Está en la página 1de 99

Solid State Electronics

Neil Broderick
March 12, 2003
Contents
1 Introduction 1
1.1 Units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Assumed Mathematics and Physics . . . . . . . . . . . . . . . . . . . . . . . . 2
1.4 List of Symbols Used . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.5 Problem Set 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2 Introduction to Quantum Mechanics 7
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 The Hydrogen Atom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.3 The Schrodinger Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.4 Solutions to the Schrodinger equation . . . . . . . . . . . . . . . . . . . . . . 14
2.4.1 Free Particle Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.4.2 Particle in an innite square well . . . . . . . . . . . . . . . . . . . . . 15
2.4.3 Finite Square Well . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.4.4 The Hydrogen Atom . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.4.5 The Story so far . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.5 Quantum tunnelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.5.1 A single potential barrier . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.5.2 Multiple Potential Barriers . . . . . . . . . . . . . . . . . . . . . . . . 23
2.6 Periodic Potentials in One dimension . . . . . . . . . . . . . . . . . . . . . . 23
2.7 Density of States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.7.1 The Density of states for a three dimensional potential well . . . . . . 29
2.8 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.9 Further Reading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.10 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3 Electrons in Crystals 34
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.2 Atomic Bonds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.2.1 Van der Waals Attraction . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.2.2 Ionic Bonds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.2.3 Metallic Bonds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.2.4 Covalent Bonds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.3 Atomic Energy Levels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.4 Crystal Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
i
3.4.1 Three Dimensional Crystal Types . . . . . . . . . . . . . . . . . . . . 39
3.5 The Fermi-Dirac Distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.6 Electrons and Holes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.7 Carrier concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.7.1 Number density of holes . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.7.2 Why we care . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.7.3 Calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.7.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.8 The motion of an electron in a electric eld . . . . . . . . . . . . . . . . . . . 53
3.8.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.9 The Hall Eect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.10 Carrier diusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.10.1 Charge Continuity Equations . . . . . . . . . . . . . . . . . . . . . . . 61
3.10.2 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.11 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4 Junctions 66
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.2 Junctions at Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.2.1 The story so far . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.3 Width of the Charge depletion Region . . . . . . . . . . . . . . . . . . . . . . 73
4.4 PN Junctions under an applied bias . . . . . . . . . . . . . . . . . . . . . . . 77
4.5 Behaviour of a diode under reverse voltage . . . . . . . . . . . . . . . . . . . . 83
4.5.1 Zener Tunneling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
4.5.2 Avalanche Breakdown . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
4.6 Transient Eects in Diodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
4.7 Capacitance of p-n Junctions . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.8 Junctions between two dissimilar materials . . . . . . . . . . . . . . . . . . . 89
4.8.1 Junctions between a metal and a semiconductor . . . . . . . . . . . . 89
4.8.2 Ohmic Contacts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.8.3 A Junction between two dissimilar semiconductors . . . . . . . . . . . 91
4.9 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
5 Interactions between light and matter 94
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
ii
Module 1
Introduction
These notes are designed for the rst year course on solid state electronics for engineers. As
such they are not meant to be read in isolation but rather I have intended that they should
read alongside the lectures. Some things will be in the lectures which are not in the notes
and there will probably be a lot in the notes which doesnt get covered in the lectures.
The aim of this course is to explain the electrical characteristics of a diode. You should
all be familar with the basic properties of a diode, i.e. they only allow current to pass in
one direction, however understanding this behaviour relys on many dierent areas of physics,
such as quantum mechanics, statistical mechanics. crystalography etc. Hence we will need to
cover all these areas in some detail and in isolation before we can put the material together.
And hopefully once you have understood the material presented here it should be relatively
easy to apply it to more complicated devices such as bi-polar transistors.
1.1 Units
In this course I will use SI units where-ever possible. The only exception will be that I will
often measure energy in electron volts (eV) rather than in Joules (J). One electron volt is
dened as the energy obtained by an electron after moving through a potential dierence of
1 Volt. Thus we have
1eV = 1.6 10
19
J. (1.1)
Electron volts are useful as one electron volt is the typical amount of energy an electron will
have in this course. For example it is much easier to remember that the band gap of silicon
is 1.1 eV than it is to remember its value in Joules. Hence, if when solving a problem you
obtain an answer that is substantially dierent from 1eV you should check it carefully as it
is likely to be wrong.
In this course all temperatures will be measured in Kelvins. 0K is absolute zero and it is
not possible to have a temperature lower than this. Measured in degrees celsius absolute zero
is equal to 273.15
o
C. To convert from the celsius scale to the absolute scale you simplely
add 273.15. Generally I will take 300K to be room temperature and it is a easy value to
remember. Those of you who would like to know more can look at http://www.sun.rhbnc.
ac.uk/~uhap057/LTWeb/Absolute.html or in any introductory thermodynamics textbook.
Lastly many textbooks in this eld measure length and thus volume in centimetres and
centimetres cubed rather than metres and metres cubed. Of course it is easy to convert
between the two as 1 cm
3
= 10
6
m
3
.
1
1.2 Notation
In general any symbol in a bold font e.g. k indicates a vector, i.e. a quantity that has both a
direction and magnitude. The magnitude of a vector will be indicated by either |k| or just k.
In places in the text you will see expressions like n Z which means that n is an integer.
Z is commonly used to represent the set of all integers. Similarly Z
+
is the set of positive
integers, R the set of all real numbers and C the set of all complex numbers.
In this course I will write complex numbers as
z = x +iy (1.2)
where i
2
= 1. When I refer to currents I will either use a capital I or it should be clear
from the context whether I am talking about currents or complex numbers. If you read other
electronics textbooks you will nd that a number of them use the symbol j to denote the
square root of 1. However in this course j represents the current density.
Lastly during the course we will deal with both positively and negatively charged particles,
i.e. electrons and holes. I will denote the charge on an electron by q and the charge on a
hole q where
q = 1.602 10
19
C (1.3)
1.3 Assumed Mathematics and Physics
You will need to be familiar with complex numbers to pass this course. The use of complex
numbers is restricted mostly to describing waves and hence you will need to very familiar
with de Morives theorem that:
e
iz
= cos(z) +i sin(z), z C (1.4)
More likely you have seen de Morives theorem presented for a real number x as
e
ix
= cos(x) +i sin(x), x R (1.5)
If we now write z = x +iy and substitute this into Eq. (1.4) we nd
e
iz
= e
i(x+iy)
(1.6)
= e
ixy
(1.7)
= e
y
e
ix
(1.8)
= e
y
[cos(x) +i sin(x)] . (1.9)
Probably one of the most used pieces of information in this course is the solution to the
dierential equation
d
2
y
dx
2
= y (1.10)
where is a complex number. You should be able to immediately see that the solution
y = A
1
e

x
+A
2
e

x
(1.11)
is correct and where A
1
and A
2
are two constants of integration which can be found by
applying the appropriate initial conditions.
2
In Eq. (1.10) I made no assumptions about whatsoever. There are two important cases
of Eq. (1.10). The rst is when =
2
in which case Eq. (1.10) describes the process of
simple harmonic motion which you covered last semester. The other case is when = k
2
where k is a positive real number. In this case the solutions are exponentially growing or
decaying and describe processes like radioactive decay or light propagation in a amplier.
However by writing the solutions in terms of complex exponentials I will not need to worry
about whether or not the solutions are oscillating or growing exponentially or a mixture of
both.
The last piece of mathematics you will need to be familiar with is the complex represen-
tation of a plane wave. In this course I will represent a plane wave by
y(x, y, z) = y(r) = e
ikr
(1.12)
where r = (x, y, z) represents the spatial co-ordinates in three dimension. k is the wavevector
dened by
k = (k
x
, k
y
, k
z
). (1.13)
From Eq. (1.12) and Eq. (1.13) it can be easily seen that
y(x, y, z) = y(x + 2/k
x
, k
y
, k
z
) (1.14)
as
e
2i
= 1 (1.15)
and thus the wavelength of the wave in the x-direction is 2/k
x
. In Eq. (1.12) the direction
of propagation of the wave is given by the unit vector k/|k|. If I want to describe a wave
propagating in the opposite direction I will write
y(r) = e
ikr
(1.16)
Comparing Eq. (1.11) and Eq. (1.12) we see that the general solution to the simple har-
monic oscillator problem is a superposition of two plane waves travelling in opposite directions.
1.4 List of Symbols Used
i The square root of -1.
q The magnitude of the charge of an electron. q = 1.602 10
19
C.
k, k The wavenumber/wavevector of a particle.
k Boltzmanns constant.
T Temperature in degrees Kelvin.
h Plancks constant.
Plancks constant divided by 2.
The angular frequency of a wave.
3
Wavelength of a particle or a photon.
The wavefunction of a particle.
V The potential (either electrical or quantum mechanically).
m

, m
eff
The eective mass of a particle in a crystal.
D(E) The density of states for a particle with energy E.
n The number density of electrons in the conduction band.
p The number density of holes in the valance band.
F(E) The Fermi-Dirac distribution.
E
f
The Fermi level.
E
c
The energy at the bottom of the conduction band.
E
v
The energy at the top of the valance band.
E
i
The intrinsic Fermi level.
N
c
The eective density of states for the conduction band.
N
v
The eective density of states for the valance band.
n
i
The intrinsic number density.
N
a
The number density of acceptor atoms.
N
d
The number density of donor atoms.

n,p
The mobility of electrons or holes.
v
th
The thermal velocity of a particle.
The conductivity of a crystal.
= 1/ The resistivity of a crystal.
The mean free time between collisions in a crystal.
J = J
n
+J
p
The total current density.
J
n
The electron current density.
J
p
The hole current density.
D
n,p
= kT/q
n,p
The electron/hole diusion coecient.
p, n Excess hole/electron distributions.
V
0
The equilbrium contact potential for a pn diode.
p
p
Hole distribution on the p-side of a pn junction.
4
p
n
Hole distribution on the n-side of a pn junction.
n
p
Electron distribution on the p-side of a pn junction.
n
n
Electron distribution on the n-side of a pn junction.
W = x
n
+x
p
Width of the charge depletion region in a pn junction.
x
n
Width of the charge depletion region on the n-side of a junction.
x
p
Width of the charge depletion region on the p-side of a junction.
E
max
Maximum strength of the inbuilt electric eld in a pn junction.
V = V
0
+V
f,a
Total voltage drop across a pn junction.
V
a,f
Applied voltage across a diode.
C Capacitance of a device.
5
1.5 Problem Set 1
1. The density of water is 1 g.cm
3
what is the value in kilograms per cubic metre?
2. Using Einsteins relation E = mc
2
where c is the speed of light what is the energy of
an electron at rest in both Joules and Electron volts?
3. How many silicon atoms are there in a cubic metre of crystalline silicon with a density
of 2.33 g.cm
3
?
4. Using the Euler identity e
2i
= 1 nd the 5 solutions to the equation
x
5
= 1
and plot them on a graph.
5. Find three independent solutions to the dierential equation
d
3
f
dt
3
+f = 0
6. Using the internet or some other source of information, nd out what you can about
Moores law and why it is relevant to this course.
6
Module 2
Introduction to Quantum
Mechanics
Learning Objectives:
By the end of this module on quantum mechanics it is expected that you should be capable
of the following:
1. Discussing the failings of classical mechanics and how these were solved.
2. Explaining what is meant by a particles wavelength.
3. Solving the Schrodinger equation for simple potential wells.
4. Discussing the nature of the solutions to Schrodinger equation for periodic potentials
and relate this to the motion of electrons in semiconductors.
5. Calculating the density of states for an electron in a 3D crystal.
2.1 Introduction
The main aim of this module is to provide a short introduction to quantum mechanics. Much
of this material is not directly relevant to the rest of the course but is essential in order to
fully understand the behaviour of semi-conductor devices. Also, hopefully, some of you will
go on design the next generation of semiconductor devices and computers and in order to do
so you will need to understand how an electron behaves. There are two common approaches
to teaching quantum mechanics the rst is a historical approach starting with the failure of
classical mechanics and the second is a more axiomatic approach. The historical approach is
the one I will adopt here since hopefully much of the introductory material will be familiar
to you.
The foundations of quantum mechanics were laid at the end of the 19th Century by Max
Planck. Planck was working on the problem of predicting the spectrum of light produced when
you heat objects (known as the black body problem). As is well known if you heat a piece
of metal it rst glows red then yellow and then eventually white. Standard thermodynamics
had predicted that such a body should radiate an innite amount of light a result which was
clearly wrong. In order to resolve this Planck made the then extremely unlikely suggestion
7
Figure 2.1: Graphs of the blackbody spectrum for dierent temperatures.
that light was emitted in discrete lumps or quanta. The amount of energy emitted in one
quanta was
E = h (2.1)
where h is Plancks constant and is the frequency of the light. More commonly nowadays
Eq. (2.1) is written as
E = (2.2)
where = h/(2) and is the angular frequency. In SI units h = 6.626 10
34
Js and
= 1.05410
34
Js. It will be important to remember the denition of since it will be used
constantly in the course.
Once Planck had made this assumption he found that he could solve the black body
problem and obtain excellent agreement between his theory and the experiments. For those
of you who are interested the formula for the radiation produced by a blackbody is given by
S(, T) = 2

c
3
h
e
h/kT
1
(2.3)
where S(, T) is the energy density per unit volume and unit solid angle, c is the speed of
light, k is Boltzmans constant
1
and T is the temperature. This is written down for the sake
of completeness and it will not be used in this course. Fig. 2.1 shows the black body spectra
for several dierent temperatures. Incidently Planck himself did not believe in the existence
of quanta but just thought they were a useful trick for solving the problem.
The next piece in the puzzle was provided by the photo-electric eect. In this experi-
ment as shown schematically in Fig. 2.2 scientists measured the current produced when light
illuminated the surface of a metal electrode. They noticed the following key points
Electrons were only produced when the frequency was greater than a certain frequency.
Increasing the intensity of light increased the number of electrons produced but did not
increase their energy.
Increasing the frequency of the illuminating light increased the energy of emitted elec-
trons.
8
Figure 2.2: Schematic of a setup to measure the photo-electric eect.
Again these experimental results did not have any ready explanation within classical
physics. In 1904 Albert Einstein extended Plancks ideas and suggested that not only was
light emitted in discrete quanta but that it could also only be absorbed in discrete quanta.
Thus the energy of the emitted electron was given by
E = (2.4)
where is the amount of energy needed to escape the surface of the metal. It can easily be
seen that that Eq. 2.4 easily explains all of the experimental observations
2
.
In doing this Einstein created the idea of a photon. A photon is a discrete lump of light
with energy and momentum given by
E = (2.5)
p = k (2.6)
where k is the wavevector of the light and |k| = 2/ in free space. Confusingly however
light is still composed of waves which diract and interfere. Einstein did not try to resolve
this.
'
&
$
%
Example Problem: Calculate the energy and momentum of a photon with a
wavelength of 1m
Solution: A photon with a wavelength of 1 m has a frequency of
= 2c/ = 1.88 10
15
s
1
. Thus its energy is 1.98 10
19
J. Alterna-
tively we can write this in electron volts as 1.23eV.
The momentum of a photon is given by p = k = 2/ = 6.62 10
28
kg.m.s
1
.
Thus you can see that a photon has comparatively little momentum compared to
its energy.
Following Einsteins lead a number of scientists in the early part of the 20th century tried
to develop ad-hoc methods of quantisation to treat various problems. Perhaps the most
successful was Bohr whose model of the hydrogen atom we will look at next.
1
We will come across Boltzmans constant again later on in the course and I will dene it properly then.
2
To explain the 2nd observation you need the additional piece of information that the intensity of light is
proportional to the number of photons
9
2.2 The Hydrogen Atom
By the end of the 19th Century it was well established that when you heated up a pure
sample of an element e.g. hydrogen or sodium it would emit light only at a number of precise
wavelengths. Similarly if you passed white light through some pure gas then it would only
absorb light at these frequencies. Today this is the basis of spectrospecy whereby molecules
are identied by their absorption lines. Scientists were naturally keen to understand this and
also to discover the internal structure of atoms. Since Thompsons discovery of the electron
it was known that atoms must posses some internal structure but no-one really knew what.
Rutherford in his scattering experiments had shown that atoms were in fact mostly empty
space with a tiny positively charged nucleus. But it was not clear where the electrons were
located
3
One of the rst proposed structures for the atom was the solar system model. As the name
suggests it was modelled on the solar system and consisted of a central positively charged
nucleus around which electrons orbited. However such a model was in direct violation of
classical electrodynamics which predicts that accelerating electrons emit radiation and hence
the solar system model would be unstable as the electron would constant lose energy and
rapidly spiral into the nucleus.
Niels Bohr then made the assumption that there existed certain orbits for which an elec-
tron would not radiate and would be stable. More precisely he suggested that the angular
momentum of the electron was quantised and that only orbits with an integral value of the
angular momentum were permitted. The angular momentum of a electron with velocity v
orbiting with a radius r is given by
I = mvr (2.7)
where m is the mass of an electron and I is the angular mometum. Bohr proposed the
following quantisation rule:
mvr = n (2.8)
where n is a positive integer. In addition Bohr proposed that an electron could only jump
from level to another by emitting or absorbing a photon with the requisite amount of energy.
From these two ideas we can nd an approximate expression for the energy levels in Hydrogen
(which you did last semester I hope). The end result is that the energy of the nth level of the
Hydrogen atom is given by
E
n
=
mq
4
32
2

2
0
n
2

2
(2.9)
where E is negative by convention as the electron is in a bound state. The agreement between
this model and the experimental results was quite good considering the relative simplicity of
the model. What the model does fail to do is explain why such privileged orbits might exist
in the rst place. It then became obvious that it was hard to go much further without a
better understanding of what quantities might be quantised and how.
The next main step came from de Broglie who, in 1924, made the quite remarkable
suggestion that if light can behave as particles then particles could behave as waves. Going
back to Einsteins equation for the momentum of a photon in terms of its wavelength [Eq. (2.6)]
3
At this time no-one knew anything about neutrons and so people had suggested that the nucleus was
composed of protons and electrons only.
10
he suggested that it could be inverted to give
=
h
|p|
. (2.10)
We now refer to the wavelength of a particle as the de Broglie wavelength in his honour.
Nowadays everyone should be familiar with the idea of an electrons wavelength as it is the
basis of devices such as the electron microscope.
'
&
$
%
Example Problem: An electron is accelerated from rest across a voltage of 10kV,
what is its nal wavelength?
Solution: The wavelength can be determined from Eq. (2.10) once we know the nal
velocity of the electron. From the conservation of energy we know that the nal
kinetic energy of the electron 1/2mv
2
must equal the loss in potential energy eV
and thus
1
2
mv
2
= eV, or v =
_
2eV
m
and hence the wavelength is given by
=
h
mv
=
h
m
_
m
2eV
=
h

2meV
=
6.63 10
34
J s
_
2(9.11 10
31
kg)(1.60 10
19
C)(10
4
V)
= 1.23 10
11
m
= 0.0123 nm
This value is much smaller than a typical optical wavelength for visible light (
500nm) and explains why electron microscopes have a better resolving power than
optical microscopes.
After de Broglies suggestion the basis of modern quantum mechanics (or at least non-
relevatistic quantum mechanics) was almost complete. It was Schrodinger who completed
the picture with his wave equation which described how the de Broglie waves propagated.
2.3 The Schr odinger Equation
To understand the Schrodinger Equation it is perhaps useful to take another look at classical
mechanics. In the 19th Century it was well known that the energy of a particle was conserved.
Also in many situations the energy of a particle could be written as a sum of the potential
and kinetic energy. Following convention we will call this the Hamiltonian of the system and
write it in terms of the position and momentum rather than the position and velocity. Thus
we have
H(x, p) =
p
2
2m
+V (x) (2.11)
11
where the rst terms represents the kinetic energy while the second term represents the
potential energy. What is quite remarkable is the fact that if we know the Hamiltonian then
we know everything about the system. Also surprising is the fact that almost all physical
laws can be put into Hamiltonian form
4
. Given a Hamiltonian then it can be shown that the
equations of motion are given by
dx
dt
=
H
p
(2.12)
dp
dt
=
H
x
(2.13)
As an example consider the case of simple harmonic motion such a particle on a spring.
Such a particle feels a restoring force (i.e. Hookes law) given by
F = kx. (2.14)
In the standard way we can dene a potential V (x) as
V (x) =
_
x
0
kx

dx

(2.15)
=
1
2
kx
2
. (2.16)
Inserting Eq. (2.16) into Eq. (2.11) allows us to write the Hamiltonian as
H(x, p) =
p
2
2m
+
1
2
kx
2
. (2.17)
The equations of motion are then
.
x =
H
p
(2.18)
= 2p/(2m) (2.19)
= v (2.20)
so far no surprises. The equation of motion for the momentum reads
.
p =
H
x
(2.21)
= kx (2.22)
but the derivation of the momentum is just the force and hence we have recovered Hookes
law [Eq. (2.14)].
So what does all of this have to do with quantum mechanics? In fact the idea of a
Hamiltonian is central to all of modern physics and especially quantum mechanics. To make
the step from classical to quantum mechanics we need to think of the Hamiltonian as being
what is known as an operator which is basically just a fancy name for a function
5
. In quantum
mechanics the Hamiltonian H operates on what is known as the wave-function, commonly
4
The one exception is the general theory of relativity and this is one of the reasons why a quantum theory
of gravity has not yet been developed
5
Technically an operator is a function that acts on other functions
12
written as (x) which is a complex number in general. In its simplest form the Schrodinger
can be written as
H = E (2.23)
where E is the energy of the wave-function. Eq. (2.23) means that if we act on the wave-
function with the Hamiltonian then we obtain the energy of the wave-function. We will
write down a more precise form of the Hamiltonian later but for now lets concentrate on the
wave-function.
The wave-function is interpreted as giving us the probability of nding the particle at
a particular point. More precisely the probability, P, of nding the particle in the vicinity of
a point x
0
is given by
P = |(x
0
)|
2
dx. (2.24)
Clearly the particle must be located somewhere and hence
_

|(x)|
2
dx = 1. (2.25)
According to the axioms of quantum mechanics once you know the wave-function (x) then
you know everything that it is possible to know about the problem. For example if we want
to know the most likely position of the particle then we need to evaluate
_

x|(x)|
2
dx. (2.26)
In fact it can be shown that any information we might want can be obtained from a similar sort
of integral. This highlights the main dierences between classical and quantum mechanics i.e.
classical mechanics deals with certainties while quantum mechanics deals with probabilities.
To write down the Schrodinger equation explicitly Schrodinger proposed that you start
by writing down the classical Hamiltonian
H =
p
2
2m
+V (x) (2.27)
and make the substitution
6
p i

x
. (2.28)
This gives the following equation:

2
2m

x
2
+V (x) = E (2.29)
Technically the Schrodinger equation is what is known as an eigenvalue equation, i.e solutions
will in general only exist for certain values of the energy E
n
where E
n
is the nth eigenvalue. For
each allowed energy eigenvales there is a corresponding eigenfunction. Eigenvalue equations
are perhaps more easily understood when dealing with matrices and vectors. In a similar way
to Eq. (2.23) we can write a general matrix eigenvalue equation as
M v = v (2.30)
6
You will have to trust me on this one.
13
where v is referred to as the eigenvector and is the eigenvalue. More concretely consider
the matrix:
M =
_
2 0
0 3
_
(2.31)
then hopefully you should be able to see that the eigenvectors are (1, 0)
T
and (0, 1)
T
with
eigenvalues 2 and 3. It can be shown that for an invertible n n matrix there exists
n eigenvalues and hence n eigenvectors. Eigenvalue equations often have discrete solutions
which naturally lead to the idea of discrete energy levels which we will nd in quantum
mechanics. If you are still having problems with the idea of eigenvectors and eigenvalues then
another example would be vibration modes on a violin string (or other musical instrument).
As is well known a string will only vibrate at certain discrete frequencies which are the
eigenvalues. In this case the eigenvectors are the shapes of the string.
As a nal word I should mention that the solutions to the Schrodinger equation are often
called quantum states.
2.4 Solutions to the Schr odinger equation
2.4.1 Free Particle Solutions
Clearly the simplest example of the Schrodinger equation is when the potential V (x) is a
constant, V
0
say. In this case we can write the Schrodinger equation as

2
2m

x
2
+V
0
= E (2.32)
which we can rewrite as

x
2
= k
2
(2.33)
where
k =
_
2m

2
(E V
0
). (2.34)
This is just a case of simple harmonic motion and hence the general solution can be written
as a sum of a forward and backward propagating wave:
(x) = Ae
ikx
+Be
ikx
. (2.35)
Where A and B are complex constants determined by the boundary conditions. Note that the
solutions are very dierent depending on whether or not E > V
0
. If E < V
0
then the energy
is less than the potential energy and the solutions are exponentionally growing or decaying.
If E > V
0
then k is real and the particle behaves like a wave. In this case solutions exists for
all values of the energy such that E > V
0
. Note that in this case the normalisation given by
Eq. (2.25) does not apply. Also if we try to evaluate Eq. (2.26) for the average position we
nd that we get a nonsensical answer. The reason for this is that |(x)|
2
= constant so the
particle is equally likely to be anywhere which is what we would expect for a wave and hence
the name free particle.
Free particle solutions will be important later on in the course so it is important to notice
two things about the solution. Firstly inverting Eq. (2.34) gives us
E = V
0
+

2
2m
k
2
(2.36)
14
(a) Potential Well (b) Waveforms
Figure 2.3: (a) The innite potential well. The particle is conned between 0 and L. Fig.
(b) shows the rst three eigenvalues.
which is the equation for the energy of a free electron in terms of its wavenumber k. E only
depends on the magnitude of k
2
and hence identical particles moving in opposite directions
with equal speeds will have the same energy. This is what we expect and it is nice to know that
our model agrees with this. Also the plane wave solutions [Eq. (2.35)] has exactly the same
mathematical form as any other wave whether it is a water wave, radio wave or light wave
and hence everything you know about waves in general such as interference and diraction
can be applied to quantum mechanics. This is all I want to say about free particles for the
moment so instead let us turn to the case of bound particles.
2.4.2 Particle in an innite square well
Consider now the potential shown in Fig. 2.3 where we assume that the particle is bound
between 0 and L. At the edges (x) 0. In the well the solutions are given by Eq. (2.35)
i.e.
(x) = Ae
ikx
+Be
ikx
for x [0, L] (2.37)
Applying the boundary conditions (0) = (L) = 0 we get
0 = A+B, and (2.38a)
0 = Ae
ikL
+Be
ikL
. (2.38b)
From Eq. (2.38a) we nd that
A = B (2.39)
Using Eq. (2.39) and the fact that e
ix
= cos(x) +i sin(x) we nd that
(x) = 2iAsin(kx). (2.40)
We can make (x) purely real by setting
A =
1
2i
. (2.41)
15
Substituting Eq. (2.40) into Eq. (2.38b) gives
sin(kL) = 0 (2.42)
which has the solutions
k =
n
L
(2.43)
where n is a positive integer. Thus the nth eigenfunction can be written as

n
(x) = sin(
n
L
x), n Z
+
(2.44)
and the nth eigenvalue of nth energy level is given by
E
n
=

2
2m
k
2
(2.45)
=

2
2m
n
2

2
L
2
. (2.46)
Here we see that the energy levels are discrete i.e. there is a one to one correspondence
between the eigenvalues and the integers. For the sake of completeness I will mention here
the concept of orthogonality. Labelling the nth solution by
n
then the following property
holds
_

m
dx =
nm
(2.47)
where
nm
is the Kronecker delta function dened by

nm
=
_
1 if n = m,
0 otherwise.
(2.48)
Technically we say that the solutions to the Schrodinger equation form a complete orthogonal
set. This means that for an arbitrary function f(x) we can write
f(x) =

n
(x) (2.49)
where
n
are uniquely dened. Although we will not use this property it will be central to
any later courses you do involving quantum mechanics.
Moving from the obviously unphysical (but easy to solve) problem of an innite square
well we now turn to the case of a nite square well.
2.4.3 Finite Square Well
The potential for the nite square well is shown in Fig. 2.4. Here the well is located between
L/2 and L/2 and has a depth of V
0
. Unlike the free particle case here we are interested
in bound solutions which means that 0 > E > V
0
. In classical mechanics such a restriction
would mean that the particle was conned to the well. In the quantum regime the situation
is more complicated as we will see.
16
(a) Potential Well (b) Solutions
Figure 2.4: (a) Finite square well and (b) the rst two eigenvalues.
As in all three regions (I,II and III) the potential is constant we can immediately write
down the solutions in the three regions using the solutions given in Eq. (2.35) as

I
(x) = Ae
x
+e
x
(2.50a)

II
(x) = Be
ikx
+Ce
ikx
(2.50b)

III
(x) = De
x
+e
x
(2.50c)
where
=
_
2m

2
|E| (2.51)
k =
_
2m

2
(E V
0
). (2.52)
Imposing the requirement that (x) remains nite for all values of x immediately implies
that we can set = = 0 as otherwise the probability of nding the particle would increase
exponentially as we moved further away from the well.
Now as Schrodingers equation is a second order partial dierential equation both and
its rst derivative must be everywhere continuous. This leads to the following boundary
conditions:

I
(L/2) =
II
(L/2) (2.53a)

I
(L/2) =

II
(L/2) (2.53b)

II
(L/2) =
III
(L/2) (2.53c)

II
(L/2) =

III
(L/2) (2.53d)
where primes denote dierentiation with respect to x. Note that we have four boundary
conditions plus the normalisation condition for four unknowns A, B, C and D and thus in
order to satisfy all the boundary conditions some restrictions will be placed on the value of
E. These boundary conditions are quite general and we will use them whenever we need to
match solutions at an interface.
We can nd the solutions to the boundary conditions by using symmetry arguments. As
the potential well is symmetrical about the origin we expect that any solution will either be
17
an even or an odd function
7
. If the wavefunction is even then one can show that D = A
and B = C while for an odd wavefunction D = A and B = C. We these simplications
substituting Eqs. (2.50) into Eqs. (2.53) gives the following transcendental
8
equation for E
k tan(kL/2) = for even wave-functions (2.54)
k cotan(kL/2) = for odd wave-functions (2.55)
Fig. 2.4[b] shows the rst few waveforms for a typical square well. It can also be shown that
the number of solutions depends on the depth of the potential well however there is always at
least one bound state. From the pictures of the solutions along with the solutions in Regions
I and III it can be seen that outside the potential well (x) decays exponentially but is not
actually zero. What this means is that there is a small probability that the particle can be
found outside the potential well.
Although one dimensional quantum wells might seem a long way from anything real in
recent years engineers have been constructing square wells in solid state devices in order to be
able to manipulate the properties of electrons. Quantum well lasers are useful as they have
transitions in the mid infra-red while in other circumstances the use of multiple quantum wells
can enhance nonlinear interactions. Later on in the course we will look at how to fabricate a
quantum well.
I am now going to jump to another type of potential well, that seen by an electron in a
Hydrogen atom. This is to illustrate the ideas of quantisation in more detail and is usefully
in explaining many of the properties of dierent elements. We will then return to more
complicated one-dimensional potentials.
2.4.4 The Hydrogen Atom
In this section I want to briey look at the solutions to the Hydrogen atom. The formal
solution of the problem is outside the scope of this course however it is useful to look at the
solution. We start o with the Schrodinger equation in three dimensions

2
2m
_

2
(r)
x
2
+

2
(r)
y
2
+

2
(r)
z
2
+V (r)(r)
_
= E(r) (2.56)
For the Hydrogen atom we can write
V (r) =
1
4
0
q
2
|r|
. (2.57)
As that the potential is circularly symmetric it makes sense to solve it using spherical
coordinates r, and as dened in Fig. 2.5. We will also assume that the solution can be
written as
(r, , ) = R(r)()() (2.58)
In addition we can write Eq. (2.56) in spherical coordinates as

2
2m
_
1
r
2

r
_
r
2

r
_
+
1
r
2
sin

_
sin

_
+
1
r
2
sin
2

2
_
+V = E (2.59)
7
Even functions satisfy f(x) = f(x) while odd functions satisfy f(x) = f(x). Examples of even
functions are x
2
and cos(x) while examples of odd functions are x
3
and sin(x)
8
transcendental is just a big word meaning an equation we cant solve
18
Figure 2.5: Diagram showing the relationship between the spherical coordinates and Cartesian
coordinates.
Although it doesnt actually look it, Eq. (2.59) is actually easier to solve than Eq. (2.56).
Substituting Eq. (2.58) into Eq. (2.59) we can solve it using the method of separation of
variables. The method of seperation of variables assumes that the equation for one variable
is independent of the other equations and hence instead of solving one complicated equation
we solve three simple ones. The equation for gives
d
2

d
2
+
2
= 0 (2.60)
which has solutions
() = e
i
(2.61)
Now we would expect that the solutions should be periodic as is limited to between /2
and /2. In addition we expect that () = ( +2) and this condition restricts the value
of to being an integer. Thus the solution for is characterised by the quantum number l.
Once we have the solution for we can solve for which leads to another quantum number
m. Finally we get get the solution for R which gives us another quantum number n which is
the principal quantum number (corresponding roughly to the quantum number n in the Bohr
model of the atom). Putting these all together the energy of the electron is given by
E
n,m,l
=
e
4
2
2
n
2
(2.62)
where is a constant. Note that the energy levels do not depend on the angular quantum
numbers m and l since the potential is circularly symmetric. In fact there is an additional
quantum number s which describes the electron spin. The electron spin is an intrinsically
relativistic eect and to properly describe it you need to use the Dirac equation which is
beyond the scope of this course. The additional quantum numbers lead to splitting of the
energy levels when the atom is placed in a magnetic eld this is known as the Zeeman eect.
Finally it can be shown that the quantum numbers can take the following values
n = 1, 2, 3, 4, 5, .... (2.63a)
m = 0, 1, .., (n 1) (2.63b)
l = m, m+ 1, ..., m1, m (2.63c)
s =
1
2
(2.63d)
19
With these quantum numbers we can explain the spectrum of the hydrogen atom. In fact
it can be shown that electrons in any atom have exactly the same quantum numbers. To
explain the periodic table we now need one additional piece of information namely the Pauli
exclusion principle which states
Principle 1 Pauli Exclusion Principle: No two electrons can be in the same quantum state
at the same time.
Particles which obey the Pauli exclusion principle are called Fermions and obey Fermi-Dirac
statistics. Other examples of Fermions are protons and neutrons. Photons on the other hand
do not obey the Pauli exclusion principle and are called Bosons. Photons can be in the same
quantum state as other photons which is the basis behind the laser.
Using the Pauli exclusion principle it can be seen that the inner shell of an atom can
hold two electrons, one with quantum numbers n = 1, m = 0, l = 0, s = 1/2 and the other
with n = 1, m = 0, l = 0, s = 1/2. The dierence in energy between these two states
9
corresponds to radiation with a wavelength of 21cm which is a very important transition for
radio astronometers since neutral hydrogen occurs throughout the universe.
Once the n = 1 shell has been lled the resulting atom, Helium, is very stable and does
not form chemical bonds. In fact all atoms with a full outer shell are stable and correspond
to the noble gases. By using Eq. (2.63) it is possible to predict the atomic numbers of the
noble gases. Similar it can be shown that atoms with the same number of electrons in the
outer shell have similar chemical properties
10
.
The last thing I want to mention about electrons in the Hydrogen atom (or in fact electrons
in any material) is that they can only jump from one energy level to the next by gaining or
losing energy equal to the dierence in energy levels. In gases this is usually achieved by
absorbing or emitting a photon while in solids it can happen in a number of ways including
absorbing/emitting a photon or a sound wave (called a phonon) or by directly stimulating
another electron.
2.4.5 The Story so far
Hopefully you should now have a basic feel for what sort of calculations are involved in quantum
mechanics. Also you should be able to explain what is meant by a quantum state with reference
to either a hydrogen atom or an innite square well. The other crucial concept from this
section is the Pauli exclusion principle and you should be able to use this to explain the
structure of the periodic table
2.5 Quantum tunnelling
2.5.1 A single potential barrier
The next problem we want to look at is the case of quantum tunnelling through a potential
barrier like the one shown in Fig. 2.6. Here again like in Section 2.4.3 we can immediately
9
This energy dierence is a consequence of relativistic quantum mechanics hence I will not derive it here.
10
This is not strictly true for transition metals but we wont worry about that here.
20
Figure 2.6: Schematic of a potential barrier of height V
0
.
write down the general solution in the three sections as:

I
(x) = A
1
e
ik
1
x
+A
2
e
ik
1
x
(2.64)

II
(x) = B
1
e
ik
2
x
+B
2
e
ik
2
x
(2.65)

III
(x) = C
1
e
ik
1
x
+C
2
e
ik
1
x
(2.66)
(2.67)
where k
1
and k
2
are dened by Eq. (2.52) using the relevant dierence in energy levels. Note
that k
2
may be real or imaginary depending upon the dierence between E and V
0
. We can
simplify the system of equations by making the following assumptions. Firstly we will set
C
2
= 0 as we will assume that we initially have particles entering from the left and moving
towards the right. In this case the quantities of interest are A
2
/A
1
which denes the reection
from the barrier and C
1
/A
1
which denes the transmission through the barrier. To nd the
remaining unknown we need to use the boundary conditions at x = 0 and x = L. These are
of course are the same as in Eq. (2.53) and give
A
1
+A
2
= B
1
+B
2
(2.68)
k
1
(A
1
A
2
) = k
2
(B
1
B
2
) (2.69)
B
1
e
ik
2
L
+B
2
e
ik
2
L
= C
1
e
ik
1
L
(2.70)
k
2
_
B
1
e
ik
2
L
B
2
e
ik
2
L
_
= k
1
C
1
e
ik
1
L
(2.71)
We now have four equations for ve unknowns. However since the equations are linear we
can arbitrarily set any one of the unknowns to equal to unity and nd the rest. When we
solve them we obtain the following expressions for the reection and transmission:
ref =
_
1 +e
2ik
2
L
_
(k
2
2
k
2
1
)
(1 +e
2ik
2
L
) 2(1 +e
2ik
2
L
)k
1
k
2
+ (1 +e
2ik
2
L
) k
2
2
(2.72)
tran =
4e
i(k
1
k
2
)L
k
1
k
2
((1 +e
2ik
2
L
) 2(1 +e
2ik
2
L
)k
1
k
2
+ (1 +e
2ik
2
L
) k
2
2
(2.73)
these are plotted in Fig. 2.7. In the classical regime we would expect that if E < V
0
then
the transmission would be identically zero. However this is not true in the quantum regime
and instead we nd that even for the strongest barrier there is still some possibility that an
electron will tunnel through (although you might have to wait for hundreds if not thousands
of years to observe an electron).
21
5 10 15 20
Energy
0.2
0.4
0.6
0.8
1
Reflectivity
5 10 15 20
Energy
0.2
0.4
0.6
0.8
1
Transmission
Figure 2.7: (a) Reection coecient for a barrier of height 5eV . Fig. (b) shows the transmis-
sion through the barrier as a function of the electrons energy.
Quantum tunnelling is also becoming more and more important (or a bigger problem
depending on your point of view) in desiging integrated circuits. Remember that Moores law
predicts that the number of transistors on an IC will double every 18 months. This implies
that the distance between wires on a IC will decrease by a factor of

2 every 18 months.
Thus while in the past the wires were suciently far away so that electron tunnelling from
one to the other was unimportant in the near future it will be a major problem. Quantum
tunnelling is in fact quite likely to put a stop to Moores law and other methods will be
needed to continue the computer revolution. Already there is a growing interest in quantum
computers fueled by the realisation that minuraturisation can only go so far
11
.
Quantum tunnelling has already found its way into practical devices, namely the Zener
diode which uses quantum tunnelling to provide the breakdown current. But we will discuss
such diodes more in a later section.
Matrix method of solving the Schr odinger equation
The point of this digression is to give a very brief outline of how you can solve Schrodingers
equation for piecewise constant potentials. Take for example the interface between regions
I and II in Fig. 2.6 this interface leads to two boundary conditions given in Eq. (2.68) and
Eq. (2.69). It is trivial to write these two equations as one matrix equation
_
A
1
A
2
_
= M
1
_
B
1
B
2
_
(2.74)
where M
1
is a 2 2 complex matrix which relates the free particle solutions on one side of an
interface to the solutions on the other side. In general if we have n interfaces we can write
_
A
1
A
2
_
= M
1
M
2
M
n
_
B
1
B
2
_
(2.75)
_
A
1
A
2
_
= M
_
B
1
B
2
_
(2.76)
where M
i
is the matrix describing the ith interface and B
1,2
are the coecients of the free
particle solutions at the end of the problem.
11
Another reason for the interest in quantum computers is the fact that researchers have found ecient
algorithims which can run on quantum computers to crack public key cryptographic code.
22
Figure 2.8: Schematic of a double potential barrier.
In many problems of interest we can assume that we only have particles incident from the
left and hence B
2
= 0 since it is the coecient of an incoming plane wave from the right.
Also since the equations are linear we can set A
1
= 1 and what we would like to know is the
transmission and reection coecients t and r. Setting A
2
= r, B
1
= t, A
1
= 1 and B
2
= 0
in Eq. (2.76) we nd that
t =
1
M
11
(2.77)
r =
M
21
M
11
(2.78)
I followed this procedure to create most of the graphs in this section. This procedure is also
often used in optics in calculating the reection and transmission of periodic structures such
as Bragg gratings or photonic crystals where it is called the transfer matrix approach.
2.5.2 Multiple Potential Barriers
In the previous section we saw that the particle wave-function decays exponentially if E < V
0
.
Thus one might expect that if we add another potential barrier that the transmission would
decrease still further. In general this is true however as we will see this is not the complete
story. Consider now the case of the potential barrier shown in Fig. 2.8. By now you should
be able to write down immediately the general form of the wave-function in all the dierent
regions. Applying the boundary conditions leads to a set of linear equations which can be
solved. In Fig. 2.9 I show the transmission for a typical situation. Here we can see that
there is a narrow band where the transmission increases to unity even when E < V
0
. This is
something which our classical experience tells us does not happen with particles. It is thus an
intrinsic wave eect which is common to all wave equations where it be quantum mechanics,
acoustics, electromagnetic waves
12
or even water waves.
2.6 Periodic Potentials in One dimension
So far much of what we have looked at in this chapter has seemed quite remote from everyday
applications. However it has been necessary to build up the tecniques we need to look at real
situations. The aim of this course is to understand the behaviour of electrical components
12
The optical equivalent is called a Fabry-Perot cavity and it is very useful as a narrow band lter
23
2 4 6 8 10 12 14
Energy
0.2
0.4
0.6
0.8
1
Reflection
(a) Reection
2 4 6 8 10 12 14
Energy
0.2
0.4
0.6
0.8
1
Transmission
(b) Transmission
Figure 2.9: (a) Reection coecient for a double barrier of height 5i eV. Fig. (b) shows the
transmission through the barrier as a function of the electrons energy. These graphs should
be compared with those of Fig. 2.7.
Figure 2.10: Schematic of a periodical potential in one dimension.
such as diodes. This can only be done by understanding rst the properties of individual
electrons and the moving on to consider the properties of collections of electrons. Crucial
to the operation of diodes and other semiconductor devies is the crystal structure of the
underlying material typically silcon. The dening property of a crystal is its periodicity,
i.e. the atoms in a crystal are located at regularly spaced well dened positions. Thus an
electron propagating through a crystal will be attracted to every nucleus and will thus see
an periodic potential. Solving the Schrodinger equation for a full three dimensional periodic
potential is beyond the scope of this course however we can simplify matters by looking at a
one dimensional periodic potenial as shown in Fig. 2.10. This is known as the Kronig-Penney
model and is the simplest model of a crystal. It does have all of the features of a real crystal
that we will want in this course.
Unlike in the case of a single potential well when we were interested in the shape of the
wavefunction for periodic potentials we are more interested in the energy and momentum of
the wavefunctions. The energy is important for as we will see later only certain energy levels
are allowed and the energy dierence between these levels is what determines the electrical
properties of semiconductors (and indeed all crystals and metals). Similarly the momentum
tells us about the speed which electrons can respond to external inuences and for example
is necessary in predicting the resistence of a material. To understand these quantities let us
rst of all go back to the free particle eigenvalues given by Eq. (2.35). Instead of writing the
wavevector k as a function of the energy E we write E as a function of k i.e.
E = E
0
+

2
2m
k
2
(2.79)
24
Figure 2.11: Energy momentum diagram for a free electron.
and this is plotted in Fig. 2.11. Graphs like Fig. 2.11 are called energy momentum diagram (or
sometimes band diagrams) and equations like Eq. (2.79) are called dispersion relationships.
The thing to note is that after twice dierentiating Eq. (2.79) we obtain
d
2
E
dk
2
=

2
m
(2.80)
or
m =
2
/
d
2
E
dk
2
(2.81)
Thus given an arbitrary dispersion relationship i.e. a function E = E(k) we can dene the
eective mass as being:
m
eff
=
2
/

2
E
k
2
(2.82)
In Silicon the eective mass of an electron has dierent values and ranges from 0.98 m
e
to
0.19 m
e
where m
e
is the rest mass of an electron. For most of the rest of the course when
we talk about the mass of an electron we will really mean the eective mass of an electron.
To try and understand how these eective masses arise let us look at the Kronig-Penney
model in more detail. The goal of the analysis will be to obtain the energy-momentum
relationship for the Kronig-Penney model i.e. we want to obtain E = E(k). Although it
is possible to solve the Kronig-Penney model exactly it is beyond the scope of this course
although any decent book on quantum mechanics should show you how to solve it.
We can however try to obtain some feeling for the solutions by looking at the solutions
to the double square well. Recall that for a double square well the transmission spectrum
consisted of sharp resonances where transmission was possible interspersed by energy regions
where the transmission was exponentially small. Adding addition wells magnies these trends.
This is shown in Fig. 2.12 which shows the transmission through 500 potential wells of depth
5eV and length 2, i.e. 500 copies of the potential in Fig. 2.9. Here it can be clearly seen
that the transmission spectrum consists of regions where the transmission is almost unity and
other regions where the transmission drops almost to zero. The energy regions where the
25
Figure 2.12: Transmission spectrum for 500 potential wells
Figure 2.13: Energy-Momentum diagram for the Kronig-Penney model. The green lines show
the solutions for a free electron.
transmission drops to zero implies that there are no solutions to the Schrodinger equation
with these energies and thus we speak about a band-gap. If we were to inject an electron
with this energy into such a structure then it would be reected.
Adding addition quantum wells does not signicantly alter the transmission spectrum
from that in Fig. 2.12 and hence in going to the case of an innite periodic potential we
would expect that there we be energy regions where we would nd some solutions separated
by energy regions where there are no solutions. Extending this to three dimensions and real
crystals this means that we would expect to nd energy regions where there are no allowed
quantum states for an electron and other energy regions where electrons can exist.
Usually in quantum mechanics one does not plot the transmission spectrum of a crystal
but rather the energy-momentum diagram, i.e. E = E(k). For the Kronig-Penney model
this is shown in Fig. 2.13 where the existence of energy gaps can be clearly seen. Each of
the energy regions where no solutions exist correspond to a zero transmission region of the
Fig. 2.12.
Looking at Fig. 2.13 we can see that for each of the bands the eective mass [Eq. (2.82)]
changes depending on the energy. This is one reason why in real crystals there is a range of
eective masses.
An alternative way to look at the energy gaps is that they correspond to what is known
as Bragg scattering. When we looked at the transmission of a single potential well we saw
26
that even if E > 0 a small fraction of the wave-function was scattered by the well. Now in
fact it can easily be shown that an abrupt change in the potential will reect some part of
the wave-function. If we think for a moment about the scattered wave-function from two
adjacent wells then clearly they can add up either in phase or out of phase. If the scattered
waveforms add up in phase then the reection will grow with the number of wells. This is
known as Bragg scattering and was rst seen with X-Ray diraction in crystals. The Bragg
condition is given by
= 2d (2.83)
The Bragg condition is again a generic property of all waves and holds for X-Rays, electrons
in crystals or even light in an optical bre. The common property which unites all these
phenomena is that when the Bragg condition is satised all of the incident light/particles is
reected and cannot propagate through the medium.
More commonly in textbooks the energy-momentum diagram is plotted in what is known
as the reduced zone scheme. This was done in Fig. 2.13 where essentially we have folded the
wave-function solutions into the region between k = /d and k = /d. We are able to do
this because of the periodicity of the system. As the potential has a periodicity d i.e.
V (x +d) = V (x) (2.84)
we would expect that we would only need to solve the Schrodinger equation in the region
between 0 and d and use periodic boundary conditions that
(x) = (x +d)e
i
(2.85)
where I have stuck in an additional phase factor since the absolute phase of is undetermined.
In momentum space the equivalent of Eq. (2.85) is that statement that solutions which dier
by /d are equivalent. Thus if we know all the solutions between k = /d and k = /d
then we know all the possible solutions.
In Fig. 2.13 we can see that a crystal has a number of energy bands. Generally in any
crystal a number of bands will be full and these are called the valence bands. The next
empty band is called the conduction band. The energy dierence between the highest valence
band and the lowest conduction band is called the bandgap. In Fig. 2.13 we see that for
the Kronig-Penney model the lowest point in the conduction band occurs at the same point
in k space as the highest point in the valence band. Materials for which this is also true is
called a direct gap semiconductor. If this is not true the material is called an indirect gap
semiconductor.
Fig. 2.14 shows the energy-momentum diagrams for Si and for GaAs. Both of these are
considerable more complicated that the simple model we have seen previously however the
basic features are still present. Again we can see that there are valence and conduction bands
and whether or not they are direct or indirect.
2.7 Density of States
In this section we are going to count the number of possible solutions to the Schrodinger
Equation. To do this let us start with the simplest possible situation that of an innite
square well. Recall that the solutions are given by

n
(x) = A
0
e
i2nx/L
(2.86)
= A
0
e
ik
n
x
(2.87)
27
Figure 2.14: Electronic band structure of silicon, Ge and AlGe. The dierent points along
the bottom axis describe dierent directions of propagation through the crystal. Taken from
Sze.
and that the energy associated with the nth solution is
E
n
=

2
2m
k
2
n
(2.88)
What we would like to do is count the number of possible solutions with an energy less than
E
f
say (the reason for this will become clearly later). To do this it is in fact simpler to count
the number of states with a wavevector k < k
f
then use Eq. (2.86) to convert to energies.
Now we can see from Eq. (2.86) that k
n+1
k
n
= 2/L so in the interval [0, k
f
] the number
of states N(k
f
) is
N(k
f
) = k
f
/(2/L) (2.89)
=
L
2
k
f
(2.90)
now we want to write it in terms of the energy so we invert Eq. (2.88) to get
k =
_
2mE

2
(2.91)
and substitute this into Eq. (2.90) to get
N(E) =
L
2
_
2mE

2
(2.92)
which is essentially the answer we are looking for. However in most applications we would
like to avoid having answers that depend on the size of the crystal and so would really like
28
to know the number density which is just the number of electrons divided by the volume (in
this case the volume is just the length) of the sample. Also Eq. (2.92) tells us about the total
number of available states starting from a zero energy level. However again in practice we
are really only interested in the number of energy levels starting from a well dened non-zero
energy level such as the bottom of the conduction level. To get around these issues people
use what is call the density of states D(E) which is dened as
D(E) =
1
V
dN
dE
(2.93)
where V is the volume of the sample. Given the density of states the number of energy levels
N in a range E around a energy E
c
is given by
N = D(E
c
)E (2.94)
Now for the 1D innite square well discussed above D(E) is
D(E) =
1
2
_
2m

2
E
1/2
(2.95)
2.7.1 The Density of states for a three dimensional potential well
Eq. (2.95) gives us the density of states for a one dimensional innite square well. However it
is rare that we meet such animals in real life (or even universities) and to get a more useful
result we need to repeat the calculations but this time in three dimensions.
The rst we need to do is to dene the potential well the particle is sitting in. The 3D
analogue of a innite square well is a cube with sides of length L. We will require that the
wave-function (r) is periodic with period L, i.e (x) = (x + L). This can be achieved by
the waveform
(r) = A
0
e
i
2
L
(n
x
x+n
y
y+n
z
z)
(2.96)
= A
0
e
ikr
(2.97)
where n
x
, n
y
, n
z
Z and we have dened the wavevector k as
k =
2
L
(n
x
i +n
y
j +n
z
k) (2.98)
In addition we will now write the energy as
E = E
0
+

2
|k|
2
2m
eff
(2.99)
The dierences between Eq. (2.99) and Eq. (2.88) arise from the fact that we have included a
constant energy E
0
and we have used the eective mass of the electron [Eq. (2.82)] obtained
from the band structure of the crystal rather than the free mass. Thus Eq. (2.99) is the energy
of a electron in a crystal rather than a free electron. Solving Eq. (2.99) for the magnitude of
the electrons wavevector gives
|k| =
_
2m
eff
(E E
0
)

2
(2.100)
29
The next thing we need is to work out how far apart are the states. Remember in 1D
there was one state for ever 2/L we covered. In 3D the argument is slightly dierent but the
result is that there is one state for every (2/L)
3
covered
13
. Now again we want to count the
number of states with an energy less than or equal to E
f
. From Eq. (2.99) we see that this
is equivalent to counting states with a value of |k| < k
f
. In 3D this is just the denition of a
sphere with radius k
f
and we now want to know how many states there are in the volume of
a sphere with radius k
f
. This gives us
N(k
f
) =
4
3
(k
f
)
3
/(
2
L
)
3
2 (2.101)
=
L
3
3
2
(k
f
)
3
(2.102)
where I included an extra factor of 2 to take into account the electrons spin. Substituting
V = L
3
and Eq. (2.102) into Eq. (2.93) we nd that
D(E) =
1
2
2
_
2m
eff

2
_3
2 _
E E
0
(2.103)
Now the important thing about Eq. (2.103) is that in 3D the density of states is proportional
to the square root of the energy. We will use this result in a later section of the course so
dont worry too much if you dont understand it at the moment as hopefully it should be
clearer when we actually use it.
The last calculation we will do in this section is look at the density of states for electrons
in the valence band. Here the shape of the energy-momentum diagram is approximately that
of an upside down parabola and hence instead of Eq. (2.99) we can write
E = E
0


2
|k|
2
2m
eff
. (2.104)
Substituting this expression into Eq. (2.101) and repeating the above calculation we nd that
the density of states in the valence band is given by:
D(E) =
1
2
2
_
2m
eff

2
_3
2 _
E
0
E (2.105)
Here as you would expect the density of states decreases with increasing energy until you
reach the bottom of the band gap at which point the density of states becomes zero.
2.8 Discussion
In this chapter I have tried to give you a short introduction to quantum mechanics and
introduce the bits of it which we will need later on in the course. The essential piece of
information that I would like you to take away from this chapter is that electrons do not
always behave like little billiard balls but instead they can behave like waves. Furthermore
many of the properties of semiconductors can only be explained by the use of the wave picture.
13
This result follows from number of atoms in a unit cell for a simple cubic lattice which is something we
will cover in a few lectures time
30
Quantum mechanics of course covers a lot more than what I have included in this course. I
have left out perhaps one of the most infamous bits of quantum mechanics namely Heisenbergs
uncertainity relationship since we will not need it in the course however it is probably worth
stating here. The uncertainity principle states that
xp

2
(2.106)
where x is the uncertainity in the position and p is the uncertainity in the momentum.
This relationship is perhaps one of the more misunderstood equations in physics and especialy
seems to confuse philosphers. What the Eq. (2.106) means is that if a particle has a denite
position (i.e. x = 0) then it does not have a denite momentum. Similarly if the particle has
a denite momentum then it does not have a concrete position. This is a simple consequence
of the wave nature of particles, if a particle has a denite momentum then it is a perfect plane
wave and hence does not have a denite position after all can you say where a water wave
is?
Despite the Heisenbergs uncertainity relationship, quantum mechanics or to be more
precise quantum electrodynamics which deals with electrons and photons is the best theory
that there is. By best I mean that it gives more accurate predictions than any other physical
theory. The prime example of this is the ne structure constant which describes the splitting
of the energy levels of Hydrogen in a magnetic eld. The current best experimental value is
= 0.007297352533(27)
where the gure in brackets gives the error in the last two digits. The best theoretical estimate
is
= 0.00729735253186
which agrees exceedly well with the measured value. Richard Feynman said that this is
equalivalent to knowing the distances across the USA to within the width of a human hair!
2.9 Further Reading
For those of you who are interested in reading more about quantum mechanics I have tried to
list here some suitable books. Most are fairly mathematical but you should not let that put
you o as it is not really possible to fully understand quantum mechanics without working
through the details.
Introduction to quantum mechanics by Fench and Taylor
P. A. M. Dirac, The principles of Quantum Mechanics, Cambridge University Press,
1981 ISBN: 0198520115
31
2.10 Problems
1. In analysing the photoelectric eect, how can we be sure that each electron absorbs
only one photon?
2. A nucleus in an excited state emits a -ray photon of energy 1MeV.
(a) What is the photons frequency?
(b) What is the photons wavelength?
(c) How does this compare with typical nuclear radii (of the order of 10
15
m) ?
3. A photoelectric surface has a work function of 4.00eV. What is the maximum speed of
the photoelectroncs produced by light with a frequency of 3 10
15
Hz ?
4. Calculate the rst three energy levels for an electron in a quantum well of width 1nm
with innite walls.
5. For crystal diraction experiments, wavelengths of the order of 0.1nm are often appropi-
ate. Fine the energy (in Joules and electron Volts), for a particle with this wavelength
if the particle is
a) An electron
b) A photon
Can you comment on these answers?
6. Repeat the density of states calculation for an electron in a two dimensional periodical
potential box. Draw graphs of the density of states for electrons in one, two and three
dimensions.
7. Bound states in square wells: The aim of this problem is to prove that square well always
have at least one bound state. Recall that the eigenvalue equation for a symmetrical
wavefunction can be written as
k tan(kL/2) =
where
=
_
2m

2
|E|
and
k =
_
2m

2
(E V
0
).
a) Prove that the boundary conditions [Eq. (2.53)] plus the symmetry requirements lead
to the above equations for E.
b) The aim is now to try and solve this equation. We can do this graphically by plotting
the left hand side and the right hand side as a function of E in the interval [0, V
0
]
where V
0
is the depth of the potential well.
Step 1. Sketch the L.H.S. as a function of E.
Step 2. Sketch the R.H.S. as a function of E
Do they intersect? Can you prove that the curves will always intersect?
Repeat the process for the odd wavefunctions. What do you nd now?
32
This problem is actually quite general. In optics exactly the same proof can be used
to show that every optical bre contains at least one bound mode, that is light can
propagate along the bre.
8. A 10kg satellite circles the earth once every 2hr in an orbit having a radius of 8060km.
a) Assuming that Bohrs angular momentum postulate [Eq. (2.8) applies to satellites
just as it does to an electron in the Hydrogen atom, nd the quantum number of the
orbit of the satellite.
b) Show from Bohrs postulate and Newtons law of gravitation that the radius of an
earth-satellite orbit is directly proportional to the square of the quantum number, i.e.
r = kn
2
where k is the constant of proportionality.
c) Using the result from part (b) nd the distance between the orbit of the satellite in
this problem and its next allowed orbit. Calculate the numerical value.
d) Comment on the possibility of observing the seperation of the two adjacent orbits.
e) Do quantised and classical orbits correspond for this satellite? Which is the correct
method for calculating the orbits?
33
Module 3
Electrons in Crystals
Learning Outcomes:
By the end of this module you should be able to
1. Describe how the crystal structure of silicon arises from the valence electrons.
2. Explain the concept of a Fermi Level
3. Discuss how the Fermi level changes with temperature and doping
4. Be familiar with the concept of holes
5. Describe both drift and diusion currents and explain when each are important.
3.1 Introduction
In the previous chapter I introduced the basics of quantum mechanics and looked at how
it was possible to solve for the energy levels of the hydrogen atom. In this chapter I aim
to look at the energy levels of more complicated atoms and also to try and developed some
understanding of how atoms bond together to form crystals.
Understanding the crystal structure of semiconductors is important for a number of rea-
sons. Firstly many of the processing techniques for manipulating silicon wafers rely on the
crystal structure. Knowledge of the crystal structure is also crucial when trying to dope
semiconductors with dierent materials to make diodes and transistors. The crystal struc-
ture also explains why some materials are good emitters of light while others (like silicon) are
poor emitters of light.
3.2 Atomic Bonds
The rst thing we need to know about ensembles of atoms is how they join together and why
matter sticks together. There are four main types of bonds, Van der Waals, Ionic, Metallic
and Covalent and we will briey look at each of them. For those of you who have done any
chemistry this should be familiar so I will only briey discuss each type of bond.
34
Figure 3.1: Diagram of two dipoles showing relative distances.
3.2.1 Van der Waals Attraction
The Van der Waals attraction is the weakest bond and is responsible for the force between
neutral atoms and molecules. The basic principle is shown in Fig. 3.1 which represents two
neutral atoms as two dipoles separated by R. Now the force on one of the dipoles from the
other is given by
F =
1
4
0
_
q
2
R
2

q
2
(R x
2
+x
1
)
2
+
q
2
(R x
2
)
2
+
q
2
(R +x
1
)
2
_

q
2
4
0
6x
1
x
2
R
4
(3.1)
when R x
1
, x
2
. The Van der Waals force is responsible for the liquidication of Helium as
well as other gases such as O
2
. Due to the weak nature of the Van der Waals force it only
becomes dominant at low temperatures and is not relevant for most semiconductor processes.
3.2.2 Ionic Bonds
In an ionic compound such as sodium chloride (i.e. salt) one atom gives a number of electrons
to the other atom so that each atom now has a full outer shell. Clearly this can only happen
when the atoms are from opposite sides of the periodic table. The atomic distances between
ions in a ionic bond can be calculated by equating the distance at which the attractive force
is balanced by the repulsive force felt by the electrons.
3.2.3 Metallic Bonds
In all atoms the outer electrons are only loosely bound to the atom since most of the attractive
force between the electron and the core is shielded by the inner electrons. In a metal the outer
electrons are in fact free and can move about the material. You can thus think about a metal
as being a crystal of positive ions surrounded by a sea of free electrons. This free electron
model can explain many of the properties of metals (and also of semiconductors with some
modications). And we will look at it in more detail later. Note that this model explains
why metals are quite exible as there are no strongly oriented bonds compared to an ionic
compound.
35
Figure 3.2: Illustration of a covalent bond between two atoms.
3.2.4 Covalent Bonds
The last type of bond I want to discuss is the covalent bond. Here instead of one atom giving
electrons to another rather they share electrons. This is shown schematically in Fig. 3.2
which shows the case of silicon (or carbon) in two dimensions. From this picture we see that
covalent bonds have a strong directionality associated with them. As we will see later this
directionality leads to a variety of crystal types. Covalent bonds tend to form between atoms
in groups III,IV, and V of the periodic table. Typically such semiconductors are formed from
such materials e.g. silicon which is a type IV semiconductor of Gallium Arsenide which is a
type III-V semiconductor.
3.3 Atomic Energy Levels
In section 2.4.4 I showed that electron levels in a hydrogen atom could be characterised by
four quantum numbers: n, l, m and s, where n is the principle quantum number, l and m
relate to the angular momentum state of the orbit and s describes the intrinsic spin of the
electron. As stated earlier these four quantum numbers suce to specify the energy levels of
any atom. In addition I introduced the Pauli exclusion principle which states that no electron
can be in the same quantum state as another one.
To understand what happens to the electrons energy levels when atoms start to interact
it is easier to consider for a moment a system of two coupled pendula (you can think of
each pendulum representing an electron and the frequency of oscillation as representing the
energy levels of the atoms). This can be seen at http://home.a-city.de/walter.fendt/
physengl/cpendula.htm Left to their own devices each pendulum would oscillate with a
frequency
0
. However when the two pendula are coupled together we now need to nd the
modes of the whole system. A moments thought will show you that there are two modes one
where the two pendula are in phase and one where they are out of phase. By playing with
the above simulation you should soon be able to see both these modes. These modes each
have a slightly dierent frequency. Similarly if we have N coupled pendula then there would
be N distinct modes of the whole system each with a dierent frequency.
The situation with atoms and electron energy levels is almost identical. Here the elec-
trons are coupled via the Coloumb interaction and as the atoms are brought closer and closer
together the electron energy levels split split and so if there are N atoms then there are 2N
36
Figure 3.3: Dierence between a metal, semiconductor and insulator.
distinct energy levels per band where the addition factor of two comes from the intrinsic elec-
tron spin. However since in most crystals N is extremely large the energy dierence between
the levels is extremely small and hence we talk about energy bands. This is an alternative
approach to the Kronig-Penney model we looked at in Section 2.6. In both approaches we see
that we would expect to nd permitted energy levels and forbidden energy levels. In terms
of our initial intuition then the picture of atoms and coupled pendula is perhaps easier to
believe but for actually calculations the Kronig-Penney model is preferable.
In a crystal the energy bands can either be full, empty or partially full. We will call the
outermost full energy band the valence band and the next highest energy band the conduction
band. Almost all of the properties of a crystal are due to the electrons in the valence band
or conduction band as electrons in a lower energy level are too tightly bound to the atom to
participate in any interactions. In addition we will refer to the energy dierence between the
valence band and the conduction band as the band-gap.
Electrical conduction in a crystal is only possible when a band is partially full. Clearly an
empty band cannot conduct while, what is also surprising, is that a full band cannot conduct
either. The reason for this is that conduction involves the acceleration of electrons. If the
electron is accelerating then it is gaining energy and in order to do so there must be an empty
energy level for it to move into since due to the Pauli exclusion principle no two electrons can
share the same quantum state. So in order for a material to conduct there must be a partially
full energy band. Given that for N atoms there are 2N energy levels it is clear that any atom
with an odd number of electrons in its outer shell must form crystals with a partially full
outer shell and hence will conduct.
Fig. 3.3 shows a number of possibilities for the dierent arrangements of valence bands
and conduction bands. The rst is where they are widely separated. The third is where
the bottom of the conduction band and the top of the valence band are touching and the
second is where there is is a small separation between the two. Not surprising the rst type
of material form insulators the third type of materials form metals and the second type form
semiconductors. As with most other distinctions the distinction between semiconductors and
insulators is fairly arbitrary as there is a fairly wide continuum as the width of the band-gap
can vary over a wide range. Typical values of restivity range from 10
8
m for a metal to
10
16
m for a good insulator.
3.4 Crystal Structure
In this section I want to talk briey about how atoms join together to form crystals. In
discussing the covalent bond I mentioned that the bonds typically had a very strong direc-
tionality associated with them. This can be easily seen if we consider a single silicon atom.
37
Figure 3.4: A simple 2D crystal structure. The basis is a hexagon which is repeated across
the plane.
Figure 3.5: Formation of a crystal from the space lattice [Fig. (a)] and the basis [Fig. (b)].
Silicon has four electrons in the outer shell and we can imagine that this is similar to having
four negatively charged particles on the surface of a sphere. Now these negatively charged
particles will all want to be as far away from each other as possible. It is easy to convince
yourself that the stable situation will be when the four particles position themselves so as to
be the four vertices of a pyramid (or tetrahedron). Thus we would expect a silicon crystal
to have four nearest neighbours in a tetrahedron arrangement. However visualising three
dimensional crystals is hard to do so I will mainly discuss two dimensional crystals.
In any crystal structure there are two important elements: the lattice and the basis. These
are shown in Fig. 3.4. The lattice is a set of points which are arranged in a regular way and
which covers the whole space. In two dimensions we can completely describe the lattice by
means of two lattice vectors a
1
and a
2
which are shown in Fig. 3. Then the lattice is the
collection of points r such that
r = na
1
+ma
2
, n, m Z. (3.2)
The basis of a crystal is the element that is repeated at each lattice point. In Fig. 3.4 the
basis is a hexagon however it can be anything. The crystal structure is formed by adding the
basis element at every lattice point. This is shown in Fig. 3.5
38
Figure 3.6: Simple crystal lattices in three dimensions. The rst is a simple cubic lattice, the
second is the body centred cubic (bcc) and the last is the face centred cubic (fcc).
3.4.1 Three Dimensional Crystal Types
Extending these denitions to three dimensions is fairly simple. A three dimensional lattice
has three lattice vectors rather than two which is the only dierence. In three dimensions it
can be shown that there are 14 dierent types of lattices however the only one I will discuss in
detail is the diamond structure which is the structure for silicon. Silicon crystals have what
is called a face centred cubic (fcc) lattice in which the basis consists of two silicon atoms.
However before discussing the diamond structure it is helpful to look at some simpler lattices.
In Fig. 3.6 I have shown three simple crystal types. The simplest is the body centered
cubic in which one atom is placed on each of the corners of the unit cell. Each atom is thus
shared by 8 unit cells and thus there is an average of one atom per unit cell. Recall that we
used this fact earlier in deriving the density of states for a particle in a three dimensional
box. The only element that crystalises into a simple cubic lattice is polonium.
The next simplest unit cell is that of the body centred cubic. Here there is an additional
atom placed at the centre of the unit cell. The last lattice structure we are going to look at
is the face centred cubic shown in Fig. 3.6. In comparison to the simple cubic lattice the fcc
has addition atoms placed in the centre of each of the 6 faces of the unit cell
1
.
1
The face centred cubic is one solution to what is known as the sphere packing problem. If you are trying
to pack the most spheres into a region then the solution which will give you the best result is to pack the
spheres into a fcc lattice. This has never actually been proven and it is often said that it is something that
every physicists knows and every mathematician believes.
39
Figure 3.7: Crystal structure of Silicon. The basis consists of two atoms which the lattice
structure is that of face centred cubic.
'
&
$
%
Problem: Find the fraction of of the unit cell that is lled with hard spheres in a
face centred cubic lattice assuming that the spheres are touching.
Solution The rst thing to do is to count the number of atoms per unit cell. In
the picture in Fig. 3.6 there are 14 atoms depicted, 8 at the corners plus 6 on the
faces. Each atom on the corners is shared by 8 other unit cells, while each atom on
the faces is shared by two other cells. Thus the total number of atoms per unit cell is:
8/8 + 6/2 = 4
Now the distance from one atom to its nearest neighbour is a

2/2 and so the radius


of the spheres is half the distance to the nearest neighbour, i.e. r =

2a/4. The
volume of each sphere is
4
3

2
4
a
_
3
=
a
3

2
24
.
The volume of the unit cell is a
3
and we have four spheres so the fraction lled is
4
_
a
3

2
_
/24
a
3
=

2
6
or 74% of the unit cell is lled.
To go from the fcc structure to the diamond structure the only dierence is that of the
basis elements. In Fig. 3.6 the basis was a single atom at each point. In a silicon crystal the
basis is two silicon atoms and this is shown in Fig. 3.7. Another way of thinking about the
crystal structure of silicon is that it consists of two simple fcc lattices interleaved together.
If you, are like me and has trouble visualising what this crystal structure looks like I would
suggest you look at http://jas2.eng.buffalo.edu/applets/index.html which has a very
good tool for looking at a variety of crystal structures. As expected for the covalent bond of
40
Figure 3.8: The band structure of silicon (again).
silicon each atom in the fcc lattice has four nearest neighbours arranged in a tetrahedron.
'
&
$
%
Problem: Calculate the density of Si given that the lattice constant a = 5.43
10
8
cm.
Solution As discussed above for the fcc structure there are 4 atoms per unit cell. In
the diamond structure the basis consists of two atoms and hence there are 8 atoms
per unit cell. This gives us a density of
8
a
3
=
8 atoms
(5.43 10
8
)
3
cm
3
= 5 10
22
atoms/cm
3
the actual density can then be obtained by multiplying the density by the atomic
weight of silicon and dividing by Avogadros number. This gives a density of
2.33g/cm
3
which agrees very well with the experimentally measured value.
The important thing to remember from this calculation is that in a piece of silicon
there are roughly 5 10
22
atoms per cubic centimetre. This is a useful number to
remember for when it comes to calculating number densities of electrons (which we
do in the next section) sensible numbers will be several orders of magnitude less than
this. Also the number of dopant atoms we can incorporate into the silicon lattice is
also signicantly less than this.
By now you are probably wondering what this discussion of crystals has to do with real
devices. In fact it turns out that the crystal structure plays an important part in determining
the properties of the electrons and hence the properties of the semiconductor. In the simplest
case if we imaging making cuts through the crystal in various directions then we would see that
the spacing between the atoms in dierent directions is dierent. Recall that in the Kronig-
Penney model of a crystal the periodicity of the potential lead to band-gaps at particular
energies. Thus if the periodicity is dierent in dierent directions then we would expect that
the band-gap is dierent in dierent directions. This is in fact the case as we can see from
the energy-momentum diagram of silicon as shown in Fig. 3.8.
We can apply what we have learn to discuss this energy-momentum band diagram in
41
more detail. Firstly the dierent points on the x-axis correspond to dierent directions of
propagation through the crystal. At the point the valence band has its highest level. Note
that there are actually two bands overlapping at this point this implies that electrons in the
valence band can have two dierent eective masses
2
. The bottom of the conduction band is
in silicon at the X point where again two eective masses are possible. We can also see that
the width of bandgap is about 1eV.
Current techniques for growing semiconductor devices such as quantum well lasers often
involve growing crystals of dierent materials on top of one another. In order to do this
the crystal properties need to be similar so again you need to know the properties of the
crystals. Similarly when doping the crystals an understanding of the crystal properties helps
to determine which materials make the best dopant.
3.5 The Fermi-Dirac Distribution
We are now ready to begin discussing how free electrons behave in a crystal or a metal. The
central ingredient of this work is what is called the Fermi distribution. The Fermi-Dirac
distribution essentially tells us how many electrons we might nd with a particular energy
which is what we need to calculate quantities like the conductivity of a semiconductor and
how it might vary with temperature. To see why we need the Fermi-Dirac distribution arises
let us consider the problem of calculating the number density of electrons in a material. If n
is the density of electrons per cubic centimetre then we can write
n =
_
E
top
0
n(E)dE (3.3)
where E
top
is the energy level at the top of the conduction band and n(E) is the number of
electrons with an energy E. We then write n(E) = D(E)F(E)E where D(E) is the density
of states function (discussed in section 2.7 ) with energy E and F(E) is the probability that
the state is occupied. F(E) is called the Fermi-Dirac distribution after the physicists Enrico
Fermi and Paul Dirac who rst looked at this problem.
To understand the Fermi-Dirac distribution we need to go back to the Pauli exclusion
principle. As discussed earlier electrons obey what is called the Pauli exclusion principle
which states that no two electrons can be in the same quantum state. Importantly for what
we want to discuss this limits the number of electrons with the same energy to two (each with
opposite spins). Thus if we have N electrons and P energy levels at 0K then the rst N levels
will be lled and the remaining levels will be empty. This is shown in Fig. 3.9. In Fig. 3.9 I
have drawn in the Fermi Level which is indicated by the dotted black vertical line. The Fermi
level is equal to the energy level of the lowest empty level at 0K. At higher temperatures the
Fermi level is the energy level for which F(E) = 1/2.
Earlier we dened the conduction band to be the lowest partially lled band while the
valence band was the highest lled band. If the valence band is lled then clearly F(E) = 1
when E is an energy in the valence band. Thus the Fermi-level for a semiconductor must lie
above the top of the valence band. In most semiconductors at zero degrees the conduction
band is empty as there are no free electrons and so F(E) = 0 for any energy level in the
conduction band at 0K. This implies that the Fermi level must lie between the top of the
valence band and the bottom of the conduction band in a semiconductor at 0K. In fact as
2
This leads to the terms light and heavy holes that you will see in some textbooks
42
Figure 3.9: Graphs of the Fermi-Dirac distribution for dierent temperatures.
we will show later on in almost all circumstances the Fermi level for a semiconductor lies
somewhere in the bandgap. The precise position depends on the levels of doping and the
temperature.
As we raise the temperature things start to become more complicated we are increasing
the total energy of the system while keeping the number of electrons constant. Clearly this
means that some electrons will move to higher energy levels and moreover this can happen
in a variety of ways. The Fermi distribution is the way to describe the various probabilities.
I will not attempt to derive the Fermi-Dirac distribution but rather simplely present it
F(E) =
1
1 +e
(EE
F
)/kT
(3.4)
where F(E) is the probability of nding an electron with energy E when the temperature is T
and the Fermi-level is E
f
. In Eq. (3.4) k is a constant referred to as the Boltzmann constant
which converts temperatures to energies. It value is
k = 1.38065 10
23
JK
1
(3.5)
= 8.617342 10
5
eVK
1
. (3.6)
Note that the units of Boltzmans constant are Joules per degree Kelvin. The Kelvin temper-
ature scale was introduced briey in the introduction and refers to the absolute temperature
scale. Boltzmans constant will be popping up a lot in this course as it is of fundamental
important in thermodynamics and we will be using a lot of the concepts of classical thermo-
dynamics to describe the behaviour of electrons.
In Fig. 3.9 I have plotted the Fermi-Dirac distribution for a number of dierent temper-
atures. From this it can be seen that as the temperature increases the probability increases
that electrons can be found with energies above the Fermi level can be found. Why is this
important?
The answer to this is that in a semiconductor the band-gap has a width of only about 1eV.
This means that although semiconductors are insulators at 0K at room temperature there is
43
a reasonable probability that electrons will be found in the conduction band. And if there are
electrons in the conduction band then the material can conduct. Thus in a semiconductor we
would expect that the conductivity is a strong function of the temperature. This is in marked
contrast to a metal where the conductivity is relatively constant over a wide temperature
range
3
.
From Eq. (3.4) we can make the following approximations
F(E) e
(EE
F
)/kT
for (E E
F
) > 3kT (3.7)
F(E) 1 e
(EE
F
)/kT
for (E E
F
) < 3kT. (3.8)
Eq. (3.7) is known as the Boltzmann distribution and particles such as photons obey the
Boltzmann distribution. The Boltzmann distribution is usually a good approximation for
electrons unless the vast majority of available energy levels are occupied. If the Boltzmann
distribution does not hold then the semiconductor is said to be degenerate. Finally Eq. (3.4)
suggests that in many applications energy should be measured in units of kT. At room
temperature (T = 300K) we nd that kT = 0.026 eV whereas the band-gap for silicon is
1 eV.
3.6 Electrons and Holes
If we can move an electron from the valence band to the conduction band then we are left we
a free energy level where an electron use to be. This then allows electrons in the valence band
to move from one quantum state to another remember that electrons are fermions and no
two fermions can be in the same quantum state. These holes can be thought of as bubbles in
water one way of describing the motion of a bubble is to describe the motion of the water
as it ows into the space left by the bubble. The other is to describe the motion of the bubble
as though it is a real particle. It is this later approach that we will take here with holes. We
will regard them as particles in their own right but with some rather surprising properties.
Firstly have a look at Fig. 3.10 which shows the motion of electrons in an electric eld. From
this we can see that if the electrons move to the right then the hole moves to the left. This
can be explained if we imagine that holes have a positive charge.
In an undoped semiconductor like silicon then the number of free electrons at room tem-
perature is insucient for many applications and it is necessary to increase this somehow.
The standard way is to dope the silicon with a dierent element. Recall that silicon has 4
valence electrons and so wants to form 4 covalent bonds. Now consider what would happen
if we replaced a silicon atom with an Arsenic atom. Arsenic has 5 valence electrons but only
4 of which will form a bond with the surrounding silicon atoms. There is thus one electron
left over which is very loosely bound to the Arsenic atom. In terms of energy levels we say
that the presence of a defect (i.e. the Arsenic atom) introduces an allowed energy level into
the band-gap
4
. Also as the electron is only loosely bound we would expect the energy level
to be close to the top of the band-gap. Fig. 3.11 shows the positions inside the band-gap of
various impurity levels caused by doping silicon with dierent atoms.
3
If you cool some metals down to a few degrees above 0K then they become superconductors and have no
resistance but that takes us too far from o the topic
4
Quantum mechanically the situation is very similar to the case of two potential wells looked at in section
2.5.2. You can think of the dip in the potential barrier in Fig. 2.8 as being a defect. This defect leads to the
presence of a transmission resonance which is exactly what happens with defects in the silicon crystal.
44
Figure 3.10: Schematic of hole transport in a doped semiconductor. The rst picture shows
a simplied 1D view at three dierent time while the second gure shows a 2D lattice at a
particular instance showing the holes caused by impurities.
Figure 3.11: Graph showing position of energy levels within the bandgap of Si for common
dopants. Taken from Szes Semicdonuctor devices.
45
For As atoms in Si the impurity level is 0.054eV below the conduction band. This means
that at 0K the impurity level is full and the conduction band is empty. However at room
temperature the thermal energy is sucient to ensure that all of the impurity electrons are
promoted to the conduction band increasing greatly the conduction of silicon. This is called
n type doping since we are adding negatively charged particles to the crystals.
The reverse case is when we add a group III element such as Boron to silicon. Boron only
has three valence electrons and however it still forms four covalent bonds leaving one of the
neighbouring silicon atoms with a missing electron for its bonds. It can be show that this
forms an additional energy level just above the bottom of the band-gap. At room temperature
this energy level becomes occupied and so the valence band is no longer full and this allows
conduction through the crystal. This is called p type doping since the holes behave as though
they are positively charged. In addition undoped semiconductors are called intrinsic while
doped semiconductors are called extrinsic. If a semiconductor is n-type then electrons are
refered to as the majority carriers while holes are called the minority carriers and via-versa
for a p-type semiconductors.
'
&
$
%
Useful Things to Remember:
Electrons follow the Fermi-Dirac distribution.
The absence of an electron is equivalent to a positively charged particle called
a hole.
Semiconductors can be doped to increase the concentration of either electrons
(n-type semiconductors) or holes (p-type semiconductors)
An undoped semiconductor is called an intrinsic semiconductor.
A doped semiconductor is called a extrinsic semiconductor.
In doped semiconductors you have majority and minority carriers. In n-type
samples electrons are the majority carriers while in p-type samples holes are
the majority carriers.
3.7 Carrier concentration
In this section we are going to look at the consequences of the Fermi-Dirac statistics. The
rst thing we are going to look at is the distribution of holes. Clearly as a hole is the absence
of an electron we must have
P
n
(E) +P
p
(E) = 1 (3.9)
where P
n
(E) is the probability of nding an electron with an energy E and P
p
(E) is the
probability of nding a hole. Substituting in Eq. 3.4 for P
n
(E) we nd that
P
p
(E) =
1
1 +e
(E
F
E)/kT
. (3.10)
This shows us that, as expected, holes are more likely at higher energies and are less likely
the deeper into the valence band you go. Note that for reasonable temperatures and energies
46
the approximations given by Eq. (3.7) hold for holes with the appropriate changes.
The next piece of information we wish to determine is how many free electrons there
are in a piece of semiconductor. The number of free electrons (or holes) will determine the
conductivity of the sample and so it is a useful piece of information. To determine the number
of free electrons we can count the number of electrons with a particular energy then count the
number with a slightly higher energy add this to the previous total and then ... . Formally
we can write the number density n of electrons per unit volume as
n =
_
E
top
E
c
(E)dE (3.11)
where E
c
is the energy at the bottom of the conduction band and E
top
is the energy at the
top of the conduction band. The number of electrons at an energy E is given by (E). We
can express (E) as being the product of the number of possible states at an energy E times
the probability that an electron has energy E, i.e:
(E) = D(E) F(E) (3.12)
where D(E) is the density of states dened in Eq. (2.102). The meaning of this integral is
shown schematically in Fig. 3.12. Putting the expressions for D(E) and F(E) into Eq. (3.11)
we get
n =
_
E
top
E
c
1
2
2
_
2m
e

2
_3
2 _
E E
0
1
1 +e
(EE
f
)/kT
dE (3.13)
where m
e
is the eective mass of the electron at the bottom of the conduction band. Now to
simplify matters we make two approximations. Firstly we set E
top
= which is justied as
F(E) 0 as E so this will not change the value of the integral by much. The second
approximation is to approximate F(E) by the Boltzmann distribution given in Eq. (3.7) which
becomes increasingly valid the greater the energy. This then gives
n =
1
2
2
_
2m
e

2
_3
2
_

E
c
_
E E
c
e
(EE
f
)/kT
dE (3.14)
setting x = (E E
c
)/kT Eq. (3.14) becomes
n =
1
2
2
_
2m
e

2
_3
2
(kT)
3
2
e
(E
c
E
f
)/kT
_

0

xe
x
dx (3.15)
Now the integral in Eq. (3.15) might not be one that you have seen before but if you look it
up in a table of integrals
5
you will nd that the answer is

/2. This then gives us our nal
answer for the number density of free electrons in a crystal.
n =
1
2
2
_
2m
e

2
_3
2
(kT)
3
2

2
e
(E
c
E
f
)/kT
(3.16)
Eq. (3.16) is however often written in a two slightly dierent but equivalent forms:
n = N
c
e
(E
c
E
f
)/kT
(3.17)
5
For example see Table of integrals, series and products by I. S. Gradshteyn and I. M. Ryzik.
47
Figure 3.12: (a) Schematic band diagram for a n-type semiconductor. (b) Density of states.
(c) Fermi distribution function. (d) Carrier concentration.
or
n = n
i
e
(E
f
E
i
)/kT
(3.18)
where N
c
is the eective density of states and the value of this can be worked out quite
simplely by comparing Eq. (3.16) and Eq. (3.17). In the second expression n
i
is the intrinsic
density of states and E
i
is the intrinsic Fermi level. For silicon at room temperature N
c
=
2.86 10
19
cm
3
and for GaAs N
c
= 4.7 10
17
/cm
3
. We shall come back to the meaning of
these terms shortly but rst lets derive similar expressions for the number density of holes.
3.7.1 Number density of holes
In this section we can go a lot more quickly since as we will see most of the maths is identical
to math we did previously for the case of electrons. The starting point is again that
p =
_
E
v
0

p
(E)dE (3.19)
where p is the number density of holes, E
v
is the energy at the top of the valence band and

p
(E) is the number of holes with energy E. We can now proceed exactly as before writing

p
(E) = D(E)P
p
(E) (3.20)
where D(E) is the density of states for holes which is identical to that of electrons in the
valence band and P
p
(E) is the probability of nding an hole with an energy E and is given
by Eq. (3.10). Substituting these expressions into Eq. (3.19) we get
p =
_
E
v
0
1
2
2
_
2m
h

2
_3
2 _
E
c
E
1
1 +e
(E
f
E)/kT
dE (3.21)

1
2
2
_
2m
h

2
_3
2
_
E
c

_
E
c
E e
(EE
f
)/kT
dE (3.22)
=
1
2
2
_
2m
h

2
_3
2
(kT)
3
2

2
e
(E
f
E
v
)/kT
(3.23)
48
As in the case of the hole density we can write
p = N
v
e
(E
f
E
v
)/kT
(3.24)
or
p = n
i
e
(E
i
E
f
)/kT
(3.25)
where N
v
is the eective density of states at the valence band. To give you a feel for what the
typically values are, N
v
= 2.66 10
19
cm
3
for silicon at room temperature while for GaAs
N
v
= 7.0 10
18
cm
3
.
3.7.2 Why we care
Now that we have done all the hard work we can start to see how it will pay o. An important
thing to realise about the equations for p and n that we have derived is that we have not
made assumption about whether or not the material was doped and hence they will hold both
for intrinsic and extrinsic semiconductors. In the rst instance let us look at the case of an
intrinsic semiconductor. Here we must have n = p since every free electron will correspond
to a hole in the valence band. Equating Eq. (3.21) and Eq. (3.16) we get
1
2
2
_
2m
h

2
_3
2
(kT)
3
2

2
e
(E
f
E
v
)/kT
=
1
2
2
_
2m
e

2
_3
2
(kT)
3
2

2
e
(E
c
E
f
)/kT
(3.26)
or upon rearranging:
_
m
e
m
h
_3
2
= e
(E
f
E
v
)/kT
e
(E
c
E
f
)/kT
(3.27)
and then taking the log of both sides and rearranging slightly
E
v
+E
c
2E
f
= kT
3
2
ln
_
m
e
m
h
_
(3.28)
or
E
f
=
E
c
+E
v
2

3kT
4
ln
_
m
e
m
h
_
(3.29)
Eq. (3.29) tells us that the Fermi level in an intrinsic semiconductor would be in the middle
of the gap except for the dierence in eective masses of holes and electrons. So now we know
how to calculate the Fermi level for an intrinsic semiconductor. Eq. (3.29) thus gives us an
explicit expression for E
i
which is
E
i
=
E
c
+E
v
2

3kT
4
ln
_
m
e
m
h
_
(3.30)
49
The next question is how we would calculate it for an extrinsic semiconductor. To do this
let us assume that for the moment our crystal is n type and is doped with a density N
D
of
donor atoms per cubic metre. We saw in the previous section that the donor atoms create
an extra energy level just below the conduction band so that at room temperature we can
assume that all of these atoms have been ionised and hence
n = N
d
(3.31)
we now stick the value for n into Eq. (3.17) and solve for E
f
. The answer you should get is
E
f
= E
c
kT ln
_
N
c
N
d
_
(3.32)
so we can see that the greater the concentration of donor atoms the close the Fermi level moves
towards the conduction band. Alternatively we can substitute Eq. (3.31) into Eq. (3.18) to
get
E
f
= E
i
+kT ln
_
N
d
n
i
_
(3.33)
For the case of p type materials we again assume that we have N
A
acceptors and that at
room temperature all of them have caught an electron leaving N
A
holes. Sticking this value
for p into Eq. (3.25) we nd that
E
f
= E
v
+kT ln
_
N
v
N
A
_
(3.34)
and hence as I hope you would have guessed the presence of acceptors moves the Fermi level
down towards the valence band. Again we can describe the position of the Fermi-level relative
to the intrinsic level using
E
f
= E
i
kT ln
_
N
A
n
i
_
(3.35)
Another useful relation comes from looking at the product of n and p. Starting from
50
Eq. (3.16) and Eq. (3.23) gives us:
np =
_
1
2
2
_
2m
e

2
_3
2
(kT)
3
2

2
e
(E
c
E
f
)/kT
__
1
2
2
_
2m
h

2
_3
2
(kT)
3
2

2
e
(E
f
E
v
)/kT
_
(3.36)
= N
c
N
v
e
(E
c
E
f
)/kT
e
(E
f
E
v
)/kT
(3.37)
= N
c
N
v
e
(E
c
E
f
+E
f
E
v
)/kT
(3.38)
= N
c
N
v
e
(E
c
E
v
)/kT
(3.39)
= N
c
N
v
e
(E
c
E
v
)/kT
(3.40)
= N
c
N
v
e
E
g
/kT
(3.41)
(3.42)
where E
g
is the size of the bandgap. If we then dene
n
i
=
_
N
c
N
v
e
E
g
/(2kT)
(3.43)
we nd that
np = n
2
i
. (3.44)
which is known as the law of mass action. The nal form is useful as it shows that the product
of electron and hole density does not depend on the position of the Fermi level at all, rather
it is simplely a function of the material parameters and the temperature. Nore that the law
of mass action holds for both intrinsic and extrinsic semiconductors. The law of mass action
tells us that for an extrinsic semiconductor if we increase the number of holes via doping then
the number of free electrons decreases. Similarly the reverse happens when we add donor
atoms to silicon.
'
&
$
%
Problem: We now want to calculate a value for n
i
for silicon at room temperature
(i.e. T=300K). You can use the values m
e
= 1.18 and m
h
= 0.59 and the width of
the bandgap is 1.1eV. Assuming that you do the calculation correctly the answer
you should get is about 8 10
9
cm
3
.
It is also interesting to consider for a moment what happens in a doped semiconductor as
the temperature is increased. Imagine that we have a n-type piece of silicon then
n = n
i
+N
D
(3.45)
where N
D
is the donor concentration and n
i
is the intrinsic electron distribution. Now N
D
is a
very weak function of temperature being basically constant and equal to the number of donor
atoms while n
i
increases exponentially with temperature. Thus we would expect that at high
temperatures n
i
N
D
while at low temperatures N
D
n
i
. This is shown in Fig. (3.13)
51
Figure 3.13: Carrier concentration as a function of temperature. Taken from Szes Semicon-
ductor devices.
where three distinct regimes can be seen. At extremely low temperatures the donor atoms
are unionised and hence they concentration approached zero. However the temperature is
suciently high all of the donor atoms have been ionised and thus the carrier concentration
is roughly constant. Then when the thermal energy is sucient to excite carriers across the
bandgap the carrier concentration starts to increase exponentionally.
The last point that I need to make in this section is that the Fermi level at equilibrium
does not change with position. The reason for this is quite simple if the Fermi level
did change with position then currents will ow as electrons and holes moved to the region
with the lowest Fermi-level. And if currents were owing then the sample would not be
in equilibrium. The other related point is that in a non equilibrium situation then strictly
speaking there is no such thing as a Fermi level however in many cases we will just pretend
that there is one.
3.7.3 Calculation
The following example is taken from Greg Parkers book Introductory Semiconductor Device
Physics which is unfortantly out of print else I would strongly recommend you get it.
A piece of silicon is doped with 5 10
16
Boron atoms per cm
3
. Calculate the electron
concentration at 300K. What is the position of the Fermi-level with respect to the valence
band edge and the intrinsic Fermi-level?
The key to answering this question is to realise that at room temperate all of the Boron
atoms are ionized and so the hole density p = 510
16
cm
3
. The using the law of mass action
we nd that
n = n
2
i
/p
52
setting n
i
= 10
10
cm
3
gives us
= 2000 cm
3
which is the answer to the rst part of the question. To solve the next part of a question
we have a large number of equations which we can use. To nd the Fermi-level I would use
Eq. (3.34) where the only unknown is N
v
which we can take equal to 10
19
cm
3
. Putting in
the numbers then gives that
E
f
E
v
= 0.137 eV.
Similarly to nd the position of the Fermi-level with respect to the intrinsic level we need
to use Eq. (3.35) this gives us
E
i
E
f
= 0.398 eV.
Adding these two results together we nd that the intrinsic level is 0.535 eV above the
valence band. Comparing this with the middle of silicons energy gap which is 0.55 eV above
the valence band we see that the dierent eective masses have shifted the level slight as
expected.
3.7.4 Summary
In the last few sections we have started from the Fermi-Dirac distribution for electrons and
combined this with the density of states (which we did in the last module) to obtain expressions
for the density of free electrons and holes in a semiconductor. We then showed that the product
of the electron and hole densities was a constant which did not depend on the doping. We
also showed how the position of the Fermi-level moved in response to the eect of doping.
This section is one of the more mathematical sections in the course however it is one of
the most essential and you should ensure that you understand all of the main features.
The minimum you should take away from this section is rstly the law of mass action
i.e. np = n
2
i
. Secondly you should remember Eq. (3.17) and Eq. (3.18) for the free electron
densities along with the corresponding equation for the hole densities. Of course you will need
to be able to apply these equations to solve problems such as the example given above.
3.8 The motion of an electron in a electric eld
We have now covered all that we need to about the carrier distributions and we now turn to
examining how an electron behaves in an electric eld. To understand this we rst need a
couple of concepts from thermodynamics. The rst is that we can think of free electrons in
a crystal behaving like an ideal gas. One of the properties of an ideal gas is that when it is
in thermal equilibrium at a temperature T the average kinetic energy of a particle moving in
three dimensions is 3/2kT where our good friend kT appears once more. Equating this to
the kinetic energy 1/2mv
2
we nd that the average velocity v
th
of an electron is
v
th
=
_
3kT
m

(3.46)
where m

is the eective mass of the electron. Importantly this is a random motion and
electrons are equally likely to be moving in any direction if this was not the case then the
crystal would not be in equilibrium since there would be a macroscopic charge separation after
53
some time. Typically v
th
10
5
m/s for both silicon and GaAs at 300K. Recall that earlier we
stated that in a doped semiconductor at room temperature all of the dopants were ionised.
We can now see why that is: since in a doped semiconductor the dierence between the donor
level and the edge of a bandgap is less than 3/2kT at room temperature the thermal energy
given by 3/2kT is sucient to ionise the atoms.
However the thermal velocity only tells us half the story, the other half of the story is that
free electrons in a crystal are constantly colliding with all sorts of things other electrons,
impurities in the crystal, thermal vibrations of the crystal
6
and so on. We can however dene
a time which is the average time between collisions. There is also the average distance an
electron moves between collisions which is given by v
th
and is called the mean free path.
Now consider what happens when we apply an electric eld to the semiconductor. An
electron will feel a force given by
F = qE (3.47)
or
m

dv
dt
= qE (3.48)
where Eq. (3.48) follows from Newtons third law. For an electron starting from rest we can
make the following approximation
dv
dt

v 0
t
=
v
d

(3.49)
where we have set t = which is the average time that an electron will be accelerated
before undergoing a collision and having its velocity reduced to zero. Substituting Eq. (3.49)
into Eq. (3.48) we nd for the drift velocity
v
d
=
q
m

E (3.50)
or
v
d
= E. (3.51)
where

q
m

(3.52)
In Eq. (3.51) I have introduced the symbol which is the called the electrons mobility and
has units of cm
2
V
1
s
1
. It tells you how fast an electron will move in response to an electric
eld. If you are looking to build high speed switches then you will want to use a material
which has a high mobility. The other thing to notice which is essential to understanding elec-
tronics is that in bulk matter the average velocity of an electron is proportional to the electric
eld, whereas in free space the acceleration of an electron is proportional to the electric eld.
6
Thermal vibrations are also called phonons and can be thought of as sound waves
54
'
&
$
%
Problem: Calculate the mean free time of an electron having a mobility of
1000 cm
2
/(V s) at 300K. Also calculate the mean free path. Assume that
m

= 0.26m
0
in these calculations.
Solution: From Eq. (3.52) the mean free time is given by:
=
m

q
=
_
0.26 0.91 10
30
kg
_

_
1000 10
4
m
2
/(Vs)
_
1.6 10
19
C
= 1.48 10
13
s = 0.148 ps
The mean free path is given by
l = v
th
= (10
5
m/s)(1.48 10
13
s) = 1.48 10
8
m
By now you should also have realised that anything we do with electrons we can do with
holes and the same calculation applies to holes and one can dene a hole mobility in an
exactly analogous way to Eq. (3.50). The only dierence will be that holes will move in the
opposite direction to electrons as they have a positive charge. One thing to note is that as
the eective mass of a hole is greater than that of an electron the mobility will be smaller.
Thus for high speed devices you will want to use n doped materials where ever possible.
We now want to go from the microscopic motion of the electrons (or holes) to macroscopic
quantities such as the current I or more usually the current density J. Now the current density
due to electrons moving at a velocity v is just qnv where n is the number density of the
electrons and q is the charge on an electron. Then from the drift velocities v
n
and v
p
for
the electrons and holes (remember that holes and electrons move in opposite directions in the
presence of an electric eld) we can calculate the current density J as
J = (q)n(v
n
) +qpv
p
(3.53)
= q(nv
n
+pv
p
) (3.54)
= q(n
n
+p
p
)E (3.55)
:= E (3.56)
where the conductivity is dened as
q(n
n
+p
p
). (3.57)
In an extrinsic semiconductor the amount of doping will ensure that either n p (n-type
doping) or p n (p-type doping). In this case Eq. (3.53) can be simplied to only include
the majority carrier, eg. J qnv
n
for n-type materials.
We can now put Eq. (3.56) into a more familiar form using Fig. 3.14. Firstly we can
express the electric eld as the ratio of the potential dierent across the crystal over its
55
Figure 3.14: Figure for use in deriving Ohms law.
length, i.e. E = V/L, secondly the current density J = I/A where I is the current and A is
the cross sectional area. Sticking these expressions into Eq. (3.56) we get
I
A
=
V
L
(3.58)
rearranging to make V the subject gives
V =
L
A
I. (3.59)
Eq. (3.59) is just Ohms law which should be familiar to all of you. Thus the resistance of a
crystal is given by
R =
L
A
(3.60)
or
R =
L
A
(3.61)
where the symbol is called the resistivity and is the inverse of the conductivity i.e.

1

=
1
q(n
n
+p
p
)
(3.62)
If we are dealing with a doped semiconductor then usually n p or p n and hence we can
ignore the eects of the minority carriers in the above equation.
3.8.1 Summary
In this section of the course we have started with the idea that the free electrons/holes in a
semiconductor move in a random fashion characterised by a time between collisions. We
then looked at what happened in the presence of an electric eld and derived Ohms law. We
also introduced the concept of electron/hole mobility which describes how an electron or hole
responses to an electric eld.
56
Figure 3.15: Diagram showing the setup for the Hall eect.
3.9 The Hall Eect
In this section I want to briey describe an experiment that can be used to determine where
a piece of semiconductor is n doped or p doped. Those of you who are not yet convinced
about the reality of holes should hopefully be converted by the end of this section.
The Hall eect arises when a piece of semiconductor is placed in a magnetic eld as shown
in Fig. 3.15. Now in the presence of a magnetic an electric eld a charged particle feels a
force given by the Lorenz law
7
F = q(E+v B) (3.63)
where q is the charge on the particle which can be either positive or negative. Using the
coordinate axis shown in Fig. 3.15 and the fact that the magnetic eld is constant and vertical
(i.e. B = Bk) we nd that a charged particle feels a force in the y direction given by
F
y
= qv
x
B
z
(3.64)
where v
x
is the x component of the particles motion. Now the the motion of charges in one
direction will lead to the buildup of an electric eld across the semiconductor since the ions
cannot move. In equilibrium the force due to the induced electric eld will cancel out the
force due to the magnetic eld i.e.
F
y
= 0 = qE
y
+qv
x
B
z
(3.65)
or
E
y
= v
x
B
z
(3.66)
but we know v
x
is proportional to the current density J via Eq. (3.53). If we now assume
that the semiconductor is p type we can write v
x
= J/(qp) and this gives us
E
y
=
JB
z
qp
. (3.67)
7
Hopefully you will remember the Lorenz law from last semesters course on physics. If you dont it is not
critical as we do not use it again after this section.
57
We can also write the current density as the current I divided by the area A. Also the electric
eld E
y
will result in a potential dierence V
h
= E
y
w which can be measured. Thus we can
rewrite Eq. (3.67) in terms of measurable quantities as
p =
IB
z
w
qV
h
A
(3.68)
which gives us a simple way of measuring the carrier density in a semiconductor. If we had
instead a n type semiconductor then the inducted hall voltage V
h
would have the opposite
sign and so this experiment allows us to determine where holes and electrons are the majority
carrier in a unknown semiconductor. In that case the analogue of Eq. (3.68) becomes
n =
IB
z
w
qV
h
A
(3.69)
'
&
$
%
Problem (From Streetman): A sample of Si is doped with 10
17
phosphorus
atoms/cm
3
. What would you expect to measure for its resistivity? What
Hall voltage would you expect in a sample 100 m thick if I
x
= 1 mA and
B
z
= 10
5
Wb/cm
2
? You can take the electron mobility to be 700 cm
2
/(Vs).
Solution: As we are dealing with a highly doped extrinsic semiconductor we can
ignore the eects of the minority carriers in the relevant equations. Thus Eq. (3.57)
gives us:
= q
n
n =
_
1.6 10
19
C
_ _
700 cm
2
/(Vs)
_ _
10
17
_
= 11.2
1
cm
1
To determine the Hall voltage we need to use Eq. (3.69) with the value of w/A =
1/(100 m). gives us
V
h
=
_
10
3
A
_ _
10
5
Wb/cm
2
_
(1.6 10
19
C) (10
17
) (10
2
cm)
= 62.5 V
3.10 Carrier diusion
So far we have looked at how electrons move in response to an electric eld and also how they
behave in thermal equilibrium (i.e. random motion). The aim of this section is to look at
how they respond to a gradient in the carrier density. Such a gradient much be created by a
current source or by a shining a light onto the semiconductor
8
.
Going back to the model of electrons as being an ideal gas then we all know what happens
to an ideal gas if there is a pressure gradient the gas diuses till the pressure gradient
8
we will consider the interaction between light and semiconductors in a later part of the course
58
Figure 3.16: Schematic veiw of a diusion process.
disappears. Similarly if we imagine that there is a large concentration of electrons in some
region then the coloumb force will ensure that these electrons repel each other till the local
concentration equals the average concentration. This process is known as diusion and in
order to gain a better understanding of this process consider the situation shown in Fig. 3.16.
In Fig. 3.16 l is the mean free path of the electrons (we will deal with electrons explicitly
and you can work out the details for holes yourselves). We then dene the average density of
electrons in region 1 as n
1
and the average density of electrons of electrons in region 2 as n
2
.
We now want to work out the number of electrons in region 1 that will cross a plane located
at x = x
0
into region 2. The total number of electrons in region 1 is
N
1
= n
1
Al (3.70)
where A is the cross sectional area. Now in a time each electron will have moved a distance
l thus if all the electrons were moving towards the right then after a time all N
1
electrons
would have moved into region 2. However the electrons are moving randomly so on average
half the electrons will be moving towards region two and the other half away from it. Thus
the number crossing into region 2 will be 1/2N
1
. By the same argument 1/2n
2
Al electrons
will move from region 2 into region 1 in a time . Thus the net ux of electrons F to the
right is
F =
1
2
(n
1
Al n
2
Al) (3.71)
F =
Al
2
(n
1
n
2
). (3.72)
Now if the distance l is very short then we can make the following approximations
9
n
1
n
0
l
dn
dx
(3.73a)
n
2
n
0
+l
dn
dx
. (3.73b)
Substituting Eqs. (3.73) into Eq. (3.72) we get
F = Al
2
dn
dx
(3.74)
9
Technically this is called a Taylor series expansion.
59
and in order to obtain the value of the ux per unit time f we just have to divide Eq. (3.74)
by the time this gives
f =
Al
2

dn
dx
(3.75)
but l/ is just the thermal velocity v
t
h
f = Av
th
l
dn
dx
(3.76)
or
f
A
= v
th
l
dn
dx
(3.77)
We can now go from f/A to the current density J
n
simplely by multiplying by the charge q
for electrons. This gives us
J
n
= qD
n
dn
dx
(3.78)
where we have dened the diusion coecient D
n
as
D
n
v
th
l (3.79)
We will now derive a relationship between the mobility and the diusion coecient. To
do this we need the additional fact that for an ideal gas each particle has on average 1/2kT
of thermal energy per degree of freedom. As we are considering one dimensional motion at
present we can write the average kinetic energy as:
1
2
m

v
2
th
=
1
2
kT. (3.80)
We can now proceed as follows:
D
n
= v
th
l (3.81)
= v
th
(v
th
) (3.82)
And from Eq. (3.52) we can express in terms of the mobility giving
= v
2
th
_

n
m

q
_
(3.83)
Then using Eq. (3.80) we obtain
=
_
kT
m

__

n
m

q
_
(3.84)
60
This then gives the Einstein relation between the diusion coecient and the mobility
D
n
=
kT
q

n
(3.85)
You should be able to repeat this derivation for the hole diusion current yourself. The
analysis is the same until you have to convert from ux to current density. Only this time
you multiple by +q rather than by q so you nd that the diusion current for holes has the
opposite sign than for electrons. This is expected since both holes and electrons move away
from regions of higher concentration but a movement of electrons results in a current owing
in the opposite direction whereas for holes the current is in the same direction as the holes
motion.
In the cases where there is both an electric eld and a concentration gradient present we
can add the drift and diusion currents to obtain
J
n
= q
n
nE +qD
n
dn
dx
(3.86)
for electrons and
J
p
= q
p
pE qD
p
dp
dx
(3.87)
for holes. One thing to note about these equation is that in many cases they predict that
the diusion of minority carriers in a material might be the dominant source of current. The
reason for this is that diusion currents are driven by gradients and even a small density of
minority carriers might have a large gradient. Drift currents however are dominated by the
majority carriers in most cases.
3.10.1 Charge Continuity Equations
In this nal section we will derive the equations for the time evolution of carrier density
uctuations. Essentially the result is identical to that of Eq. (3.86) and Eq. (3.87) since the
current density is proportional to the velocity of the carriers and hence Eq. (3.86) for example
implicitly describes how the position of carriers changes with time. However we would like
something more explicit.
The rst thing we need to consider is recombination. This is the process by which an free
electron in the conduction band loses energy and drops down to the valence band. Due to the
Pauli exclusion principle this can only happen if there is an empty energy level in to valence
band i.e. a hole. Thus recombination is a process by which one free electron and one hole
recombine. The energy lost by the electron can be converted into a number of forms such as
light, heat or sound. In a later section of the course we will examine light emitting diodes in
some detail and this is one example of carrier recombination.
To treat recombination we need to dene a characteristic lifetime which is the average
time which a hole or an electron spends wandering about before it recombines. This is
61
Figure 3.17: Picture for the derivation of the charge continituity equations.
exactly analogous to the half life of a radioactive element. If we then have an excess carrier
distribution C then we would expect it to decay according to
dC
dt
=
C

(3.88)
where the carrier density C equals the equilibrium carrier density C
0
plus an excess distribu-
tion C.
If we now consider all the processes involved we can write the following equation
Rate of change of the number of carriers in volume V =
rate at which carriers arrive rate at which carriers leave rate of recombination (3.89)
We can express the above more mathametically using Fig. 3.17. We will also assume for
the purposes of this discussion that the current is being carried by holes. Now in Fig. 3.17 the
volume of the region we are considering is Ax and so the number of carriers in that volume
is Axp where p is the hole density. We can thus write the left hand side of Eq. (3.89) as
(Axp)
t
= Ax
p
t
(3.90)
Now we need to work out the right-hand side of Eq. (3.89). The rate of increase in carriers
is given by I
p
(x)/q where I
p
is the current. Similarly the rate at which the carriers leave is
given by I
p
(x + x)/q. From Eq. (3.88) we know that the rate of decay is given by p/
p
where p is the excess carrier density, i.e. p = p
0
+p. Putting this all together we get
Ax
p
t
=
1
q
(I
p
(x) I
p
(x + x))
p

p
(3.91)
dividing both sides by the volume we get
p
t
=
1
q
_
I
p
(x)/AI
p
(x + x)/A
x
_

p
(3.92)
but the current divided by the area is just the current density so
(p
0
+p)
t
=
1
q
_
J
p
(x) J
p
(x + x)
x
_

p
(3.93)
62
Now p
0
is assumed constant in time as it is the equilibrium distribution hence we can simplify
the l.h.s. of Eq. (3.93) giving
(p)
t
=
1
q
_
J
p
(x) J
p
(x + x)
x
_

p
(3.94)
Upon taking the limit as x 0 we realise that the rst term on the r.h.s. is just the spatial
derivate of J
p
, i.e.
(p)
t
=
1
q
J
p
x

p

p
(3.95)
Similarly one can repeat the process for electrons (and I hope you do) to obtain
(n)
t
=
1
q
J
n
x

n

n
(3.96)
where the only dierence is the sign change before the derivate of J as electrons ow in the
opposite direction to the current.
In the case where there is there little drift and the current is carried strictly by diusion
we can write
J
p
(x) = qD
p
(p)
x
(3.97)
and substituting this result into Eq. (3.95) we get
p
t
= D
p

2
p
x
2

p

p
(3.98)
and again similarly for electrons we get
n
t
= D
n

2
n
x
2

n

n
. (3.99)
These are the diusion equations for electrons and holes in a semiconductor. They will be
needed when we look at carrier injection into a device. However for the moment we will only
look at a simple steady state solution. This will however introduce the concept of the diusion
length.
Suppose now that our system is in a steady state i.e. all the time derivates are identically
zero and that we are injecting a steady stream of carriers into the material (exactly how we
do this is unimportant). We can then write Eq. (3.98) as
0 = D
p

2
p
x
2

p

p
(3.100)
or upon rearranging

2
p
x
2
=
p
D
p

p
(3.101)
63
after dening the hole diusion length L
p
=
_
D
p

p
we can write Eq. (3.101) as

2
p
x
2
=
p
L
2
p
. (3.102)
Again we can of course dene an electron diusion length L
n
=

D
n

n
which allows us, in
the steady state regime to rewrite Eq. (3.99)

2
n
x
2
=
n
L
2
n
. (3.103)
The diusion length L
n,p
can be thought of as the average length that a carrier can propagate
before recombining.
Now hopefully you should immediately be able to solve Eq. (3.102) or Eq. (3.103) since
it is the same equation as the free particle version of Schrodingers equation. The general
answer (for holes) is
p(x) = A
1
e
x/L
p
+A
2
e
x/L
p
(3.104)
where A
1
and A
2
can be found from the boundary conditions of the problem at hand. For
now all we need to know that if we inject carriers into a semiconductor at some point then
they will decay exponentionally away from that point as we move into the semiconductor.
3.10.2 Summary
In the last two sections we have looked at the diusion of carriers due to gradients in the
concentration. This lead us to the concept of diusion currents which as we will see are fun-
damental to the operation of diodes. We also looked at the transient response to concentration
gradients and saw how the eect of recombination lead to a exponentional decay of the carrier
density in space.
64
3.11 Problems
1. Discuss the meaning of the Fermi-Dirac distribution, use diagrams if necessary.
2. Starting with the Fermi distribution for electrons and the fact that
P
n
(E) +P
p
(E) = 1
where P
n
and P
p
are the probabilities for nding electrons and holes respective, show
that
P
p
(E) =
1
1 +e
(E
f
E)/kT
.
Also nd simplied expressions for P
p
(E) similar to those given for electrons in Eq. (3.7).
3. Prove Eq. (3.32) i.e.
E
f
= E
c
kT ln
_
N
c
N
d
_
4. Calculate the hole and electron densities in a piece of p-type silicon that has been doped
with 5 10
16
acceptor atoms per cm
3
?
5. What fraction of silicon atoms are ionised at room temperature? How does this change
if the temperature is increased to 600K? In GaAs the bandgap is smaller than in silicon,
how do you expect this to alter the number of atom ionised?
6. Find the electron and hole conentrations and Fermi level in silicon at 300K (a) for
10
15
boron atoms/cm
3
and (b) 3 10
16
boron atoms/cm
3
and (c) 2.9 10
16
arsenic
atoms/cm
3
.
7. A semiconductor is doped with N
D
(N
D
n
i
) and has a resistence R
1
. The semicon-
ductor is then doped with an unknown amount of acceptors N
A
(N
A
n
i
) yielding a
resistence of 0.5R
1
. Find N
A
in terms of N
D
if D
n
/D
p
= 50.
8. Consider a semiconductor that is nonuniformly doped with donor impurity atoms N
D
(x).
Show that the induced electric eld in the semiconductor in thermal equilibrium is given
by
E(x) =
_
kT
q
_
1
N
D
(x)
dN
D
(x)
dx
9. Consider a electron with an energy of energy 2eV impinging on a potential barrier of
height 20eV and a width of 30nm. What is the tunnelling probability?
10. Which elements can be added to silicon to create n and p type semiconductors?
65
Module 4
Junctions
Learning Objectives In this module we will learn about the properties of semiconductor
junctions, i.e diodes and in particular by the end of this section you should:
1. Know what a diode is, and give examples of where it is used.
2. Be able to explain how a diode works
3. Be familiar with the concepts of drift and diusion currents, contact potential, reverse
bias, breakdown potential and give brief descriptions of them.
4. Understand the derivation of the diode equation and be able to use it to solve problems.
5. Explain what happens to a diode under reverse breakdown and how this can be used in
realistic circuits.
Web resources To help with this section of the course I would strongly recommend you
look at http://jas2.eng.buffalo.edu/applets/index.html which has an excellent set of
applets showing the behaviour of pn junctions. This is one of the best sites I have found for
semiconductor physics.
If you also do a quick search on google you can nd a wide selection of notes on semi-
conductor physics. I would suggest you read these for an alternative viewpoint to my own as
dierent people explain things in dierent ways. However I would urge some caution as not
everything you will nd will be accurate.
Directed Reading This module mostly covers Chapter 5 of the core text (Solid State
Electronic Devices) and I would strongly suggest that you read this chapter. In particular I
will not cover the work on the fabrication of solid state devices as I am not active in that
area and hence any information I give you would only be a summary of what is in the text.
4.1 Introduction
In the last module we looked at the basic properties of semiconductor crystals both in equi-
librium and when there was an applied electric eld or concentration gradient. Hopefully you
should now be familiar with the idea of a Fermi level and how the position of the Fermi level
determines the number of free carriers whether they are holes or electrons. In addition you
understand the two types of currents that can exist drift currents and diusion currents
and know when each are important.
66
Figure 4.1: Schematic of a p-n junction before joining the two pieces of semiconductors
In this module we are going to use these ideas to look at the properties of junctions,
whether they are between two dierently doped semiconductors (i.e. diodes) or between a
semiconductor and a metal. Diodes or pn junctions are fundamental to the design of all
semiconductor devices so we will concentrate on these to begin with. However the junction
between a semiconductor and a metal is also important as once we have built our semi-
conductor devices we need to be able to connect it to the outside world if we want to use
it.
4.2 Junctions at Equilibrium
The basic junction we will be considering is the p-n junction
1
as shown in Fig. 4.1. In Fig. 4.1
I have shown schematically two separated pieces of semiconductor, one p doped and the other
n doped. I have also tried to indicate where in each material the Fermi level is located within
the bandgap. Note that I have implicitly assumed that each crystal is made of the same
material and hence the size of the bandgap in both pieces is identical.
The question we now want to answer is what happens when we bring the materials to-
gether
2
? First of all when everything has settled down and the system is in thermal equilib-
rium then the Fermi level will be constant across both materials. As the material properties
of each sample can not change by being in contact with another material this means that the
whole band gap must shift rigidly up or down so that the Fermi levels match up. This is
shown in Fig. 4.2 where I have only shown the positions of the energy levels.
From Fig. 4.2 it can be seen that there is a dierence between the energy level of the
conduction band in the p-type material E
cp
and the dierence in the energy level of the
conduction band in the n-type material E
pn
. If this dierence is E
pn
then
E
pn
= E
cp
E
cn
= E
vp
E
vn
qV
0
(4.1)
where we have dened the contact potential V
0
which is the potential dierence that an
electron or a hole sees in the vicinity of the junction. If we measure the energy dierence in
electron volts then the contact potential is just the dierence in energy levels.
1
In every textbook I have seen it is always called a p-n junction and never a n-p junction. I have no idea
why but if any of you come across a reason I would like to hear from you.
2
This is not a practical way of making junctions but it does illustrate the physical processes rather well.
67
Figure 4.2: Schematic of the energy levels in a p-n junction.
Figure 4.3: Schematic of the space charge buildup in a p-n junction.
The next thing to note about the junction is that we now have an extremely large con-
centration gradient for both holes and electrons in the region of the junction. Hence from
the current continuity equations we derived early [Eq. (3.86) and Eq. (3.87)] we would expect
that this gradient will drive an electron diusion current from the n-type region to the p-type
region while there will be a hole diusion current from the p-type region to the n-type region.
This diusion current clearly cannot last indenitely and in fact what happens is that the
diusion currents create a region of space charge near the junction as the ions in each region
are xed and cannot move. Recall that in a p-type material the acceptor ions are negatively
charged as they have trapped an electron while in n-type material the donor ions are positively
charged. Thus we would expect a space charge to build up as shown in Fig. 4.3
This space charge will create an electric eld that in turn will drive a drift current that
at equilibrium will exactly cancel out the diusion current and hence no current will ow.
Equating the two currents will allow us to nd an expression for the contact potential as we
will see. I will again work with the equation for holes and I hope that you will do the same
for electrons and we should both end up with the same answer. Our starting point is again
Eq. (3.87):
J
p
(x) = q
_

p
p(x)E D
p
dp
dx
_
(4.2)
68
but as we know J
p
(x) = 0 we can equate the two terms on the r.h.s. of Eq. (4.2) giving us

p
pE = D
p
dp
dx
(4.3)
or

p
D
p
E =
1
p
dp
dx
(4.4)
where we have put all the terms concerning p on one side. We can simplify the l.h.s. of
Eq. (4.4) by making use of the Einstein relationship [Eq. (3.85)] between
p
and D
p
. This
gives:
q
kT
E =
1
p
dp
dx
(4.5)
We now want to solve Eq. (4.5) for the electric eld in terms of the equilibrium hole densities
far from the junction. The easiest way to do this is by writing the electric eld in terms of
the potential V (x), i.e.
E =
dV
dx
. (4.6)
Then substituting this into Eq. (4.7) gives
q
kT
dV
dx
=
1
p
dp
dx
. (4.7)
Now to solve Eq. (4.7) we use a trick that if two expressions are equal then the integral of
the two expressions must also be equal. We integrate both sides of Eq. (4.7) from one side of
the diode to the other (taken here as ). This gives us
q
kT
_

dV
dx
dx =
_

1
p
dp
dx
dx. (4.8)
Now on the l.h.s. of Eq. (4.8) we can change the variable of integrate from x to V while on
the r.h.s. we change the variable of integration from x to p this gives us:
q
kT
_
V ()
V ()
dV =
_
p()
p()
1
p
dp (4.9)
both of which integrals we can easily solve. Writing V () = V
p
, V () = V
n
, p() = p
p
and p() = p
n
, allows us to write Eq. (4.12) as
q
kT
[V
n
V
p
] = ln(p
n
) ln(p
p
) (4.10)
But V
n
V
p
is just the contact potential V
0
and so
V
0
=
kT
q
[ln(p
n
) ln(p
p
)] (4.11)
69
or
V
0
=
kT
q
ln(
p
p
p
n
). (4.12)
In Eq. (4.12) we have expressed the contact potential in terms of the ratio of the equilibrium
hole carrier densities in each material. More commonly we invert Eq. (4.12) to give
p
p
p
n
= e
qV
0
kT
(4.13)
where we have no expressed the ratio of the hole densities in each material in terms of the
contact potential at the junction. The advantage of Eq. (4.13) is that it is still valid if we
replace V
0
by V
0
+ V where V is an applied voltage. Similarly when you do the calculations
for the electron distribution on each side of the junction you get
n
n
n
p
= e
qV
0
kT
(4.14)
The more alert of the readers should by now have noted that we have two dierent de-
nitions of the contact potential. The rst is the shift in the Fermi levels shown schematically
in Fig. 4.2. The second derives from the requirement that there be no net current ow in
equilibrium. We will now show that these are the same. To see this we need with Eq. (3.24)
which relates the equilibrium hole density to the position of the Fermi level. Starting with
Eq. (4.13)
e
qV
0
/kT
=
p
p
p
n
(4.15)
we now use Eq. (3.24) to expand p
p
and p
n
=
N
v
e
(E
fp
E
vp
)/kT
N
v
e
(E
fn
E
vn
)/kT
(4.16)
= e
((E
fn
E
vn
)(E
fp
E
vp
))/kT
(4.17)
but E
fn
= E
fp
in equilibrium and so
e
qV
0
/kT
= e
(E
vp
E
vn
)/kT
(4.18)
or
qV
0
= E
vp
E
vn
(4.19)
and Eq. (4.19) is exactly how we dened the contact potential in Eq. (4.1) and hence our two
dierent denitions of the contact potential are in fact the same, which is reassuring to know.
70
Of course if instead of starting with Eq. (4.13) we had started with Eq. (4.14) and repeated
the derivation we would have obtained
qV
0
= E
cp
E
cn
(4.20)
and you should be able to prove this yourselves without too much trouble.
Finally I would like to re-examine Eq. (4.12) and to express p
p
and p
n
in terms of the
doping concentrations. In an ideal diode the p-type region will be doped N
A
acceptors and
so p
p
= N
A
whereas the n-type region will be doped with N
d
donors. However we can use the
law of mass action [Eq. (3.44)] to write p
n
= n
2
i
/N
d
. Putting these expressions into Eq. (4.12)
we nd
V
0
=
kT
q
ln
_
N
A
N
d
n
2
i
_
. (4.21)
Eq. (4.21) allows one to calculate fairly easily the inbuilt voltage of a diode in terms of the
design specications i.e. the doping levels.
71
'
&
$
%
Problem: An Abrupt Si p-n junction has N
a
= 10
17
cm
3
on the p side and
N
d
= 10
16
cm
3
on the n side. At 300K, (a) calculate the Fermi levels, draw an
equilibrium band diagram and nd V
0
from the diagram. (b) compare the result from (a)
with V
0
calculated from Eq. 4.21
Solution: In the p-type region p
p
= 10
17
cm
3
an we can use Eq. (3.35) to calculate the
dierence between the Fermi level and the intrinsic level giving:
E
i
E
f
= kT ln
_
N
A
n
i
_
= 0.0259 ln
10
17
10
10
= 0.417 eV
Similarly in the n-type region n
n
= 10
16
cm
3
and Eq. (??) relates the intrinsic level to
the Fermi level giving
E
f
E
i
= kT ln
_
N
d
n
i
_
= 0.0259 ln
10
16
10
10
= 0.357 eV
The situation is shown in the gure below
From the gure we can see that the contact potential is given by 0.417 +0.357 = 0.774 V.
(b) From Eq. (4.21) we nd that
qV
0
= kT ln
N
a
N
d
n
2
i
= 0.774 eV
which agrees with the answer in part (a).
4.2.1 The story so far
Up to this point we have considered what happens to the equilibrium electron and hole densities
in an unbiased pn junction. The central result was that the realisation that a pn junction has
an inbuilt potential dierence between the two sections. The size of this potential depends on
the logarithm of the ratio of the free carriers on either side of the junction.
72
Figure 4.4: Schematic of the space charge buildup in a p-n junction.
4.3 Width of the Charge depletion Region
In this section we are going to look at the width of the depletion region more closely. One of
the reasons for doing this is that the width of the depletion region determines the capacitance
of the device and this can be changed by applying a voltage to the diode. Diodes which are
used as variable capacitors are called varactors. In addition the capacitance of a device sets
a fundamental limit on how quickly it can be switched so when designing high speed devices
you will need to know how to determine the capacitance. Lastly in optical applications of pn
junctions such as photocells or photodetectors the width of the depletion region determines
the eciency of the device. The other reason for looking at the depletion region is that once
we have some equations we can get a feel for the numbers involved.
Let us start with the schematic picture of the space charge given in Fig. 4.4. Here I have
indicated the direction of the electric eld as going from right to left (from the positive charge
to the negative charge). The gure also denes the variables x
n
and x
p
as being the width
of the charge depletion regions for the n-type and p-type materials respectively. The width
of the depletion region is given by
W = x
n
+x
p
. (4.22)
The rst thing we need to realise about the charge depletion region is that there are no
free carriers in this region since if there were any they would be swept out by the electric
eld. Hence all of the space charge is built up from donor and acceptor atoms. Thus the total
negative charge in the p-type region is
Q
p
= qN
A
x
p
A (4.23)
where N
A
is the acceptor density in the p-type region and A is the cross sectional area of the
device. Similarly the total positive charge is given by
Q
n
= qN
d
x
n
A (4.24)
where again N
d
is the donor density. Now as the device itself has zero total charge since it
was formed from two electrically neutral pieces of silicon the amount of positive charge must
be equal and opposite to the negative charge, i.e.
qN
A
x
p
A = qN
d
x
n
A (4.25)
73
Figure 4.5: Schematic of the space charge buildup in a p-n junction.
The next step is to calculate the electric eld due to this space charge. This can be done
quite simplely as we have a uniform charge density in each region of either qN
A
or qN
d
and
the electric eld is just the integral of the charge density. For x < 0 this gives us
E(x) =
q

_
x

N
A
dx

(4.26)
but the charge density is zero for x < x
p
so we have
E =
q

_
x
x
p
N
A
dx

(4.27)
=
qN
A

(x +x
p
) (4.28)
Similar nding the electric eld in the n-type region gives
E(x) =
qN
d

(x x
n
) (4.29)
and equating the two electric elds at x = 0 gives us the same equation as Eq. (4.25) for the
overall charge neutrality. In Fig. 4.5 I sketch the electric eld in a typical junction. Note that
the maximum electric eld E
max
is given by
E
max
=
qN
A

x
p
=
qN
d

x
n
(4.30)
From the electric eld we can now calculate the potential dierence across the junction.
As the potential is minus the integral of the electric eld it corresponds to the area under the
curve in Fig. 4.5. As the curve is a triangle the area is just one half the base or W times the
height E
max
or
V =
1
2
WE
max
(4.31)
=
1
2
W
qN
d

x
n
. (4.32)
74
Now we eliminate x
n
from Eq. (4.32) by using Eq. (4.22) and Eq. (4.25) which imply that
x
n
=
N
A
W
N
d
+N
A
. (4.33)
Putting Eq. (4.33) into Eq. (4.32) we nd that
V =
1
2
W
2
q

N
A
N
d
N
A
+N
d
(4.34)
or
W =

2V
q
_
N
A
+N
d
N
A
N
d
_
(4.35)
Similarly we can nd expressions for x
n
and x
p
and these are
x
n
=

2V
q
_
N
A
N
d
(N
A
+N
d
)
_
(4.36)
and
x
p
=

2V
q
_
N
d
N
A
(N
A
+N
d
)
_
(4.37)
Eq. (4.34) and Eq. (4.35) are useful in practical situations as they express either the inbuilt
voltage or the depletion width in terms of design parameters like the doping density. The
other thing to note about Eq. (4.35) is that it is valid even when an external voltage is applied
to the device. In that case it can be seen that increasing the voltage across the diode will
increase the depletion region while decreasing the voltage will decrease the depletion region.
75
'
&
$
%
Example: A silicon diode has a doping density of N
d
= 10
16
cm
3
for the n-type and
N
A
= 10
18
cm
3
in the p-type region. Assuming that an abrupt junction is formed calculate i)
The Fermi levels in the p-type and n-type regions, ii) the contact potential and iii) the depletion
widths on each side of the junction.
To nd the Fermi levels we have a host of equations to chose from. For the p-type region lets
start with Eq. (3.34)
E
f
= E
v
+kT ln
_
N
v
N
A
_
ndequation where N
v
= 2.66 10
19
cm
3
and T = 300K. This gives us that
E
f
= E
v
+ 0.026 ln
_
2.66 10
19
10
18
_
eV
or
E
f
= E
v
+ 0.09 eV
Similarly to nd the position of the Fermi level in the n-type sample we use Eq. (3.31) which
gives us
E
f
= E
c
kTln
_
N
c
N
d
_
where N
c
= 2.86 10
19
cm
3
. This gives us
E
F
= E
c
0.207 eV
and this completes the rst part of the question.
To nd the inbuilt potential we need to combine the two answer we got earlier with the size of
the band gap. Remember that the Fermi level is constant across the junction and we know how
far it is from the conduction band on side and the valence band on the other. The other piece of
information we need is the size of the bandgap which is E
g
= 1.1 eV. The inbuilt potential eV
bi
is just the dierence in the position of the conduction bands on each side. In the p-type region
the position of the conduction band is E
v
+E
g
= E
f
0.09 +E
g
while in the n-type region the
conduction band is at E
f
+ 0.207 and subtracting the two we get
eV
bi
= E
g
0.09 0.207 eV
or
eV
bi
= 0.803 eV
which is the answer to the second part of the question.
To answer the last part of the question we use Eq. (4.36) and Eq. (4.37) with = 11.9 8.84
10
12
F/m. These give us
x
p
= 3.2 10
9
m
and
x
n
= 3.2 10
7
m = 0.32 m
Note that the depletion region is longer in the n-type region as it is less heavily doped by a
factor of 100.
76
4.4 PN Junctions under an applied bias
We are now ready to analyse the behaviour of a PN junction under an applied voltage. You
are probably familiar with the basic behaviour of a diode in that it only lets current pass
through in one direction. Thus it behaves in a profoundly nonlinear way and this is crucial
when it comes to designing logic gates. The two most important facts that we will need in
this section are
1. The drift current across the junction is independent of applied voltage.
2. An applied voltage V changes the potential drop across the junction to V
0
+V
At rst sight the statement that the drift current is independence of applied voltage is
surprising. In Sect. 3.8 when we introduced the idea of the an electrons mobility we showed
that the drift current was proportional to the electric eld. Yet here in the very rst chance
that we get to apply these results to a real device I am claiming that it isnt true. However
a closer inspection of the drift current shows that it is in fact proportional to the product
of the free carrier density and the electric eld. However near the junction region of a p-n
diode there is a region of charge depletion and hence the number density of free carriers is
approximately zero. Hence any change in the electric eld will not produce any change in
the drift current. The free carriers which are responsible for the drift current are generated
within the junction region by thermal excitation and rate that these are produced does not
depend on the applied voltage. In the next module when we look at the interaction of light
with semiconductors we will see how to dramatically alter the drift current, but for now we
can assume that it is a constant.
The second of the two facts we need comes from the fact that we can assume that far
from the junction currents can easily ow and that the eective resistance of the undepleted
regions is zero and hence there is no voltage drop across them. Thus all of the voltage drop
occurs across the charge depleted region.
We can now calculate very simplely the current and voltage dependence of an ideal diode.
For any device the total current is the sum of the drift and diusion currents. For a diode
the drift current is approximately constant and is equal and opposite to the diusion current
at equilibrium. Thus the only additional piece of information we need is how the diusion
current changes with applied voltage. The change in the electron density under an applied
voltage can be understood with reference to Fig. 4.6 which is based upon a similar one in
Singh. The important fact to understand and remember is that in a diode under forward
bias the current is due to the diusion of carriers and not the drift current. Now the rate of
diusion depends exponentionally on the size of the potential barrier between the two regions.
When a voltage is applied to a p-n junction the height of the barrier is reduced (or increased)
and the current ow changes accordingly. More precisely if we change the voltage across the
depletion region from the contact potential V
0
to V
0
V then the free electron density in the
p region will change from n
0
to n
1
. The reason for this increase in the free electron density
is that the diusion current across the junction has increased. Thus we can approximate the
diusion current by the electron density in the p-type region. The total current owing across
77
Figure 4.6: Schematic showing the diusion of electrons across a junction under diering
applied voltages.
the system will then be the dierence between the new diusion current (given by n
1
) and
the new drift current. However the new drift current is identical to the old one (by the rst
assumption above) which is equal to the old diusion current (given by n
0
). Thus the current
owing across the diode will, proportional to the dierence between the two, i.e
I
n
n
1
n
0
, (4.38)
while there will be a similar hole diusion current going in the opposite direction. We can
now use Eq. (4.14) to express the electron density in the p region in terms of the voltage drop
and the electron density in the n region. This gives
I
n
n
n
e
q(V
0
V )/kT)
n
n
e
qV
0
/kT
(4.39)
= n
n
e
qV
0
/kT
_
e
qV/kT
1
_
(4.40)
or
I
n
= I
0
_
e
qV/kT
1
_
(4.41)
where we have introduced the proportionality constant I
0
. In Fig. 4.7 I have plotted a typical
I V graph given by Eq. (4.41). In Eq. (4.41) there were no assumptions made about the
sign of the voltage and it could be either positive (forward bias) or negative (reverse biased).
In the reverse biased case we can see that the current should approach a constant value I
0
irrespective of the applied voltage. Similarly as the voltage increases we see that the current
increases exponentially.
The derivation above introduces all of the important physics involved in the operation of a
p-n diode and it is important that you understand it as it is one of the core objectives of this
78
(a) I-V graph (b) PN diode under applied voltage
Figure 4.7: (a) I-V characteristics of a p-n diode, (b) Schematic for the derivation of the I-V
relationship.
course. The essential thing to remember is that in a p-n diode under forward bias the current
is primarily due to the diusion of carriers across the junction. However it is possible to
analyse the behaviour of the diode in greater detail and in particular we will now obtain an
expression of the constant of proportionality I
0
in Eq. (4.41).
For this derivation we will look at the behaviour of holes and I expect that you will repeat
the derivation for electrons on your own. We will start with Eq. (4.13) which relates the hole
densities on each side of the junction at equilibrium to the contact potential V
0
,
p
p
(V = 0)
p
n
(V = 0)
= e
qV
0
/kT
(4.42)
We will now apply a forward voltage V
f
to the diode as shown in Fig. 4.7b. In this case
the ratio of the carriers far from the junction is given by
p
p
(V = V
f
)
p
n
(V = V
f
)
= e
q(V
0
V
f
)/kT
. (4.43)
Next we divide Eq. (4.42) by Eq. (4.43) to get
p
p
(V = 0)
p
n
(V = 0)
p
n
(V = V
f
)
p
p
(V = V
f
)
= e
qV
f
/kT
(4.44)
Now in most applications the density of injected holes is very much less than the intrinsic
density of holes in a piece of p-type semiconductor. Hence we can assume
p
p
(V = 0) = p
p
(V = V
f
) (4.45)
which allows us to simplify Eq. (4.44) resulting in
p
n
(V = V
f
)
p
n
(V = 0)
= e
qV
f
/kT
(4.46)
79
Figure 4.8: Schematic of the non-equilibrium charge densities in a p-n junction.
or
p
n
(V = V
f
) = p
n
(V = 0)e
qV
f
/kT
(4.47)
Eq. (4.47) gives the non-equilibrium hole distribution in the n-type region. What we need is
the size of the excess hole distribution so we now subtract p
n
(V = 0) from both sides giving
p
n
(V = V
f
) p
n
(V = 0) = p
n
(V = 0)e
qV
f
/kT
p
n
(V = 0) (4.48)
or
p
n
= p
n
(V = 0)
_
e
qV
f
/kT
1
_
(4.49)
In Eq. (4.49) p
n
is the departure from equilibrium of the hole carrier density on the n side
of the junction. Similarly you can show that the change in electron density on the p side is
given by
n
p
= n
p
(V = 0)
_
e
qV
f
/kT
1
_
. (4.50)
Similarly there is a non-equilibrium electron distribution in the p-type region and both the
hole and electron excess distributions are shown in Fig. 4.8. Now we know that any departure
from equilibrium densities drives a diusion current in the material and we now want to work
out the value of this diusion current. If we assume that the voltage supply has been turned
on for a long time then the current ow must be constant through the device and we thus
only need to determine the value of the current at one point in the device.
To determine the current we need to solve the diusion equation Eq. (3.98) for the holes
and Eq. (3.99) for electrons. As we can assume again that the system is in a steady state
then the excess hole/electron distributions decay exponentionally with a characteristic length
L
n,p
. Thus the solutions given in Eq. (3.104) are valid and we can write
p(x) = p
n
(V = V
f
)e
(xx
n
)/L
p
. (4.51)
The solution given in Eq. (4.51) is only valid for x > x
n
where x
n
is the end of the charge
depletion layer in the n type region. Using Eq. (4.49) we nd that
p(x) = p
n
(V = 0)
_
e
qV
f
/kT
1
_
e
(xx
n
)/L
p
(4.52)
80
We can now calculate the diusion current due to the excess hole distribution using
Eq. (3.97):
I
p
(x) = qAD
p
p
x
(4.53)
= qAD
p
1
L
p
p
n
(V = 0)
_
e
qV
f
/kT
1
_
e
(xx
n
)/L
p
(4.54)
where A is the area of the junction, D
p
the hole diusion coecient and L
p
the hole diusion
length. We now evaluate Eq. (4.54) at x = x
n
to give us the total hole current that is injected
into the n type region. This gives us
I
p
=
qAD
p
L
p
p
n
_
e
qV
f
/kT
1
_
(4.55)
Similarly starting with Eq. (4.50) one can work out the electron current injected into the
p type region and you should obtain
I
n
=
qAD
n
L
n
n
p
_
e
qV
f
/kT
1
_
(4.56)
As the hole and electron currents are owing in opposite directions the total current I is then
I = I
p
I
n
(4.57)
or
I = qA
_
D
n
L
n
n
p
+
D
p
L
p
p
n
_
_
e
qV/kT
1
_
(4.58)
I
0
_
e
qV/kT
1
_
(4.59)
which is the same equation as the one we derived earlier.
81
'
&
$
%
Calculation: Consider an ideal diode model for a silicon p-n diode with
N
d
= 10
16
cm
3
and N
a
= 10
18
cm
3
. The diode area is 10
3
cm
2
. Calcu-
late the reverse current given the following transport parameters at 300K:
For the n-side,
p
= 300 cm
2
V
1
s
1
,
n
= 1300 cm
2
V
1
s
1
, D
p
= 7.8 cm
2
s
1
,
D
n
= 33 cm
2
s
1
and for the p-side,
p
= 100 cm
2
V
1
s
1
,
n
= 280 cm
2
V
1
s
1
, D
p
= 2.6 cm
2
s
1
,
D
n
= 7.3 cm
2
s
1
Finally take
n
=
p
= 10
6
s.
Solution: The rst thing we need to do is to calculate the minority carrier diusion
length on each side of the junction. The hole diusion length in the n-type region
is given by
L
p
=
_
D
p

p
= 2.79 10
3
cm
Similarly the electron diusion length in the p-type region is L
n
= 2.7 10
3
cm.
The next thing we need to calculate are the minority equilibrium carrier densities
n
p
and p
n
on each side of the barrier. On the n-side we can use the law of mass
action with n = N
d
to calculate p
n
giving
p
n
= n
2
i
/N
d
= 10
4
cm
3
similarly n
p
= 10
2
cm
3
. We can now put these values into Eq. (4.58) to nd a value
for I
0
. The resulting answer is
I
0
= 10
14
A.
Strictly speaking Eq. (4.58) denes the current across the junction not the current through
the diode. However in the steady state regime the current through the device must be constant
at all points as otherwise the would be charge buildup at some point and this would lead to
extra elds being generated and the problem would not be independent of time. Thus we can
take Eq. (4.58) as dening the current for the whole diode.
The other factor to notice about the current across the junction of a diode is that it is
driven by the diusion of minority carriers, i.e. electrons diuse into the p-type region while
holes diuse into the n-type region. However far from the junction the current must due to
the majority carriers at the extreme end of the device electrons enter the n-type region from
the battery and holes enter the p-type region. So throughout the device the current must
switch from being due to the drift of majority carriers to the diusion of minority carriers.
The majority carrier current can be easily worked out given that the total current is xed
and we know that the excess minority carriers recombine with the majority carriers over a
typical length scale given by the diusion length. Hence the hole and electron currents must
look something like that given in Fig. 4.9.
82
Figure 4.9: Schematic of the dierent currents in a p-n junction. The solid line shows the
electron current while the dashed line shows the hole current. Notice that in each case
the current is constant across the depletion region and that the minority currents decay
exponentially.
4.5 Behaviour of a diode under reverse voltage
In the previous section we derived an equation for the I-V relationship of an ideal diode.
In the derivation no assumptions were made about the sign of the applied voltage and so it
applies equally well to both forward and reverse biased diodes. When the applied voltage is
negative Eq. (4.59) implies that the current through the diode should be almost independent
of voltage and have a value of
I
r
= qA
_
D
n
L
n
n
p
+
D
p
L
p
p
n
_
. (4.60)
However in real devices this is only true up to a certain applied voltage value. Once this
threshold value is reached the current rapidly increases without limit and the voltage at
which this happens is called the breakdown voltage. However this should not be taken to
imply that the diode itself breaks, it the current is limited by an external circuit to prevent
over heating the diode itself will not be damaged. There are two processes by which a diode
can breakdown and these are:
1. Zener Tunneling
2. Avalanche Breakdown
and we will look at each in turn.
4.5.1 Zener Tunneling
Zener tunneling is a quantum eect and is caused by electrons in the valance band in the
p-type region tunneling through the depletion region to the relatively empty conduction band
of the n-type region. This is exactly the same process as we looked at in Sect. 2.5 when we
83
Figure 4.10: Schematic of the bands in a Zener diode at zero applied voltage and a reverse
voltage sucient for tunneling to occur.
derived an expression for the transmission through a rectangular barrier. In Fig. 4.10 we
show the situation for electrons tunneling through the diode.
From Fig. 4.10 it can be seen that tunneling will only occur when the reverse bias is
suciently great so that the conduction band in the n-doped region is below the top of the
valence band in the p-doped region. This is as the electron cant gain any energy during
the tunneling process and it cannot tunnel into the bandgap as there are no energy states
available for it to occupy. It can also be seen that the potential barrier the electron has to
tunnel through is roughly triangular in shape. Using standard techniques it can be shown
that the probability of tunneling through a triangular barrier of height E
g
is given by
T e

2m

E
3/2
g
2eF
(4.61)
where m

is the eective mass of the electron and F is the strength of the electric eld in the
depletion region. The electric eld strength also implicitly denes the width W of the barrier
since the peak height of the barrier is E
g
and the slope is given by the strength of the electric
eld. From Eq. (4.61) we can see that the higher the electric eld the greater the probability
of tunneling. This is due primarily to the decrease in width of the depletion region. Thus
the best way to make ensure that Zener tunneling happens is to use diodes which are heavily
doped and so have a small depletion region.
One attraction of using Zener diodes is that the breakdown voltage can be set by the doping
concentrations and so it is relatively simple to fabricate diodes with any desired breakdown
voltage. In a Zener diode during breakdown the voltage across the diode is xed and hence
can be used as voltage regulators.
4.5.2 Avalanche Breakdown
The second eect that can occur in a diode under reverse bias is that of avalanche breakdown.
This typically occurs at higher voltages than Zener tunneling and is the predominant eect
in non heavily doped diodes. The process of avalanche breakdown is shown schematically in
Fig. 4.11. The process starts with a single electron being accelerated across the depletion
region. If the electric eld is high enough then it will have acquired sucient energy to ionise
an atom upon collision. This will create an addition electron hole pair and each of these new
particles is also accelerated and upon collision create new electron hole pairs. Thus if there
84
Figure 4.11: Schematic of the avalanche breakdown process. Each electron is accelerated
across the depletion region and acquires sucient energy to ionise atoms upon collision.
are n collisions then we would expect 2
n
particles to have been created. The process thus
snowballs and results in a large current owing through the device.
To try and further understand the process we can use a simple model in which an electron
has a probability P of creating an electron-hole pair. Then the probability of creating two
electron-hole pairs is given by P
2
and similarly the probability of creating n electron-hole pairs
is given by P
n
. Thus if we have n
in
electrons to begin with the total number of electrons n
out
at the output of the device will be given by
n
out
= n
in
_
1 +P +P
2
+P
3
+
_
(4.62)
and we can thus dene a multiplication factor M as
M =
n
out
n
in
=
1
1 P
. (4.63)
Now clearly M must depend on the applied voltage in some fashion. If we dene the break-
down voltage V
br
as being the voltage for which M then empirically it is found that
M =
_
V
V
br
_
n
(4.64)
where the exponent n ranges from three to six depending on the material used.
However a moments thought should be enough to tell you that what I have said so far
about avalanche breakdown is only half the story. Since avalanche diodes can be used in real
circuits as voltage rectiers then current through such a diode must be constant else a charge
would build up over time and this is not what is observed. Hence in any steady state device
the multiplication factor M must equal unity. What happens in a real device is even as some
of the electrons are being accelerated and ionising the atoms in the depletion region other
electrons must be entering the depletion region only to be trapped by the ionised atoms and
combine to form neutral atoms which can then be ionised again.
However in the non-steady state regime the avalanche eect can be observed and does
result in a transient current which is greater than the input current. The most common use
85
Figure 4.12: Current passing through a diode being abruptly switched o at time t = 0.
for such diodes is as high gain optical detectors. The other factor that has been missing in
this discussion is any mention of where the initial electron appears from. In most devices the
answer is simplely due to the thermal tail on the Fermi-Dirac distribution which ensures that
there are always a few free electrons in the conduction band. The second possibility is that
an electron hole pair was created by the absorption of a photon. In this case the avalanche
diode (or avalanche photo-diode as it often called) acts as an extremely high gain detector
producing a pulse whenever a photon is absorbed. In this way extremely high quantum
eciency detectors can be made, the record eciency for a single photo detector is greater
than 96% which means that 96% of the time when a single photon is incident on the device
it will be detected.
Commonly diodes which are used as voltage rectiers are called Zener diodes even though
often the breakdown mechanism is through avalanche breakdown rather than Zener tunneling.
4.6 Transient Eects in Diodes
Up until this point I have mainly considered the behaviour of semiconductors in the steady-
state regime, i.e. nothing changes with time. And in the steady state regime we have shown
that diodes behave like rectiers, i.e. they only pass current in one direction. However we
now want to consider what happens when we applied a time varying signal to a diode.
To understand how the transient behaviour of diodes occurs it is necessary to keep in mind
that under forward bias the current across the junction is primarily due to the diusion of
minority carriers and in the steady state regime there is an excess hole distribution p(x) in
the n-doped region and a excess electron distribution n(x) in the p-doped region. These were
shown schematically in Fig. 4.8 and dened by Eq. (4.49) and Eq. (4.50). When discussing
excess carrier distributions in Section 3.10.1 it was stated that such carrier distributions
decayed exponentially with a time constant
n,p
. When the voltage across the diode is changed
then we would expect that it would take some time for the excess carrier distributions to evolve
to the correct values for the new applied voltage and thus we can expect transient behaviour
of some sort.
In Fig. 4.12 we show the eects of a turning o the current through a diode abruptly.
For t < 0 the diode is forward biased and a current I ows through the diode and then at
time t = 0 the circuit is broken and we want to know what how the voltage across the diode
varies with time. Now there is an excellent discussion of this in Streetman so I will not go
into too much detail but the essential physics is that the voltage across the device depends
on the excess minority carrier distributions which relax with a time evolution being primarily
86
Figure 4.13: Switching circuit for a diode along with the driving voltage as a function of time.
determined by the relaxation constant . The typical time evolution behaviour can be see in
Fig. 4.12 which is similar to Fig. 5-27 in Streetman. Anyway the nal result that you arrive
at is that the voltage across the diode v(t) decays as
v(t) =
kT
q
ln
_
I
p
qAL
p
p
n
e
t/
p
+ 1
_
(4.65)
The important thing to remember is that in a real diode (or any other device for that
matter) the voltage cannot be switched instantaneously. In a diode however there are several
ways to speed up the response. The rst is to use short devices since a short diode will store
less charge than a long diode and so will have a quicker response. The second is to alter the
relaxation time by altering the material somehow. The most common way to do this is to
introduce defect states into the band gap with energy levels near the middle of the bandgap.
These defect states can then capture an electron and a hole which then recombine. Practically
this can be achieved by doing the device with gold or some other metal during the fabrication
process.
The next form of transient behaviour I want to discuss is the current owing through a
diode when it is switched from being forward biased to reverse biased. A typical circuit is
shown in Fig. 4.13 along with the driving voltage. In Fig. 4.13 the diode is being driven by a
square wave which switches between E V. The aim is to try and understand the the time
response of the current when the voltage is switched.
If we initially assume that the voltage is E V then the diode is forward biased and we
know that its resistance is eectively zero. In this case the voltage drop will be mostly across
the resistor R and hence the current in the circuit will be i = I
f
E/R. If the driving
voltage now switches from E V to E V the current owing through the diode must reverse
to I
r
E/R. To understand this recall that previously when discussing the voltage across
a diode we saw that it decayed with a time constant and hence there will be a short period
of time when the diode has a small positive voltage across it allowing a large negative current
to ow.
However this situation cannot last as the voltage across the diode is due to the presence
of minority carriers which will recombine with the majority carriers now that the voltage is
reversed. This will gradually cause the depletion layer to increase and we would expect that
the voltage across the diode will eventually be approximately equal to E and hence very
little current will ow in the circuit. Thus the current response will be as shown in Fig. 4.14.
87
Figure 4.14: Driving voltage and associated diode current for the circuit in Fig. 4.13.
4.7 Capacitance of p-n Junctions
In the previous section we saw how the stored charge in the form of excess minority carriers
resulted in transient behaviour when a diode was switched. The presence of a stored charge
can be associated with a capacitance and in this section we are going to look at the capacitance
of a p-n junction. There are in fact two main sources of capacitance in a p-n junction, the
rst is due to the presence of the depletion layer which was discussed in Section 4.3 and the
second is due to the excess minority carrier distributions which was we have been discussing
in the previous section. The rst contribution is often called the depletion layer capacitance
while the second contribution is refered to as the diusion capacitance.
The depletion layer capacitance is perhaps the most easily understand. If you have another
look at Fig. 4.4 which shows the space charge build up near the junction then you can see
that it looks very much like a picture of a parallel plate capacitor except that the gap between
the plates is zero. Hence we would expect that such a device might act like a capacitor under
some circumstances. To nd an expression for the capacitance of a diode we need to start
with the denition of capacitance C which is
C =

dQ
dV

. (4.66)
We now need to use the equations we derived earlier in Sect. 4.3 for the width of the charge
depletion region W i.e.
W =

2(V
0
V )
q
_
N
A
+N
d
N
A
N
d
_
(4.67)
where V
0
is the contact potential and V is the applied voltage which can be positive or
negative. Now the charge stored in the region was derived earlier [Eq. (4.24)] which we can
write as
Q = A
_
2q(V
0
V )
N
A
N
d
N
A
+N
d
_
1/2
(4.68)
where we have used Eq. (4.36) to express the width of the depletion region on the n-doped
side in terms of the doping densities. The capacitance can then be easily worked out using
Eq. (4.66) and the result is
C =
A
2
_
2q
V
0
V
N
A
N
d
N
A
+N
d
_
1/2
. (4.69)
88
The important thing to note about Eq. (4.68) is that the capacitance is proportion ot (V
0

V )
1/2
and thus it is a strong function of the applied voltage. In circumstances where it is
useful to have a voltage varying capacitor then a diode can be used in reverse bias and such
devices are call veractors. One such application is in the tuning of an LC resonator circuit
where the resonant frequency = 1/

LC is for the case of an ideal diode proportional to


the applied voltage.
Another application of the deletion layer capacitance occurs when one side of the diode
is very heavily doped so that for example we have a p
+
n junction in which N
A
N
d
. In
this case we can approximate Eq. (4.68) as
C =
A
2
_
2q
V
0
V
N
d
_
1/2
. (4.70)
Now if we applied a large reverse bias so that |V | >> V
0
then the capacitance is proportion
to the minority carrier density and this method allows us a simple means to measure N
d
.
The second form of capacitance in a p-n junction is due to the presence of excess minority
carriers which is responsible for the transient eects described before. This diusion capac-
itance dominates over the depletion layer capacitance for forward bias as Eq. (4.69) shows
that the diusion capacitance goes to zero as V V
0
. In contrast the diusion capacitances
goes to zero when the diode is reverse biased and in that case the depletion layer capacitance
dominates. I am not going to spend any time in discussing the details of the diusion capac-
itance as there is an excellent discussion in Streetman and those of you who are interested in
knowing more should read the relevant sections. However you all should be able to explain
the two dierent types of capacitance in a diode and give the physical origin of each one.
4.8 Junctions between two dissimilar materials
Up until this point we have always been discussing a junction between two similar pieces of
semiconductor. In Sec. 4.2 when discussing the bandgap of a p-n junction I stated that at
thermal equilibrium the Fermi level must be constant throughout the material. This principle
then allowed us to determine the correct way to join up the band diagrams of the two samples.
In this section we are going to look at junctions between both a semiconductor and a metal
and between two dissimilar semiconductors. The aim of this section we be to try and gain
an understanding of how to draw the band diagram for such structures and how to then
determine their basic electronic properties.
The starting point for this analysis will be the principle that in thermal equilibrium the
Fermi-level of any two samples must be the same. Otherwise if this were not the case then
electrons would ow from the sample with the higher Fermi-level to the other sample and the
presence of a current would imply that the samples werent in thermal equilibrium.
4.8.1 Junctions between a metal and a semiconductor
In any practical semiconductor device there must be some way to wire it up to an external
circuit. This implies that at some point there must be a junction between a metal and the
semiconductor. Indeed looking at a typical integrated circuit under a microscope reveals that
most of the surface is covered with metallic layers which act as thin wires connecting dierent
parts of the chip. Clearly we when connect a semiconductor to a wire we would ideally like
89
Figure 4.15: Band diagrams for a metal-semiconductor junctions. The gure on the left shows
the isolated band diagrams relative to a common vacuum level while the gure on the right
so the results when the materials are joined.
that current can easily ow in both directions and that the resistance of the junction should
be independent of the voltage across it. Such a contact is called an ohmic contact. However
as we will see things are not always this simple.
We now consider the simplest situation, that of a metallic contact between a metal and
a n-doped piece of semiconductor. The separate band diagrams for each material are shown
in Fig. 4.15(a) where I have included a common vacuum level as a reference point for both
materials. We now need to consider what happens when the materials are joined. As stated
above the rst thing we need to do is to bring the Fermi-levels to the same level. Note that
as the semiconductor is n-doped and the conduction band of the semiconductor lies above
the Fermi-level of the metal initially electrons will ow from the semiconductor to the metal.
This has the eect of reducing the number of free electrons in the semiconductor near the
junction and hence the Fermi-level must lie deeper in the bandgap close to the metal than it
does further away from the metal. This is the explanation for the upward curvature of the
bands in Fig. 4.15(b).
From the gure and the level of the bands we can immediate see that after forming the
junction a space charge region will be set up near the junction with excess negative charges
being present in the metal and an absence of free electrons leading to a positive charge in the
semiconductor. Physically this is very similar to the charge depletion layer in a conventional
p-n junction and it has the same eect. As in a p-n junction this space charge leads to an
electric eld which grows until the drift current induced by the electric eld exactly cancels
out the diusion current and a steady state is reached.
From what we know about how a p-n junction behaves we can understand what the eects
of applying a voltage to the metal-semiconductor junction will be. If we apply a forward bias
the the potential barrier that electrons have to cross to reach the metal will be reduced and so
a current will ow. However if we apply a reverse bias then the barrier height will be increase
and we would not expect any current to ow. Thus we would expect a I-V relationship of the
form:
I = I
0
_
e
qV/kT
1
_
. (4.71)
A metal-semiconductor junction which behaves in such a manner is called a Schottky Barrier
diode. The main dierence between a Schottky barrier diode and a p-n diode is that in a
90
Schottky barrier diode the current is always carried by the majority carriers in each region,
i.e. in this case by electrons. In practice this means that such diodes have a smaller diusion
capacitance than a conventional p-n junction as there is no charge storage due to the eects
of minority carriers. Thus Schottky barriers can be switched faster than a conventional diode.
In addition Schottky barriers require fewer fabrication steps than a conventional p-n junctions
and can be made smaller. Thus in some applications it is preferable to use a Schottky barrier
diode rather than a p-n junction.
4.8.2 Ohmic Contacts
We saw how in the previous section a junction between a semiconductor and a metal can
result in a recting junction. However in many applications we would want a ohmic contact
in which current can easily ow in either direction with little resistance. From the previous
section you might think that creating a Ohmic contact is as simple as nding a metal whose
work function
3
is smaller than that of the semiconductor. Indeed if life was simple then that
would be all there was too it. However most common metals have work functions greater that
that of most semiconductors. In addition there are additional eects which tend to ensure
that any simple junction between a metal and a semiconductor forms a recting junction.
The most important of these are surface defects on the edge of the semiconductor. These
surface defects create additional levels inside the bandgap of the semiconductor which tend
to x the Fermi-level close to the top of the bandgap of the semiconductor.
The most common way to create a ohmic contact is to heavily dope the semiconductor
near the metallic contact. The idea behind this is very similar to the tunneling eect we
discussed in relation to Zener diodes. In a metal semiconductor junction we again have a
charge depletion region and the more heavily doped the sample is the smaller the size of the
depletion region is. In practice the creating of a good ohmic contact is more an art than a
science.
4.8.3 A Junction between two dissimilar semiconductors
The last junction I want to discuss is that between two dierent semiconductors. In practice
a dierent semiconductor can be grown on top of another one if the lattice spacing (i.e. the
distance between the atoms) is similar for both semiconductors. One good example of this
the growth of Al
x
Ga
1x
As which for low levels of aluminum has essentially the same lattice
constant irrespective of the actual composition. Such junctions are called hetrojunctions
and play an important role in the fabrication of optoelectronic devices such as lasers and
ampliers.
To understand how to draw the band diagram for a hetrojunction the only important
thing to remember is that as always the Fermi-level must be a constant. We will have a
look at the band diagram for a junction between a wide bandgap p-type semiconductor and a
narrow bandgap n-type semiconductor. The individual band diagrams are shown in Fig. 4.16.
Here again there will be a discontinuity in the energy levels as the bandgap has a dierent
value on each side of the junction.
The attraction of using hetrojunctions arise when samples with multiple hetrojunctions
can be fabricated. In Fig. 4.17 I show typical band diagrams for a simple multiple hetro-
3
Recall back in Module 1 we dened the work function of a metal as being the energy needed for an electron
to escape its surface.
91
Figure 4.16: Band diagrams for a typical hetrojunction. The gure on the left shows the
isolated band diagrams relative to a common vacuum level while the gure on the right so
the results when the materials are joined.
Figure 4.17: A typical use of multiple hetrojuctions to create an potential well for electrons.
junction. Looking at the structure of the conduction band it should hopefully remind you of
the nite potential well that we studied in Module. 2. And in fact the structure in Fig. 4.17
behaves exactly like a nite quantum well and has all the properties that you would expect
from it. Fabricating hetrojunctions in this manner allows us to localise the electrons in a de-
vice. This is often used in optoelectronic devices such as lasers and ampliers. In an optical
device typically one fabricates a waveguide which is the optical equivalent of a potential well
and connes the light to a well dened region of space. If in addition you can conne the
electrons to the same region by use of hetrojunctions then you can create strong interactions
between the light and the electrons and build compact lasers, ampliers and detectors.
92
4.9 Problems
1. Prove that the widths x
n
and x
p
of the depletion regions are given by Eq. (4.36) and
Eq. (4.37).
2. Why does the potential in a p-n diode fall mainly across the depletion region and not
across the neutral region?
3. An Abrupt silicon p-n diode at 300K has a doping of N
a
= 10
18
cm
3
and N
d
=
10
15
cm
3
. Calculate the built in potential and the depletion widths in the n and p
regions.
4. Assume that the diode in question 2 is circular with a diameter of 50 m. Calculate the
charge in the depletion region and plot the electric eld prole in the diode.
5. A Ge p-n diode has N
A
= 5 10
17
cm
3
and N
d
= 10
17
cm
3
. Calculate the built in
voltage at 300K stating what assumption you use. At what temperature does the built
in voltage decrease by 1 %?
6. Explain, using physical argument, why the reverse current in a p-n diode does not
change with bias (before breakdown). Would this be the case if the electron and holes
had a constant mobility independent of the electric eld?
93
Module 5
Interactions between light and
matter
Learning Outcomes:
1. Explain the dierence between stimulated and spontaneous emission.
2. Explain the size of the bandgap determines which wavelengths a semiconductor can
absorb.
3. Explain the dierence between a direct and indirect semiconductor.
4. Explain how pn junctions can be used to detect and emit light.
5. describe how a light emitting diode works and how its spectral properties can be im-
proved.
5.1 Introduction
In the last module we looked at the basic electrical properties of a pn junction. In this module
we aim to look at its optical properties. PN junctions form the basis for a wide variety of
optoelectronic devices such as solar cells, photodiodes and lasers. In order to understand
the optical properties of pn junctions it is necessary to start with the basics of interactions
between photons and matter.
Recall that at the very start of the course I stated that Einstein introduced the idea of
a photon and that a photon with a frequency has an energy E = . He used this to
explain the photoelectric eect whereby an electron would gain a quanta of energy when it
absorbed an electron. Bohr extended this idea to the hydrogen atom with his postulate that
when an electron moved from a higher energy level to a lower energy level it would emit a
photon with a wavelength corresponding to the energy dierence of the two levels. Similar an
electron could only absorb a photon if the energy of the photon matched the energy dierence
between the electrons state and another higher energy level.
Nowadays we would extend this to any two energy levels. Giving rise to the three diagrams
shown in Fig. ??. Here we have two unspecied energy levels called the ground state and
the excited state. The energy dierence between these two levels corresponds to a frequency
94
dierence of . The rst diagram in Fig. ?? corresponds to absorption of a photon by an
electron resulting in the electron moving from the ground state to the excited state.
The second diagram in Fig. ?? corresponds to the inverse process, that of spontaneous
emission. Here the electron emits a photon and moves back down to the ground state. It is
this process that results in the spectral lines seen in the emission of light from gases.
The last diagram in Fig. ?? corresponds to the case when an electron is in the excited state
and a photon of frequency propagates through the material. This photon stimulates the
electron to move from the excited state to the ground state and in the process emit a photon
with the same direction and phase as the original photon. This process is called stimulated
emission and is the basic physical principle behind the laser
1
Of course the rst thing we need to do is to apply this to semiconductors. And to do
so we need to identify appropriate energy levels corresponding to the ground and excited
levels in Fig. ??. The obvious candidates are the valence band for the ground state and the
conduction band for the excited state. In other devices the ground and excited states might
be two energy levels in a nite square well or quantum hetrojunction. Once we do this we can
immediately see that the size of the bandgap in a semiconductor will determine whether or
not a photon will be absorbed, since if the photons energy is less than the size of the bandgap
then it will not be able to excite an electron from the valence band to the conduction band
and will instead pass through the material. On the other hand if the energy of the photon is
greater than the size of the bandgap then it is likely to be absorbed.

Work out the wavelength corresponding to the size of the bandgap in Silicon and
Germanium.
To understand this process in some more detail we need to write down some simple
equations. We will consider the three processes in Fig. ??. If we consider light passing
through a semiconductor then the rate of change in intensity
2
I can be written as:
dI
dz
= A
10
N
1
I A01N
0
I +B
10
N
1
(5.1)
where N
1,0
are the number densities of electrons in the excited and ground states respectively.
The coecients A
10
, A
01
and B
10
describe the relative probabilities of stimulated emission,
absorption and spontaneous emission respectively. Now as stimulated emission and absorption
are the opposite process we have A
10
= A
01
. And we will ignore the value of B for now
3
.
In many cases of interest the spontaneous emission can be neglected in which case we can
write Eq. (5.1) as
dI
dz
= A
10
(N
1
N
0
) I (5.2)
which has the simple solution:
I(z) = I
0
e
z
(5.3)
where gamma is given by
= A
10
(N
1
N
0
) . (5.4)
1
It was Einstein who rst realised on thermodynamical grounds that such a process was necessary and thus
he is often regarded as the father of the laser
2
The intensity is just proportional to the number of photons
3
By using some elementary thermodynamics one can nd a similar relationship but that would take us too
far o course.
95
Note that the solutions to Eq. (5.3) are very dierent depending on whether or not is greater
or less than zero. If is positive then the intensity of the light will increase exponentially
with distance, i.e the material acts as an amplifer. This can only happen if there are more
electrons in the excited state than the ground state and this is called a population inversion.
If there are more electrons in the ground state than the excited state then is negative and
light will be absorbed as it passes through the crystal.
So the next question to ask is which is the normal situation? The answer to that is there are
more electrons in the ground state than the excited state. The reason for this comes from the
Fermi-Dirac distribution. Remember that when a sample is in thermodynamical equilbrium
then the electrons obey the Fermi-Dirac distribution which decreases monotonically with
energy and hence there will always be more electrons in the ground state than the excited
state.
However the picture is not really that bleak. The Fermi-Dirac distribtion only holds for
materials in thermoequilbrium and not all materials are. In a semiconductor for instance you
can create nonequilbrium distributions by injecting currents into the device. This will locally
give you an excess number of free electrons and so you will be able to see gain.
96

También podría gustarte