)
The associated normalised wavefunction is thus
(x) =
_
2/b sin(x/b) (6.15)
We can now calculate the probability where the particle can be found in the interval between
b/2 x 3b/4 using (6.8) as follows,
P =
2
b
_
3b/4
b/2
1
2
_
1 cos(2x/b)
_
dx =
1
b
_
_
3b/4
b/2
dx
_
3b/4
b/2
cos(2x/b) dx
_
= 1/4 + 1/2 = 0.41
(6.16)
The total probability within the intervals 0 x b/2 and 3b/4 x b is then unity.
6.4 Hilbert Space and Superposition Principle
We introduce the concept of a space of functions in quantum mechanics, the so-called a
Hilbert space in which unlimited numbers of functions exist to serve as basis vectors for a
specied wavefunction (also commonly called as a statevector). In Cartesian coordinates,
each point is described by a position vector r = x x+y y+z z, with x, y, and z being unit or
basis vectors that span the vector space in the Cartesian system. The Hilbert space is much
the same as this system except that the basis vector that span the Hilbert space is unlimited
such that a statevector (x) can be written as a linear superposition of basis vectors
i
,
(x) =

i=1
c
i

i
= c
1

1
+ c
2

2
+ . . . . . . (6.17)
where c
i
are the superposition coecients that may be complex.
The third postulate relates the modulus squared [ c
i
[
2
to the probability T
i
of nding the
system in a particular state denoted by
i
. Thus, it can be written
P
i
= [ c
i
[
2
(6.18)
for which
P =

i
P
i
=

i
[ c
i
[
2
= 1 (6.19)
as required for the total probability (see selected examples of 2.4 and 2.12 of McMahon, 2005).
50 6. THE BASIC PHYSICS PRINCIPLES OF QUANTUM MECHANICS
6.5 Expectation Values
The fourth postulate of quantum mechanics describes the way we nd the average value of
repeated measurements of an observable A relevant to an ensemble of identically prepared
systems represented by a statefunction . The average is known as the expectation value,
dened as
A =
_


Adx (6.20)
where

A is the associated operator for the observable A and

is the complex conjugate of .


Note that the expectation value is not the average of repeated measurements on one and
the same system. It is not necessary a value that could be actually measured. Hence, it may
not turn out to be equal to one of the eigenvalues associated with

A. The physical meaning
of the expectation value above involves a large number N of replica experiments in which
individual measurements yield a set of values A
i
(i = 1, 2, 3, . . . A
N
). The average of A is then
calculated from
A =
1
N

i
A
i
(6.21)
The expectation value given in (6.20) will be especially true if the deviation of measured values
of A from the mean value A is not large. A measure of the spread about the mean value is
given by the variance or mean-square deviation, dened as
(A)
2
= A
2
A
2
(6.22)
The above relation is used to prove the Heisenberg uncertainty principle. It is good for exercise
to work out the following problems taken from McMahon, 2005: 2.15, 2.16, 2.18 and 2.22.
6.6 Exercises
1. Consider a particle whose normalised wavefunction depends on position only,
(x) = 2

x e
x
x > 0
= 0 x < 0
where is an arbitrary constant. Find the probability that the particle is found between
(a) x = and x = 0
(b) x = 0 and x =
(c) x = 0 and x = 1/
(taken from Ch.3, Quantum Physics, Stephen Gasiorowicz, 1996).
6.6. Exercises 51
2. Consider a particle of mass m trapped in a one-dimensional, innite box of width a with
a potential surrounding the box described by
V (x) = x < 0
= 0 0 x a
= x > a
The boundary conditions at the walls and outside the box applied to this particle is
(x) = 0 x 0
= 0 x a
Find for the particle in the box,
(a) the possible eigenvalues of energy
(b) the normalised constant
(c) the allowed normalised eigenfunctions
Give comments on your results (see Example 2.4, Quantum Mechanics demystied,
McMahon, 2005 and also the eigenvalue problem in Ch.4, Quantum Physics, Stephen
Gasiorowicz, 1996).
3. Consider again a particle of mass m trapped inside the box described in Problem 2
above. The corresponding wavefunction is found to be
(x) =
i
2
_
2
a
sin
x
a
+
_
1
a
sin
3x
a

1
2
_
2
a
sin
4x
a
If energy is measured,
(a) what are the possible results?
(b) what is the probability of obtaining each result?
(c) what is the most probable energy for this state?
(d) what is the mean energy for this system?
(see Examples 2.12 and 2.18, Quantum Mechanics demystied, McMahon, 2005).
4. A particle of mass m is located in a one-dimensional box bounded by interval 0 x a
with no potential inside the box. The coresponding wavefunction is found to be
(x) =
1

10a
sin
x
a
+ A
_
2
a
sin
2x
a
+
3

5a
sin
3x
a
52 6. THE BASIC PHYSICS PRINCIPLES OF QUANTUM MECHANICS
(a) Find A so that (x) is normalised.
(b) What are the possible results of measurements of the energy?
(c) What are the respective probabilities of obtaining each result?
(d) The energy is measured and found to be 2
2

2
/ma
2
. What is the state of the system
immediately after measurements?
(see Example 2.13, Quantum Mechanics demystied, McMahon, 2005).
5. A particle of mass m is located in a one-dimensional box of 0 x a with no potential
inside the box. If the particle is in the ground state,
(a) nd x and x
2

(b) nd p and p
2

(c) nd x and p
Give comments on your results (see Examples 2.15 and 2.22, Quantum Mechanics de-
mystied, McMahon, 2005).
6. A particle is known to be in the state
(x, t) = Aexp
_
(x x
o
)
2
4a
2
_
exp
_
i (k
o
x +
o
t)

where the lengthscales x


o
and a are constants, as are the momentum k
o
and angular
frequency
o
. The real constant A can be determined through normalisation procedure.
Find from the above statefunction
(a) the constant A
(b) x and x
2

(c) p and p
2

(d) x and p
Comments on your results (see Problem 3.11, Ch.3, Introductory Quantum Mechanics,
Richard Libo, 1980).
Chapter 7
THE WAVE-MECHANICS
APPROACH
In Chapter 6, we discussed a physical state through a given wavefunction, from which we
could draw any information of the system by applying an eigenvalue equation, normalising
the wavefunction and calculating expectation values of observables. In doing so, we did not
need to specify the wavefunction, whether it depended on both space and time or it was a
spacial function only. It is time to ask for the details of the wavefunction and the detailed
wavefunction is used to govern an equation that captures much of the essence of the dynamics
of sub-atomic particles through wave-mechanics approach. This approach is derived from
an energy eigenvalue equation and hence, an operator for the total energy, the Hamiltonian
operator is introduced here. The governing equation is the second order partial dierential
equation and is known as the Schr odinger equation. Once developed, it is solved to describe
the dynamics of the system. This chapter is thus devoted to solving the time-independent
Schr odinger equation for a variety of given physical systems with simple forms of potential.
7.1 Stationary States
The following are preliminary steps for developing the governing equation used to describe
the dynamics of microscopic particles. Firstly, we re-write a symbol for a one-dimensional
wavefunction, (x, t), and separate it into spatial and temporal components as follows,
(x, t) = (x) T(t) (7.1)
Secondly, we write the Hamiltonian (total energy) operator as

H =

K +

V (7.2)
53
54 7. THE WAVE-MECHANICS APPROACH
where

K = p
2
x
/2m ( p
x
= i /x) is the kinetic energy operator and

V is the potential
energy operator, assumed to be independent of time t and hence

V is a function of space only,

V = V (x). This is useful for deriving time-independent eigenstates, (x). These states are of
great interest in most cases of one-dimensional potential problems and will be determined for
various physical situations, as shown below.
The time-independent state (x) is derived using the method of separation of variables
developed by multiplying both sides of (7.2) by (x, t) as follows

H (x, t) =

K (x, t) +

V (x, t)
i
dT
dt
=

2
2m
d
2

dx
2
T + V T
(7.3)
The above equation is the time-dependent Schr odinger equation. Dividing both sides of
(7.3) by T yields
i
1
T
dT
dt
=

2
2m
1

d
2

dx
2
+ V (7.4)
The left-hand side of (7.4) is a function of t alone and the right-hand side is a function of x
alone. This can possibly true if both sides are constant. For some reasons that will appear in
a moment, we write the separation constant as E. Thus, the right-hand side of (7.4) can be
written as
0 =
d
2

dx
2
+
2m(E V )

2
(7.5)
which is known as the time-independent Schr odinger equation. Without knowing the
explicit form of V (x), it is impossible to solve the Schr odinger equation (7.5).
We can, however, go further by equating the left-hand side of (7.4) to the scalar constant
E such that it can be re-written as
0 =
dT
dt
+
iE

T (7.6)
The solution of (7.6) is in the form of e
iEt/
. Thus, for each time-dependent Schr odinger
equation that involves (x, t) there corresponds a stationary solution (x) and an oscillating
term e
iEt/
. In this way, (7.1) can then be re-written as
(x, t) = (x) e
iEt/
(7.7)
Recall that from (6.17) the complete wavefunction (x, t) is a linear combination of (x),
and thus it is written as
(x, t) =

n=1
c
n

n
(x) e
iE
n
t/
= c
1

1
e
iE
1
t/
+ c
2

2
e
iE
2
t/
+ . . . . . . (7.8)
7.1. Stationary States 55
Suppose now we have an eigenvalue equation for energy in which the Hamiltonian operator

H operates on the full wavefunction (x, t) as follows,

H (x, t) = E (x, t) (7.9)


As

H is a function of x alone, along with (7.8), the above equation can be written as

H
n
(x) = E
n

n
(x) (7.10)
where E
n
is the eigenvalues of

H, providing the possible values of (total) energy of the system
and
n
(x) is the eigenfunctions for stationary states. The subscribe n in both E
n
and
n
(x)
refers to a particular solution represented by one of these states. Thus, there is a dierent
wavefunction for each allowed level of energy.
Below are some interesting discussion regarding the use of both (x, t) and (x) in basic
postulates of quantum mechanics. We here, along with (7.7), show that the probability density
does not depend upon time.
[ (x, t) [
2
=

e
iEt/
e
iEt/
=

= [ (x) [
2
(7.11)
The same thing occurs when we calculate the expectation value of any observable, as shown
below.
A =
_


Adx =
_

e
iEt/

A e
iEt/
dx =
_


A dx (7.12)
Both the probability density and expectation value are constant in time, as expected from a
physical point of view. Thus, we might expect that in any measurement of a physical quantity
the steady (constant) value is obtained.
Moreover, every measurement of the total energy is certain to return the value of E. This
can be easily proven as follows.
H =
_


H dx =
_

E dx = E
_

[ (x) [
2
dx = E (7.13)
and
H
2
=
_


H
2
dx =
_


H E dx = E
_


H dx = E
2
(7.14)
Therefore, the mean-square deviation of measurements of total energy is given by
(H)
2
= H
2
H
2
= 0 (7.15)
56 7. THE WAVE-MECHANICS APPROACH
x=a
V(x)= V(x)=
x=0
-

Figure 7.1: The innite square well potential for a trapped particle of mass m.
giving zero spread distribution, as the standard deviation H of total energy is obviously zero.
Thus, the stationary states represented by
n
(x) are states with denite total energy E
n
for
each associated eigenstate.
7.2 The Free Particle and Innite Square Well
Suppose a particle of mass m conned to an innite, square well of width a along the x-axis
(see Figure 7.1), with the surrounding potential given by
V (x) = 0 0 x a
= otherwise
The particle inside the innite well is then said to be completely free (with having only kinetic
energy) except at the endwalls at both x = 0 and x = a, where an innite force prevents it
from escaping. A similar situation in a classical model would be an object moving back and
forth on a frictionless surface bounded by rigid walls. In both cases, there is no way out for
the particle to go through the walls. Thus, outside the well, (x) = 0 and hence the dynamics
is applied only to the region inside the well, where V (x) = 0. Therefore, the time-independent
Schr odinger equation inside the well is
0 =
d
2

dx
2
+
2mE

2

=
d
2

dx
2
+ k
2

(7.16)
where
k =
1

2mE (7.17)
7.2. The Free Particle and Innite Square Well 57
The general solution of (7.16) is in fact a linear superposition of sinusoidal functions as
follows
(x) = Asin kx + Bcos kx (7.18)
where A and B are arbitrary constants. These constants are determined by the boundary
conditions at the walls, where (0) = (a) = 0. Applying one of these conditions results in
B = 0 and hence (7.18) becomes
(x) = Asin kx (7.19)
The second condition applied to the solution gives the following
ka = n for n = 1, 2, 3, . . . . . . (7.20)
in which k is proved to be dependent upon positive integer numbers n. Since the dependence
of total energy E on k is clear from (7.17), the total energy of a particle in an innite square
well potential is quantised. It follows that the magnitude of E is determined by n, as seen
below,
E
n
=
n
2

2
2ma
2
for n = 1, 2, 3, . . . . . . (7.21)
Thus, the above equation provides the eigenvalues of energy, that is the possible, allowed
values of total energy for the particle inside the innite square well, or equally important, the
free particle. It is clear that the total energy increases proportionally to n
2
, that is E
2
= 4E
1
and E
3
= 9E
1
, where E
1
is the lowest energy (see Figure 7.2).
Although we have already found the eigen energies for the particle inside the well (7.21),
we have not determined explicitly the eigen states,
n
(x). The rst thing to do is to x the
constant A through the normalisation procedure (see 6.3 for details). Here we only pick the
starting procedure and the resulting constant A as follows,
1 = A
2
_
a
0
sin
2
nx
a
dx (7.22)
giving A =
_
2/a. Substituting this into (7.19) yields

n
(x) =
_
2
a
sin
n
a
x for n = 1, 2, 3, . . . . . . (7.23)
This is the allowed eigenstates for the particle inside the well (see Figure 7.2). The lowest sate,
or commonly known as the ground state, is associated with n = 1. The excited states are
states with n > 1 (for which the rst and second excited states are associated with n = 2 and
n = 3, respectively). These states have increased energies in proportion to n
2
, compared with
the ground state level.
58 7. THE WAVE-MECHANICS APPROACH
E
E
3
=9E
1
E
2
=4E
1
E
1
x=a x=0

3
(a) (b) (c)
x=a x=0

1
Figure 7.2: Diagrams of (a) eigen energies; (b) eigen states and (c) probabilities for the
trapped particle in an innite square well of width a.
In addition, the eigen states
n
(x) show, from mathematical point of view, some interesting
properties. The states are alternately even and odd functions with respect to the centre of
the well, i.e.,
1
(x) is an even function,
2
(x) is an odd function, and so on. They are also
orthonormal in the sense that
_

m
(x)
n
(x) dx =
mn
for m, n = 1, 2, 3, . . . . . . (7.24)
where
mn
is known as the Kronecker-delta, dened to be equal to 1 if m = n and to be
zero if m ,= n.
7.3 Simple-Step Potential
Having discussed the dynamics of a particle conned to a well of limited width with no
potential applied (i.e., V = 0), we here examine a beam of particles with energy E incident
upon a potential barrier of constant height V , where E may be greater or less than V . Thus,
the complete Schrodinger equation is given by
0 =
d
2

dx
2
+
2m

2
(E V ) (7.25)
This is generally considered as a one-dimensional problem of step potential, which involves
some simple forms of a given, constant value of potential. As the dynamics is constrained by
the conservation of mass and energy, the following discussions are then necessary.
7.3. Simple-Step Potential 59
7.3.1 The Continuity Equation and Current Density
Mass is conserved in any of physical phenomena, including the problem of moving particles
from one region to another. The conservation of mass is frequently written as the continuity
equation as follows,

t
+ . J = 0 (7.26)
where is the probability density of the particles, dened as = [ [
2
= [ [
2
(see also 7.11),
and J is the current density, describing the rate at which the particles are owing past a
closed region of surface area S and volume V in space. In principle, the continuity equation
is a statement of the conservation of matter. It tells about the way in which the total number
of particles inside the volume V is xed constant such that there is no net ux of particles
owing through the surface S. This can be mathematically expressed as

t
_
V
dV =
_
S
J. n dS (7.27)
where n is the unit vector perpendicular to the surface S. Here the left-hand side of (7.27)
describes the number of particles coming into the volume V through the surface S, while the
right-hand side denotes the number of particles owing out of the surface.
As required in the continuity equation, we derive the explicit form of the current density.
For the sake of simplicity, we write the continuity equation in one-dimensional motion only,
from which the one-dimensional current density J
x
is derived (the x-axis is chosen as the
directional motion). To begin with, let us recall again that the Hamiltonian operator can be
written as

H = i /t. The associated eigenvalue equations for energy are then given by

t
=
i

H and

t
=
i

(7.28)
Re-writing(7.3) for the stationary states (x) in place of (x, t) yields

H =

2
2m
d
2

dx
2
+ V and

H

2
2m
d
2

dx
2
+ V

(7.29)
from which (7.28) becomes

t
=
i

_

2
2m
d
2

dx
2
+ V
_
and

t
=
i

_

2
2m
d
2

dx
2
+ V

_
(7.30)
The time-varying probability density is given by

t
=

t
[ (x) [
2
=


t
+

t
(7.31)
60 7. THE WAVE-MECHANICS APPROACH
Substituting (7.30) into (7.31) yields

t
=
i

_

2
2m
d
2

dx
2
+ V
_

i

_

2
2m
d
2

dx
2
+ V

_
=
i
2m
d
dx
_

d
dx

d

dx
_ (7.32)
from which we can construct the one-dimensional continuity equation in the form of

t
+
d
dx

2mi
_

d
dx

d

dx
_
= 0 (7.33)
Equation (7.33) can be simply re-written as

t
+
dJ
x
dx
= 0 (7.34)
where J
x
is the current density in the-x direction, given by
J
x
=

2mi
_

d
dx

d

dx
_
(7.35)
This quantity will be used to calculate the coecients of both reection and transmission
which describe the proportion of incident energy reected from and transmitted through the
potential barrier.
7.3.2 Reection and Transmission Coecients
The stream of incoming particles along the x-axis with initial energy E on the wall of
potential can be considered as a one-dimensional scattering problem, where the particles are
in the form of plane waves with denite momentum. These waves involve incident, reected,
and transmitted current densities denoted by J
inc
, J
ref
and J
tra
, respectively,
J
inc
=

2mi
_

inc
d
inc
dx

inc
d

inc
dx
_
J
ref
=

2mi
_

ref
d
ref
dx

ref
d

ref
dx
_
J
tra
=

2mi
_

tra
d
tra
dx

tra
d

tra
dx
_
(7.36)
Given the wavefunctions relevant to incindent, reected and transmitted waves then one
may calculate the corresponding current densities (7.36). The coecients of reection and
transmission describe how much part of the incident energy is scattered as the reected and
7.3. Simple-Step Potential 61
transmitted energies. These coecients are respectively dened as
R =

J
ref
J
inc

and T =

J
tra
J
inc

(7.37)
As earlier mentioned, the fundamental conservation of energy must hold for the scattering
problem. This leads to a simple relationship between R and T as follows,
R + T = 1 (7.38)
The above equation is independent of the details of potential and geometry of the wall,
meaning that it is true for all one-dimensional potential problems. This can be easily proved
by setting /t = 0 as the scattering process is assumed to be steady. Hence, the continuity
equation becomes simply dJ
x
/dx = 0. Integrating this yields
_

dJ
x
dx
dx = 0
J

= 0
(7.39)
Now assuming that an incident wave travels along the x-axis with its wavefunction,

inc
(x) = Ae
ik
1
x
, and eventually hits the wall at x = 0 (k
1
is the propagation constant in
domain 1, i.e., the left side of the wall). This wave is reected from and transmitted through
the wall with the reected and transmitted wavefunctions are given by
ref
(x) = Be
ik
1
x
and

tra
(x) = Ce
ik
2
x
, where k
2
is the propagation constant in domain 2, i.e., the right side of the
wall. Here A, B and C are the amplitudes of the incident, reected, and transmitted waves,
respectively. The net ux of particles in domain 1 is thus represented by J

= J
inc
J
ref
,
while in domain 2 it is true that J

= J
tra
. Substituting these into (7.39) gives
J
ref
+ J
tra
= J
inc
(7.40)
as expected. Dividing both sides of (7.40) by J
inc
yields

J
ref
J
inc

J
tra
J
inc

= 1 (7.41)
which is, along with (7.37), exactly the same as (7.38).
We can prove the conservation of energy using a dierent approach. Let the incident
inc
,
reected
ref
and transmitted
tra
waves be the ones given above. Then we can write
d
inc
dx
= ik
1
Ae
ik
1
x
d
ref
dx
= ik
1
Be
ik
1
x
d
tra
dx
= ik
2
Ce
ik
2
x
62 7. THE WAVE-MECHANICS APPROACH
Boundary conditions at x = 0 require that

1
=
2

inc
+
ref
=
tra
A + B = C
(7.42)
and
d
1
dx
=
d
2
dx
d
inc
dx
+
d
ref
dx
=
d
tra
dx
k
1
(A B) = k
2
C
(7.43)
where subscripts 1 and 2 refer to the domains 1 and 2, respectively. It is easy to derive from
(7.42) and (7.43) that
A =
1
2
k
1
+ k
2
k
1
C and B =
1
2
k
1
k
2
k
1
C (7.44)
These are needed to calculate both the coecients of reection and transmission. But rst,
with the help of (7.36), we construct the current density for each wave as follows
J
inc
=

2mi
2ik
1
A
2
J
ref
=

2mi
2ik
1
B
2
J
tra
=

2mi
2ik
2
C
2
(7.45)
Thus, it is clear from (7.37), (7.44) and (7.45) that the reection and transmission coecients
are given by
R =

J
ref
J
inc

=
_
B
A
_
2
=
_
k
1
k
2
k
1
+ k
2
_
2
(7.46)
and
T =

J
tra
J
inc

=
_
C
A
_
2
k
2
k
1
=
_
2k
1
k
1
+ k
2
_
2
k
2
k
1
(7.47)
from which we obtain
R + T =
(k
1
k
2
)
2
+ 4k
1
k
2
(k
1
+ k
2
)
2
=
k
2
1
+ 2k
1
k
2
+ k
2
2
(k
1
+ k
2
)
2
= 1 (7.48)
as expected from the perspective of the conservation of energy.
7.3. Simple-Step Potential 63
x=0
-

E
V
o
Figure 7.3: Sketch of a simple-step potential where incoming energy E is greater than
the height V
o
of the potential barrier.
7.3.3 Cases where E is greater than V
Now we are ready to consider the problem of a simple-step potential for the case of E > V
(see Figure 7.3). We wish to obtain the solution to the the time-independent Schr odinger
equation for this case in both regions (on the left and right-hand sides of the wall at x = 0).
Note that the potential function is zero for x < 0 (domain 1) and is constant at V
o
for x 0
(domain 2).
In the domain 1, the dynamics is independent of potential hence the Schr odinger equation
is given by
0 =
d
2

1
dx
2
+
2mE

2

1
=
d
2

1
dx
2
+ k
2
1

1
(7.49)
where
k
1
=
1

2mE (7.50)
with the general solution is in the form of

1
(x) = Ae
ik
1
x
+ Be
ik
1
x
(7.51)
where A and B are the amplitudes of the incident and reected beams.
Whereas in the domain 2, the potential exists at a constant value of V
o
hence the Schr odinger
equation appears to be
0 =
d
2

2
dx
2
+
2m(E V
o
)

2

2
=
d
2

2
dx
2
+ k
2
2

2
(7.52)
64 7. THE WAVE-MECHANICS APPROACH
x=0
-

E
V
o
Figure 7.4: Sketch of a simple-step potential where incoming energy E is less than the
height V
o
of the potential barrier.
where
k
2
=
1

_
2m(E V
o
) (7.53)
with the general solution is in the form of

2
(x) = Ce
ik
2
x
(7.54)
where C is the amplitude of the transmitted beam.
Since any wavefunction and its rst derivative must be continuous at the boundary (i.e.,
at x = 0), then the following relations (see detailed discussion in 7.3.2) are obtained,
B
A
=
k
1
k
2
k
1
+ k
2
and
C
A
=
2k
1
k
1
+ k
2
(7.55)
for which the coecients of both reection and transmission are
R =
_
k
1
k
2
k
1
+ k
2
_
2
and T =
4k
1
k
2
(k
1
+ k
2
)
2
(7.56)
giving the required conservation of energy, R + T = 1 (see also 7.48).
7.3.4 Cases where E is less than V
In the case of E < V (see Figure 7.4), the region of dynamics is divided into two similar
domains to the previous case. In the domain 1, the dynamics is again developed for the
potential-independent Schr odinger equation,
0 =
d
2

1
dx
2
+
2mE

2

1
=
d
2

1
dx
2
+ k
2
1

1
(7.57)
7.3. Simple-Step Potential 65
where
k
1
=
1

2mE (7.58)
with the general solution is given by

1
(x) = Ae
ik
1
x
+ Be
ik
1
x
(7.59)
where A and B are the amplitudes of the incident and reected beams, respectively.
Unlike the rst case, this time we need to be careful about the dynamics in the domain 2
as E < V
o
. In this domain, the Schrodinger equation appears as
0 =
d
2

2
dx
2
+
2m(E V
o
)

2

2
=
d
2

2
dx
2
+ k
2
2

2
(7.60)
where
k
2
=
1

_
2m(E V
o
) < 0 (7.61)
We here introduce another constant, namely such that (7.61) becomes
k
2
=
i

_
2m(V
o
E) = i (7.62)
where
=
1

_
2m(V
o
E) > 0 (7.63)
Thus, the general solution of (7.60) is

2
(x) = Ce
ik
2
x
= Ce
x
(7.64)
where C is the amplitude of the transmitted beam. Note that in classical physics this domain
is forbidden, but in quantum mechanics it is still possible for the beam to penetrate the barrier.
The wavefunction in the domain 2,
2
(x), is no longer an oscillation term. Instead, it decreases
with x as it is required for the physical meaning to be true (i.e., the wavefunction in this region
must dissapear at x = ).
The next step is the same as in the previous case. Again, applying boundary conditions
at x = 0 results in
B
A
=
k
1
i
k
1
+ i
and
C
A
=
2k
1
k
1
+ i
(7.65)
which is similar to (7.55). It follows that R = 1 and T = 0, hence there is total reection with
zero transmission (as expected from a classical point of view). To obtain the latter result,
we have to calculate the associated density current for the transmitted wave (if any). The
66 7. THE WAVE-MECHANICS APPROACH
wavefunction in the region 2,
2
, is in the form of a complex amplitude times a real function
of x. Such a wavefunction does not represent a propagating wave as it has no an oscillating
term (compared with the wavefunction in the region 1,
1
). This type of wave is known as the
evanescent wave, describing a wave with no current with it and thus no energy transported.
This can be easily proved by calculating the transmitted current density as follows,
J
tra
=

2mi
_

tra
d
tra
dx

tra
d

tra
dx
_
=

2mi
[ C [
2
_
e
x

x
e
x
e
x

x
e
x
_
= 0
(7.66)
where
tra
is given by (7.64). It is left for students to calculate the density current, J
ref
for
the reected beam and to prove that energy is conserved for this case.
7.4 The Harmonic Oscillator
A classical harmonic oscillator is all about an object of mass m attached to a spring of
constant k. If an external force is exerted on the object such that the object, along with
the spring, is stretched or compressed from its equilibrium position, then a simple oscillation
begins. The dynamics is governed by the Hookes law in the form of
0 =
d
2
x
dt
2
+
k
m
x
=
d
2
x
dt
2
+
2
x
(7.67)
where =
_
k/m is the angular frequency of the oscillation. The general solution of (7.67) is
known as the displacement x of the object from the equilibrium position at any time t, which
is given by
x(t) = Asin t + Bcos t (7.68)
with the total energy for the oscillatory motion is
E =
1
2
m x
2
+
1
2
m
2
x
2
(7.69)
The above equation is the one we want to make a point. Unlike simple-step potential cases,
the potential energy of the oscillatory motion produced by harmonic oscillator is no longer
constant. Rather, it is a function of x squared. This leads to a special treatment for solving
the time-independent Schrodinger equation, as opposed to the previous cases with no potential
applied (innite square well) or with a given constant potential (simple-step potential).
7.4. The Harmonic Oscillator 67

n
(a) (b)

n+1
a
+

n+1
a
-
Figure 7.5: Sketches of the (a) raising and (b) lowering operators in quantum mechanics.
In quantum mechanics, harmonic oscillator can, to some degree, be used as an alternative,
simple approach to describe the dynamics of microscopic systems by setting up the time-
independent Schr odinger equation for the space-dependent potential as follows,


2
2m
d
2

dx
2
+
1
2
m
2
x
2
= E (7.70)
Thus, this section is devoted to solving the Schrodinger equation above. Here, two dierent
approaches will be introduced, namely the algebraic and analytic methods.
7.4.1 The Algebraic Method
The idea is to introduce ladder operators as they allow us to climb one level up and down in
energy. These operators involve the raising a
+
and lowering a

operators, as a combination
of both the displacement x and momentum p
x
operators, as seen below
a
+
=

2
_
x
i p
x
m
_
and a

2
_
x +
i p
x
m
_
(7.71)
where
_
m/. As described in Figure 7.5 above, the properties of the raising and
lowering operators are as follows,
a
+

n
=

n + 1
n+1
a

n
=

n
n1
(7.72)
where

n + 1 and

n are the eigenvalues of the raising and lowering operators, respectively,
for the associated wavefunction
n
.
The following relations are derived from the denitions of the raising and lowering operators
given in (7.71),
a
+
+ a

2 x and a

a
+
= i

2
p
x
m
(7.73)
68 7. THE WAVE-MECHANICS APPROACH
for which the displacement x and momentum p
x
operators can be written in terms of the
raising and lowering operators as follows
x =
1

2
(a
+
+ a

) and p
x
=
m
i

2
(a

a
+
) (7.74)
In quantum mechanics, the fundamental commutator relation between x and p
x
is given
by [ x, p
x
] = x p
x
p
x
x = i. Using this, it can be shown that the raising and lowering
operators meet the following commutator relation, [ a

, a
+
] = a

a
+
a
+
a

= 1. Let us
now dene a new operator, namely the number operator

N = a
+
a

, dened as such for


the following reason,

N
n
= a
+
a

n
(x)
= a
+

n
n1
=

n a
+

n1
=

n
n
= n
n
(7.75)
where we have used the properties of the ladder operators given in (7.72) and n is an integer
number. We stop discussing this operator at the moment, and will be back to use it to derive
the energy level of the harmonic oscillator.
We write the Hamiltonian, along with the space-dependent potential, for the harmonic
oscillator as

H =
p
2
x
2m
+
1
2
m
2
x
2
=

4
_
(a

a
+
)
2
(a

+a
+
)
2
_
= (a
+
a

+ 1/2)
= (

N + 1/2)
(7.76)
since [ a

, a
+
] =1 and

N = a
+
a

, as previously dened. Then, we can construct the eigen


energy equation for the harmonic oscillator using the Hamiltonian in (7.76),

H
n
= (

N + 1/2)
n
E
n
= (n + 1/2)
n
(7.77)
for which the energy level of the harmonic oscillator can be deduced to be
E
n
= (n + 1/2) for n = 0, 1, 2, 3, . . . (7.78)
7.4. The Harmonic Oscillator 69
It follows that the ground state energy for the harmonic oscillator is E
o
= /2, the energy
of the rst excited state is then E
1
= 3/2, and so on.
Although the algebraic method provides evidence for energy quantisation in the harmonic
oscillator, it does not give us an explicit form of the normalised eigenfunction. As this is also
part of the complete solution to the Schr odinger equation for the oscillatory motion, then we
need to examine another method, i.e., the analytic method, to nd out what the normalised
eigenfunction looks like.
7.4.2 The Analytic Method
The idea behind this method is to solve the time-independent Schrodinger equation in (7.70)
by introducing a new parameter, x. In terms of this parameter, the Schr odinger equation
for the harmonic oscillator reads
d
2

d
2
+
_
2E


2
_
= 0 (7.79)
In the process to solve the equation above, we obtain the allowed values of E, which is exactly
the same as that of (7.78). Thus, we recover completely the result from the previous method,
that is the fundamental quantisation of energy.
Note that the raising and lowering operators given in (7.71) can also be written in terms
of ,
a
+
=

2
_
x

m
d
dx
_
=
1

2
_

d
d
_
a

2
_
x +

m
d
dx
_
=
1

2
_
+
d
d
_ (7.80)
by which we can derive the ground state wavefunction
o
of the simple harmonic oscillator.
Using the characteristics of the lowering operator, a

o
= 0, we can write simply
_
+
d
d
_

o
= 0 (7.81)
which has the general solution in the form of

o
= A
o
e

2
/2
= B
o
e

2
x
2
/2
(7.82)
depending upon which parameter we use. Here A
o
and B
o
are the normalisation constants for
each physically acceptable solution with dierent variables, that is and x. These constants
are determined by the normalisation procedure as follows. For the -dependent wavefunction,
70 7. THE WAVE-MECHANICS APPROACH
the normalisation procedure for the ground state implies that
1 = A
2
o
_

2
d (7.83)
for which A
o
= ()
1/4
is obtained. Thus, the ground state in this variable is written as

o
() =
_
1

_
1/4
e

2
/2
(7.84)
For the space-dependent wavefunction, on the other hand, the normalisation procedure for
the ground state implies that
1 = B
2
o
_

2
x
2
dx (7.85)
for which B
o
= (
2
/)
1/4
is obtained. Thus, the ground state in terms of x is written as

o
(x) =
_

_
1/4
e

2
x
2
/2
(7.86)
The ground state wavefunction is found to be a purely, exponentially decaying function, having
no oscillatory component, regardless of the parameter used (see 7.84 and 7.86).
The remaining normalised wavefunctions, or the higher-energy eigenstates, of the simple
harmonic oscillator can be generated using the raising operator dened in (7.72) and (7.80).
We provide here the rst three excited-state wavefunctions,

1
=
1

1
(a
+
)
1

2
=
1

2
(a
+
)
2

3
=
1

6
(a
+
)
3

o
(7.87)
for which we can deduce the general formula for nding the normalised wavefunction as

n
() =
1

n!
(a
+
)
n

o
=
1

n!
(a
+
)
n
_
1

_
1/4
e

2
/2
= A
n
_

d
d
_
n
e

2
/2
= A
n
1
n
() e

2
/2
(7.88)
7.4. The Harmonic Oscillator 71
where A
n
is the normalisation constant for the -dependent wavefunction, given by
A
n
=
_
1

_
1/4
1

2
n
n!
for n = 0, 1, 2, 3, . . . (7.89)
and 1
n
() is the well-known Hermite polynomial. Here are the rst three values of this
polynomial, 1
o
() = 1, 1
1
() = 2 and 1
2
() = 4
2
2.
The general solution for the x-dependent wavefunction is depicted in Figure 7.6, which is
mathematically similar to the last line of (7.88),

n
(x) = B
n
1
n
(x) e

2
x
2
/2
(7.90)
where B
n
is the normalisation constant for the x-dependent wavefunction, given by
B
n
=
_

_
1/4
1

2
n
n!
for n = 0, 1, 2, 3, . . . (7.91)
and 1
n
(x) is the Hermite polynomial, where 1
o
(x) = 1, 1
1
(x) = 2x and 1
2
(x) = 4x
2
2.
In addition, we also provide a number of important properties of the simple harmonic
oscillator wavefunction. These include some integrals as follows
_

n
(x)
d
dx

m
(x) dx =
_
n + 1
2
for m = n + 1
=
_
n
2
for m = n 1
= 0 for m = n
(7.92)
_

n
(x) x
m
(x) dx =
1

_
n + 1
2
for m = n + 1
=
1

_
n
2
for m = n 1
= 0 for m = n
(7.93)
_

n
(x) x
2

m
(x) dx =

(n + 1)(n + 2)
2
2
for m = n + 2
=
_
n(n 1)
2
2
for m = n 2
=
2n + 1
2
2
for m = n
(7.94)
72 7. THE WAVE-MECHANICS APPROACH
Figure 7.6: Sketches of (a) the rst three wavefunctions of the harmonic oscillator and
(b) the corresponding probability densities (taken from Figure 7.10, Ch.7, Libo, 1980).
The above properties are useful for deriving the principle of uncertainty. It is then left for
the students to prove that the Heisenberg uncertainty principle given in (5.11) holds for the
simple harmonic eigenstates (7.90).
7.5 Exercises
1. At t = 0 it is known that of 1000 neutrons in a one-dimensional box of width 10
5
cm,
100 neutrons have energy 4E
1
and the other 900 have energy 225E
1
.
(a) Construct a state function that has these properties.
(b) Use the state you have constructed to calculate the density (x) of neutrons per unit
length.
(c) How many neutrons are in the left half of the box?
(see Problem 5.5, Ch.5, Introductory Quantum Mechanics, Richard Libo, 1980).
7.5. Exercises 73
2. Measurement of the position of a particle in a one-dimensional box with walls at x = 0
and x = L nds the value x = L/2.
(a) Show that in the subsequent measurement, it is equally probable to nd the particle
in any odd-energy eigenstate.
(b) Show that the probability of nding the particle in any even-eigenstate is zero (an
eigen
n
is even if n is even and odd if n is odd).
(see Problem 5.9, Ch.5, Introductory Quantum Mechanics, Richard Libo, 1980).
3. Let us consider the particle described in problem 2 is in the ground state. The right
wall then is then moved to x = 2L, in a time relatively short compared with the natural
period 2/, where E
1
=
1
. If the energy of the particle is measured soon after this
expansion, what value of energy is likely to be found? How does this energy compare
with the particles initial energy?
(see Problem 5.11, Ch.5, Introductory Quantum Mechanics, Richard Libo, 1980).
4. (a) Show that for a particle in a one-dimensional box with walls at (L/2, L/2)
p
min
=
_
p
2

min
=
h
2L
and x
max
=
_
x
2

max
=
L
2

3
(b) In which states are p
min
and x
max
realised?
(see Problem 5.41, Ch.5, Introductory Quantum Mechanics, Richard Libo, 1980).
5. (a) Show that the fundamental commutator relation [ x, p
x
] = i
(b) Using the result in (a) show that [ a

, a
+
] =1
(see Problem 7.5, Ch.7, Introductory Quantum Mechanics, Richard Libo, 1980).
6. (a) Show that the normalised nth eigenstate
n
is generated from the normalised ground
state
o
through

n
=
1

n!
(a
+
)
n

o
(b) Show that part (a) implies the following relations
a
+

n
=

n + 1
n+1
a

n
=

n
n1
(see Problem 7.9, Ch.7, Introductory Quantum Mechanics, Richard Libo, 1980).
74 7. THE WAVE-MECHANICS APPROACH
7. Show that in the nth eigenstate of the simple harmonic oscillator,
(a) x
2
= (x)
2
and p
2
= (p)
2
(b) the Heisenberg uncertainty principle applies, p x /2
(c) Use the result in (b) to derive the zero-point energy E
o
= /2 of a harmonic
oscillator with natural frequency .
(see linked Problems 7.14, 7.18 and 7.24, Ch.7, Introductory Quantum Mechanics, Richard
Libo, 1980).
8. Electrons in a beam of density = 10
15
electrons/m are accelerated through a potential
of 100 V. The resulting current then impinges on a potential step of height 50 V. What
are the incident, reected and transmitted currents?
(see Problem 7.40, Ch.7, Introductory Quantum Mechanics, Richard Libo, 1980).
9. A uniform homogeneous beam of electrons is incident on a rectangular barrier of height V .
Each electron in the beam has energy E > V and the unit amplitude wavefunction

inc
= e
ik
1
x
. If the transmitted beam has wavefunction
tra
= 0.97e
ik
1
x
,
(a) What is the total wavefunction of the electrons in region 1?
(b) If E = 10 eV and V = 5 eV, what is the minimum barrier width compatible with
the information given above?
(see Problem 7.76, Ch.7, Introductory Quantum Mechanics, Richard Libo, 1980).
10. At time t = 0, a space-time dependent wavefunction representing a one-dimensional
harmonic oscillator case is in the state below,
(x, 0) =
1

0
+
1

1
+
1

2
(a) If the energy is measured, what values can be found and with what probabilities?
(b) Find the average value of the energy H .
(c) Find the explicit form of the stationary states
n
(x) and the basis functions for this
expansion.
(d) Write the form of the wavefunction at time t, i.e., the time-dependent wavefunction.
(see Example 9.8, Ch.9, Quantum Mechanics demystied, McMahon, 2005).
Chapter 8
THE SCHR

ODINGER EQUATION
IN SPHERICAL COORDINATES
Previous analyses of the time-independent Schrodinger equation in Chapter 7 were limited to
only cases of one-dimensional, simple potential problem using a Cartesian coordinate system.
In this chapter, we extent the discussion to three-dimensional formulation of the Schr odinger
equation using spherical coordinates as this is more relevant to real cases. We rst begin
our discussion with a brief description of the Schrodinger equation in three dimensions and
separate it into radial and angular components. Then, we re-examine the hydrogen atom (see
Chapter 3) by introducing a more comprehensive method for nding the complete solution to
the spherical coordinate-based Schrodinger equation. The primary goal of the new method is
to show that at least two more quantum numbers are needed to describe the dynamics of the
hydrogen atom and hydrogen-like atoms. The new quantum numbers introduced here are the
orbital quantum number and the magnetic quantum number m. These two quantum
numbers are derived from the solution to the angular component of the complete solution.
The radial component plays a role in that it provides a physical reason of why the description
of the hydrogen is always something like a circle with a certain radius, namely a shell K with
only a single electron orbitting the nucleus. We then put both the Heisenberg uncertainty
principle and Born interpretation into the right place in this matter. In the process of nding
the complete solution and understanding better the problem in question, we also discuss the
properties of angular momentum, an important quantity necessary to better understand the
dynamics of the hydrogen atom and hydrogen-like atoms. We here consider both orbital and
spin angular momenta; orbital angular momentum is closely akin to that in classical physics,
whereas spin angular momentum has no classical counter-part. These two momenta are of
fundamental interest in quantum mechanics, which will be discussed in terms of operators at
the end of this chapter.
75
76 8. THE SCHR

ODINGER EQUATION IN SPHERICAL COORDINATES


8.1 The Schr odinger Equation in Spherical Coordinates
The generalisation of the governing equation to a three-dimensional formulation is straight
forward as follows,
p
x
p
i
d
dx
i = i
_
x

x
+ y

y
+ z

z
_
(8.1)
Thus, the time-dependent Schrodinger equation in three dimensions is written as
i
(r, t)
t
=

2
2m

2
(r, t) + V (r, t) (8.2)
where

2
=

2
x
2
+

2
y
2
+

2
z
2
(8.3)
is the Laplacian operator in Cartesian coordinates and (r, t) is the complete wavefunction.
As can be written as a linear superposition of innite stationary eigenstates (see 7.1),
the general solution to the time-dependent Schrodinger equation (8.2) is
(r, t) =

n=1
c
n

n
(r) e
iE
n
t/
= c
1

1
(r) e
iE
1
t/
+ c
2

2
(r) e
iE
2
t/
+ . . . . . . (8.4)
where [ c
n
[
2
denotes P
n
, the probability for nding a particle in the nth stationary state
n
(r).
Using the Laplacian and separate solutions, the time-independent Schr odinger equation given
in (7.5) is thus re-written in three-dimensional formulation as
0 =
2
(r) +
2m(E V )

2
(r) (8.5)
where both the stationary wavefunction (r) and potential V are a function of space only, i.e.,
(r) = (x, y, z) and V = V (x, y, z), respectively. What we have is the three-dimensional,
time-independent Schr odinger equation in the Cartesian representation. It is, however, more
convenient to use spherical coordinates for a comprehensive description of the dynamics of
the hydrogen atom. In this coordinate system, the solution to the Schr odinger equation is
separated into radial and angular components.
8.1.1 Separation of Dynamic Variables
Assume that the potential is only a function of the distance from the origin. In this situation,
it is then common to use spherical coordinates (r, , ) as dynamic variables in the governing
8.1. The Schr odinger Equation in Spherical Coordinates 77
equations. In the spherical coordinates, the Laplacian takes the form of

2
=
1
r
2

r
_
r
2

r
_
+
1
r
2
sin

_
sin

_
+
1
r
2
sin
2

_

2

2
_
(8.6)
Using (8.6) above, the time-independent, three-dimensional Schrodinger equation in (8.5) is
re-written as


2
2m
_
1
r
2

r
_
r
2

r
_
+
1
r
2
sin

_
sin

_
+
1
r
2
sin
2

_

2

2
_
_
+ V = E (8.7)
with the general solution is given by
(r, , ) = R(r) Y (, ) (8.8)
where R(r) is a radial function and Y (, ) is a spherical harmonic function, depending upon
angular variables and . Substituting R and Y into (8.7) yields


2
2m
_
Y
r
2
d
dr
_
r
2
dR
dr
_
+
R
r
2
sin

_
sin
Y

_
+
R
r
2
sin
2

2
Y

2
_
_
+ V RY = E RY
(8.9)
Dividing both sides by RY and multiplying by -2mr
2
/
2
results in
_
1
R
d
dr
_
r
2
dR
dr
_

2mr
2

2
(V E)
_
+
1
Y
_
1
sin

_
sin
Y

_
+
1
sin
2

2
Y

2
_
= 0 (8.10)
It is clear that the rst part of the left-hand side of (8.10) depends only on r whereas the
second part contains two variables, and . Because the sum of the two is zero, they must
be a constant and accordingly have to be opposite. This leads to the followings equations for
each part,
1
R
d
dr
_
r
2
dR
dr
_

2mr
2

2
(V E) = ( + 1) (8.11)
1
Y
_
1
sin

_
sin
Y

_
+
1
sin
2

2
Y

2
_
= ( + 1) (8.12)
where is a separation constant. For some reasons at the moment, we do not need to specify
this constant.
In the next section, we will discuss the above equations (8.11) and (8.12) separately and
develop new ideas about how many quantum numbers are actually required for describing
the hydrogen atom and hydrogen-like atoms. Recall that in the heart of the Bohr model of
the atom it is n the principal quantum number the only one quantum number necessary
78 8. THE SCHR

ODINGER EQUATION IN SPHERICAL COORDINATES


for formulating discretised levels of both momentum and energy of the hydrogen atom and
hydrogen-like atoms (see discussion on Chapter 3).
8.1.2 The Angular Component
The dependence of the spherical harmonic function Y on and is obvious from (8.12).
Multiplying both sides of (8.12) by Y sin
2
yields
sin

_
sin
Y

_
+

2
Y

2
= ( + 1) Y sin
2
(8.13)
Now write the spherical harmonic function Y as a function of a latitude part multiplied
by a longitude part , Y = () (), and put them into (8.13), giving
1

_
sin
d
d
_
sin
d
d
__
+ ( + 1) sin
2
+
1

d
2

d
2
= 0 (8.14)
The rst two terms on the left-hand side of (8.14) is dependent upon only while the third
is a function of only. Thus, we get ready for nding the separate solutions for each. Again,
we here assume that the two parts are opposite constants, canceling each other to be zero.
Thus, it makes sense to write both as
1

_
sin
d
d
_
sin
d
d
__
+ ( + 1) sin
2
= m
2
(8.15)
1

d
2

d
2
= m
2
(8.16)
for convenience, where m is another separation constant.
As indicated in (8.15), there must be a relation between and m but for now it is not
necessary to nd what lies between them. We consider the second part (8.16) rst as this is
easy to solve and re-write it as
d
2

d
2
+ m
2
= 0 (8.17)
which has the general solution in the form of an oscillating function,
=
1

2
e
im
=
1

2
e
im(+2)
(8.18)
as 2 is the periodicity of the function, and hence the solution. It is clear from (8.18) that
e
im2
= 1. It follows that m is constrained to be an integer,
m = 0, 1, 2, . . . . . . (8.19)
8.1. The Schr odinger Equation in Spherical Coordinates 79
At the end of this chapter, we will derive this result and its relation to using a special
technique for nding the eigenvalues of angular momentum operator. For now it is enough to
say here that m is the third quantum number required to describe the hydrogen, the so-called
magnetic quantum number.
Having found the solution of the function and its constraint on m, now we turn to (8.15)
and re-arrange it to obtain the following form,
1
sin
d
d
_
sin
d
d
_
+
_
( + 1)
m
2
sin
2

_
= 0 (8.20)
Introducing a new parameter cos into (8.20) gives
d
d
_
(1
2
)
d
d
_
+
_
( + 1)
m
2
1
2
_
= 0 (8.21)
where 1 1.
Let us outline the method by which (8.21) is solved. Setting m = 0 and ( + 1) =
results in the well-known Legendres equation below,
d
d
_
(1
2
)
d

d
_
+

= 0 (8.22)
A solution to (8.22) may be sought as a series in powers of . The requirement that the series
solution remains bounded by 1 1 implies (1) the eigenvalues must be in the form of
( + 1), where is an integer satisfying 0 and (2) the solution for

consists of a nite
number of terms in , hence cos Thus, the solution requires a polynomial of order , namely
the Legendre polynomial written in terms of as follows,
T

() =
1
2

!
d

(
2
1)

(8.23)
The solution to (8.21) is obtained by constructing the associated Legendre polynomial
as follows,
T
m

() = (1)
m
(1
2
)
m/2
d
m
T

()
d
m
(8.24)
where m . Thus, the associated Legendre polynomial varies with both and m.
In summary, the solution to (8.21) is given by
m

(), corresponding to T
m

(). The precise


relation between these polynomials is derived from the normalisation procedure,
_
4
[ Y
m

(, ) [
2
d = 1 (8.25)
80 8. THE SCHR

ODINGER EQUATION IN SPHERICAL COORDINATES


where Y
m

(, ) =
m

()
m
() is again the spherical harmonic function and d is the
element of the solid angle, dened as the dierential area of a unit sphere, satisfying
_
,
d =
_
2
0
d
_

0
sin d = 4 (8.26)
The normalisation gives

() =
_
2 + 1
2
( m) !
( + m) !
_
1/2
T
m

() (8.27)
for which, along with (8.18), provides the complete form of the spherical harmonics,
Y
m

(, ) =
_
2 + 1
4
( m) !
( + m) !
_
1/2
T
m

(cos ) e
im
(8.28)
Here are the rst few values of the associated Legendre polynomials, T
0
= 1, T
0
1
= cos ,
T
1
1
= sin and T
1
1
=
1
2
sin . With these, the rst few spherical harmonics are Y
0
0
=
_
1
4
_
1/2
,
Y
0
1
=
_
3
4
_
1/2
cos and Y
1
1
=
_
3
8
_
1/2
sin e
i
.
Note that (8.28) holds for all possible values of m, as given by (8.19). It can also be shown
that the spherical harmonics with positive and negative values of m are correlated each other
through
Y
m

(, ) = (1)
m
Y
m

(, ) for m > 0 (8.29)


8.1.3 The Radial Component
The actual shape of the potential V (r) aects only the radial component R(r) shown in (8.11).
The radial equation can be re-written as
d
dr
_
r
2
dR
dr
_

2mr
2
R

2
(V E) = ( + 1)R (8.30)
Introducing u(r) rR(r) into the above equation simplies it to be


2
2m
d
2
u
dr
2
+
_
V +

2
2m
( + 1)
r
2
_
u = Eu (8.31)
This equation is similar to the one-dimensional, time-independent Schr odinger equation, with
x and being replaced with r and u, respectively, and an additional term for the potential V
such that
V
e
= V +

2
2m
( + 1)
r
2
(8.32)
8.2. The Hydrogen Atom 81
where the extra-term is called as the centrifugal term. This term, if exerted on a body, tends
to drive the body away from the origin (i.e., radially out).
Note here that the normalisation procedure for the u function, or equally the same R,
applies in a similar manner to the angular part, as shown in (8.25). Thus, we can write
_

0
[ u[
2
dr =
_

0
[ R[
2
r
2
dr = 1 (8.33)
This cannot proceed further unless the explicit form of the central potential V (r) is specied.
In other words, the solution to (8.31) is only obtained if a specic form of V (r) is given.
8.2 The Hydrogen Atom
The hydrogen atom contains a relatively motionless, positively charged proton centred at the
origin as a nucleus and a single, negatively charged electron moving around the nucleus. Thus,
the potential Coulomb, owing to the attractive force between the two particles, applies and is
writtten in the SI unit as
V (r) =
e
2
4
o
r
(8.34)
The radial equation (8.31) then becomes


2
2m
d
2
u
dr
2
+
_

e
2
4
o
r
+

2
2m
( + 1)
r
2
_
u = Eu (8.35)
Our task is to solve the above equation for u(r) and hence R(r) and nd the allowed values
of the energy of the electron.
The rst thing to do is to tidy up the radial equation by introducing some new parameters,
such as

1

2mE where E < 0 (8.36)


for bound states, as is the case of the hydrogen atom. With this, (8.35) becomes
1

2
d
2
u
dr
2

_
1
me
2
2
o

1
r
+
( + 1)
(r)
2
_
u = 0 (8.37)
Let us further introduce
r and
o

me
2
2
o

(8.38)
so that (8.37) becomes
d
2
u
d
2

_
1

o

+
( + 1)

2
_
u = 0 (8.39)
82 8. THE SCHR

ODINGER EQUATION IN SPHERICAL COORDINATES


The next step is to simplify (8.39) by assuming that is relatively large compared with
both
o
and ( + 1). We nally obtain a simplied form of (8.39),
d
2
u
d
2
u = 0 (8.40)
with the general solution is given by
u() = Ae

+ Be

Be

(8.41)
since A must be zero for the u() not blowing up as goes to innity. On the other hand, if
0, then the centrifugal term in (8.39) dominates over the other two such that
d
2
u
d
2

( + 1)

2
u = 0 (8.42)
with the general solution is given by
u() = C
+1
+ D

C
+1
(8.43)
since the second term causes u() to be blown up as 0. With these two asymptotic forms
at hand, we are prepared to solve (8.39) by writing a proposed solution in the form of
u() = e

+1
F() (8.44)
where the constants B and C are absorbed into F(), dened as
F() =

i=0
c
i

i
(8.45)
Substituting (8.44) into (8.39) yields

d
2
F
d
2
+ 2( + 1 )
dF
d
+ [
o
2( + 1) ] F = 0 (8.46)
The determination of the coecients c
i
in (8.45) involves (1) dierentiating F() term by
term, (2) inserting the results into (8.46) and (3) equating the coecients of like powers.
These will give
c
i+1
=
_
2(i + + 1)
o
(i + 1) (i + 2 + 2)
_
c
i
(8.47)
A careful examination on the recursion formula leads to a conclusion that the series must
terminate at some point, i.e., a maximal integer i
max
exists, beyond which all coecients
8.2. The Hydrogen Atom 83
vanish automatically. As a result, the numerator on the RHS of (8.47) must be zero, or
2(i
max
+ + 1)
o
= 0 (8.48)
We may dene n i
max
+ + 1, allowing us to write
o
= 2n. This, along with both (8.36)
and (8.38), gives
E =

2
2m
=
me
4
8
2

2
o

2

2
o
=
me
4
2
2
_
1
4
o
_
2
1
n
2
(8.49)
The above equation can also be written in terms of the cgs unit as
E
n
=
me
4
2
2
1
n
2
=
1
n
2
for n = 1, 2, 3, . . . (8.50)
which is exactly the same as (3.7) the famous Bohr formula for the energy levels of the
hydrogen atom obtained from the primitive quantum theory (see discussion on Chapter 3).
Here 1 is the Rydberg constant, dened to be 1 = me
4
/2
2
= 13.6 eV. The negative sign
for the electrons total energy in (8.50) suggests that the electron is bound to the nucleus
and thereby requiring some amount of energy to liberate it from the Coulomb attraction of
the nucleus. The lowest possible stationary state corresponding to quantum number n = 1 is
generally called the ground state having the smallest energy of E
1
= 13.6 eV.
When Bohr proposed his theory in 1913, the spectral lines of Balmer series (n
f
= 2 and
n
i
> 2) and Paschen series (n
f
= 3 and n
i
> 3) had already been found two years before. The
Rydberg constant 1 theoretically derived from the Bohrs quantum theory for the hydrogen
is found to be in good agreement with an empirical value for a constant used to describe
Balmer and Paschen lines. This was followed by the Lymann series (1916) where n
f
= 1 and
n
i
> 1. In the years to come since the rst three spectral series, Brackett (1922) and then
Pfund (1924) observed the spectral lines for n
f
= 4 and n
f
= 5, respectively. Impressed by his
successful explanation to the experimental observations of the hydrogen spectral lines, Bohr
then immediately extended his theory to hydrogen-like atoms, such as He
+
and Li
2+
, in
which all but one electron had been removed from a nucleus of positive charge Ze, where Z
denotes the number of proton. For a single electron of negative charge e orbiting the nucleus
of a hydrogen-like atom, the quantized orbit radii is given by
r
n
=
n
2

2
Zme
2
= n
2
a
o
Z
for n = 1, 2, 3, ... (8.51)
and the corresponding discrete energy levels is given by
E
n
=
Z
2
me
4
2
2
n
2
= Z
2
1
n
2
for n = 1, 2, 3, ... (8.52)
84 8. THE SCHR

ODINGER EQUATION IN SPHERICAL COORDINATES


Another important result from this analytical method is obtained when we combine (8.38)
with
o
= 2n, giving
=
me
2
4
o

2
1
n
=
1
a
o
1
n
(8.53)
for which it is easy to say that
a
o
=
4
o

2
me
2
= 0.53

A (8.54)
which is well-known as the Bohr radius. In (3.5), the Bohr radius is written in terms of the
cgs unit so that 4
o
disappears (1/4
o
= 1). It is left for the students to prove that (8.54)
is the most possible location of the hydrogens electron. By deriving this, the students will
do understand why the general description of the hydrogen appears to be in a single shell K
with a single electron moving around the nucleus.
It also follows from (8.38) and (8.53) that = r/a
o
n, and hence u() u(r) = rR(r).
With the help of (8.44), we can thus write
R
n
(r) =
1
r
u() =
1
r
e

+1
F() (8.55)
where R
n
(r) is the radial component of the wavefunction. It is clear that the radial solution
is determined by both n and , as r depends upon n. Now apart from the normalisation
constant, the F() can be written as
F() = /
2+1
n1
(2) (8.56)
where /
2+1
n1
(2) can be calculated using
/
p
qp
(x) = (1)
p
_
d
dx
_
p
/
q
(x) (8.57)
where /
p
qp
(x) is the associated Laguerre polynomial, determined by the values of the
Laguerre polynomial /
q
(x) dened as
/
q
(x) = e
x
_
d
dx
_
q
e
x
x
q
(8.58)
The rst few values of the associated Laguerre polynomials (8.57) are /
0
0
= 1, /
0
1
= x + 1,
/
1
0
= 1 and /
1
1
= 2x + 4.
After a very long discussion, we here conclude that the spatial wavefunction given in (8.8)
can be re-written as

n m
(r, , ) = R
n
(r) Y
m

(, ) (8.59)
8.2. The Hydrogen Atom 85
with the complete form of the stationary, normalised hydrogen wavefunctions are

n m
(r, , ) =

_
2
na
o
_
3
(n 1) !
2n[ (n + ) ! ]
3
e
r/na
o
_
2r
na
o
_

/
2+1
n1
Y
m

(, ) (8.60)
where Y
m

(, ) is the spherical harmonics given by (8.28). The wavefunction given by (8.60)


satises the following normalisation procedure,
_

[
n m
(r, , ) [
2
dV = 1
_
space

m
(r, , )
n m
(r, , ) r
2
sin dr d d =
nn


mm

(8.61)
Now, it is obvious that three quantum numbers are at least required for characterising the
hydrogen wavefunction. They are called as the principal quantum number n, the azimuthal
quantum number and the magnetic quantum number m, respectively dened as
n = 1, 2, 3, . . . . . .
= 0, 1, 2, . . . . . .
m = 0, 1, 2, . . . . . .
(8.62)
For an arbitrary number n, the possible values of are = 0, 1, 2, . . . . . . , n1. For each value
of , there are (2 + 1) possible values of m as m runs from to through zero. Hence, the
total degeneracy of the energy level E
n
of the hydrogen atom is
T
n
=
n1

=0
2 + 1 = n
2
(8.63)
It follows that for a particular n there will be n
2
eigenstates with the same level of energy.
For example, the ground state level (i.e., n = 1) has energy E
1
= -13.6 eV and is associated
with a single wavefunction
100
. Then, the rst excited state, n = 2, has energy E
2
= -3.4 eV
and is associated with four wavefunctions
200
,
211
,
210
and
211
.
Although we here have already obtained the explicit form of the normalised hydrogen
wavefunctions with three quantum numbers attached, as well as the values of these numbers
and the constraints on them, there remains unsolved questions. Firstly, when splitting the
stationary wavefunction into radial and angular parts (see 8.11 and 8.12) it was given without
proof that the RHS of both the equations was in the form of ( +1). Secondly, when dividing
the angular part into two components (see 8.15 and 8.16), it was again given without proof
that the RHS of both the equations was in the form of m
2
. These will be clear when we
consider the eigenvalues and eigenstates of the angular momentum operator.
86 8. THE SCHR

ODINGER EQUATION IN SPHERICAL COORDINATES


The signicance of angular momentum in classical physics is that it is, together with linear
momentum and energy, one of the fundamental constants of motion for an isolated system. As
we will nd later, the counter-part of this property also holds for isolated quantum mechanical
systems. The angular momentum itself, based on the dynamics of a rotating object around
the centre of the motion, can be separated into two components, namely orbital and spin
angular momenta. An orbital motion for macroscopic objects is something like the Earth
moves around the Sun as the central point whereas spin is associated with the rotation of the
Earth about its own axis. Thus, the two motions are of great importance hence needs to be
discussed separately rst before considering them as a combined motion.
Orbital angular momentum is derived from the space and linear momentum coordinates
of a particle and closely akin to classical angular momentum. Contrary to this, spin angular
momentum does not relate to the particles space and linear momentum coordinates. Spin
is, as earlier mentioned, an internal, intrinsic property of a particle, such as inertial mass or
electric charge. It is an extra degree of freedom attached to a quantum mechanical particle,
and must be prescribed together with the values of all the other three quantum numbers in
order to designate properly the particular state of the particle. Thus, the three quantum
numbers obtained are in fact not enough to describe fully the dynamics of the hydrogen atom.
What is needed now is the special discussion on the characteristics of angular momentum,
particularly the properties of orbital and spin angular momenta. We focus on the derivation
of eigenvalues and eigenfunctions of the angular momentum. This will be discussed in terms
of operators.
8.3 Angular Momentum
Classically, angular momentum L of a particle of mass m depends upon the particles linear
momentum p and its displacement from the origin r. It is mathematically written as
L = r p (8.64)
If we speak about a system of particles, then we may add the angular momentum of all
particles in such a system to obtain the total angular momentum of that system.
8.3.1 Orbital Angular Momentum
8.3.2 Spin Angular Momentum
8.3.3 Total Angular Momentum
8.3. Angular Momentum 87
F
i
g
u
r
e
8
.
1
:
T
h
e

r
s
t
f
e
w
e
x
a
m
p
l
e
s
(
f
o
r
n
=
1
,
2
,
a
n
d
3
)
o
f
t
h
e
s
t
a
t
i
o
n
a
r
y
,
n
o
r
m
a
l
i
s
e
d
r
a
d
i
a
l
,
a
n
g
u
l
a
r
a
n
d
t
o
t
a
l
w
a
v
e
f
u
n
c
t
i
o
n
s
o
f
t
h
e
h
y
d
r
o
g
e
n
a
t
o
m
(
t
a
k
e
n
f
r
o
m
t
h
e
s
a
m
e
t
a
b
l
e
o
f
B
e
i
s
e
r
,
1
9
8
8
)
.
88 8. THE SCHR

ODINGER EQUATION IN SPHERICAL COORDINATES


8.4 Problems
1. (a) Normalise R
10
(r) and construct the complete hydrogen wavefunction
100
(r, , ).
(b) Normalise R
20
(r) and construct the complete hydrogen wavefunction
200
(r, , ).
(c) Normalise R
21
(r) and construct the complete hydrogen wavefunction
211
(r, , ),

210
(r, , ) and
211
(r, , ).
(taken from Problem 4.11, page 140, Chapter 4, Griths, 1995).
2. Find r and r
2
for an electron in the ground state of the hydrogen atom. Express
your answer in terms of the Bohr radius a
o
.
(taken from Problem 4.13, page 142, Chapter 4, Griths, 1995).
3. The Laplacian operator
2
in cylindrical coordinates appears as

2
=
1

_
+
1

2
+

2
z
2
Consider now the cylindrical potential well given by
V () = 0, < a
V () = , a
What is the time-independent Schrodinger equation for an arbitrary potential V (, , z)
in cylindrical coordinates?
(taken from Problem 10.15, page 379, Chapter 10, Libo, 1980).
4. Show that u
2
10
(r) has its maximum at r = a
o
, the Bohr radius.
(taken from Problem 10.32, page 405, Chapter 10, Libo, 1980).
5. In what sense does the Bohr analysis of the hydrogen atom give erroneous results for
the magnitude of angular momentum L?
(taken from Problem 10.39, page 407, Chapter 10, Libo, 1980).
6. The initial state of a hydrogen atom is given by
(r, 0) =
1

1s
+ A
2s
+
1

3s
(a) Find the constant A so that the sate is normalised.
(b) What is the average energy of the state?
(taken from Example 12.4, page 352, Chapter 12, McMahon, 2005).
Bibliography
Beiser, A. 1988 Perspective of Modern Physics. London, UK: McGraw-Hill.
Gasiorowicz, S. 1996 Quantum Physics. New York, US: John Wiley and Sons.
Griffiths, D. J. 1995 Introduction to Quantum Mechanics. New Jersey, US: Prentice Hall.
Harris, R. 2007 Modern Physics. California, US: Pearson Addison-Wesley.
Liboff, R. L. 1980 Introductory Quantum Mechanics. Reading, US: Addison-Wesley.
McMahon, D. 2005 Quantum Mechanics demystied. New York, US: McGraw-Hill.
McMahon, D. 2006 Relativity Demystied. New York, US: McGraw-Hill.
Morin, D. 2003 Introductory Classical Mechanics. New Jersey, US: Prentice Hall.
Serway, R. A., Moses, C. J. & Moyer, C. A. 2005 Modern Physics. California, US:
Thomson Learning Inc.
Tipler, P. A. 1999 Physics for Scientists and Engineers. New York, US: W. H. Freemann.
89

Das könnte Ihnen auch gefallen