Why Things Are The Way They Are - B. S. Chandrasekhar
Why Things Are The Way They Are - B. S. Chandrasekhar
Why Things Are The Way They Are - B. S. Chandrasekhar
The book is intended for general readers, and uses mainly words,
pictures and analogies, with only a minimum of very simple
mathematics. Even so, the author explains the basic ideas of
quantum mechanics, the laws .of which, although unfamiliar to most
people, govern the behaviour of all matter in the universe. He
explains how it is possible to understand the basic properties of
matter using these ideas, and translates the technical jargon of
physics into a language that can be understood by anyone with an
interest in science who wants to know why the world around us
behaves in the way that it does.
B. S. CHANDRASEKHAR
WHY
THINGS
ARE
THE
WAY
THEY
ARE
Cambridge
UNIVERSITY PRESS
A catalogue record for this book is available from the British Library
Library of Congress Cataloging-in-Publication Data
Chandrasekhar, B. S. (Bellur Sivaramiah), 1928—
p. cm.
QC173.454.C49 1998
CONTENTS
PREFACE IX
I INTRODUCTION I
II CRYSTALS 8
IV THE ATOM 63
V STATISTICAL PHYSICS 84
XI MAGNETS 2O2
GLOSSARY 238
INDEX 251
PREFACE
A friend once remarked ruefully to me, “Whenever I hear someone
talk about a well-rounded person, I think of a spherical object with no
distinguishing features on it.’ A closer examination of such a person
does reveal some features; these, however, are likely to be
concentrated in areas that are labelled literature, art, music,
philosophy, history and others generally called the liberal arts or the
humanities. The area called physics, on the other hand, usually
tends to be blank, although physics is as impressive an achievement
of the mind as any of the others. A reason for this state of affairs is
contained in the response I often get from someone who hears that I
am a physicist: ‘Physics was my most difficult subject in school. I
never understood it.’ Instead of giving up, I try to explain to the
person some problem I am working on at the moment, maybe in
superconductivity. I use words, analogies, pictures drawn on a paper
serviette: anything but mathematics that is more advanced than
simple arithmetic. I am then, even if not always, rewarded by the
person with a dawning of interest and a desire to learn more.
Having enjoyed this activity during all these years that I have myself
been learning physics through teaching and research, I decided to
gather and organize these ideas into a book for people (and I believe
there are many) who are curious about the physicist’s picture of the
world, especially the part which forms our immediate surroundings.
In writing this book, I have avoided certain features which are found
in some books on popular science for the lay reader. To follow the
line of reasoning in an unfamiliar field requires some effort, and I
have left out things which seem to me irrelevant or distracting. I have
drawn the figures so that they illustrate a point in the text, and
nothing IX
PREFACE
more. They are in effect the sketches on a serviette that I mentioned
earlier. The reader will find no photographs of physicists or of
equipment; it is not clear to me that looking at a picture of Enrico
Fermi helps one to understand what a fermion is, or that gazing upon
a photograph of tubes, cables, electronic instruments and assorted
hardware clarifies the photoelectric effect. I describe the subject as it
is today, and not its history. I name a few of the physicists who made
signal contributions, but do not try to be exhaustive. I do not identify
their nationalities or prizes won: physics is an endeavour which
knows no national boundaries, and is not some kind of Olympics. I
do not list books for further reading, because I think that readers
whose interest is aroused will do much better browsing in libraries
and bookshops and making their own discoveries.
B. S. CHANDRASEKHAR
GROBENZELL, GERMANY
INTRODUCTION
When you use a plastic spoon to stir hot coffee, the handle does not
get hot, but if you use a silver spoon, it does. A sheet of glass is
transparent so that you can see right through it, but a much thinner
sheet of aluminium foil is opaque and lets no light through. You can
bend a piece of copper wire back and forth without breaking it, but a
glass rod is brittle and will break if you try to bend it. A magnet will
attract an iron screw, but not a brass screw. A piece of copper is
reddish in colour, while silver is -
Thus all the different atoms, one for each of the approximately one
hundred elements that exist in the universe, consist of just three
kinds of objects: electrons, protons and neutrons. Yet when they are
put together to give us the things we see around us, the result is a
great range of different properties. One may well wonder how such
complexity can result from putting together atoms composed of just
three simple entities: electrons, protons and neutrons. Indeed, we
shall see later that we need to consider only two entities, namely
electrons and nuclei, to understand such properties of matter. That a
nucleus is itself composed of protons and neutrons will turn out to be
largely irrelevant for this purpose, though it is of crucial importance
when we want to understand specifically nuclear properties such as
radioactivity and nuclear fission.
There must be some other force between the atoms that has nothing
to do with their mass, and this force must lead to the rigidity of the
solid. We shall see later that electrons and nuclei have a property
called electric charge, or sometimes simply charge, and this is the
source of the force that we are seeking.
So pervasive is the chip now, that one can give the old saying a new
meaning: ‘Chips with everything!’ Some other examples of practical
applications which grew out of the understanding of the physics of
condensed matter are: new magnetic materials used in electric
motors, glass fibres carrying telephone conversations, liquid crystals
and light-emitting diodes which display the numbers in pocket
calculators, superconducting magnets for magnetic imaging systems
which are used in medical diagnosis, high-strength alloys for jet
engines.
I can go one further step and denote the result of each of the above
operations also by a symbol. For example, I could write the equation
x +y = z.
x = vxt
where x stands for the distance, v for the speed and t for the time of
travel. This equation is read ‘x equals v times f’. If the object
happens to be moving at 106 centimetres per second, then I get the
distance it travels in 2 seconds, say, by multiplying its speed,
106cm/sec, by 2 sec. The equation applies equally well for any
speed and any time of travel. You can see that the equation with its
symbols gives us a convenient shorthand for the general statement,
‘The distance travelled by any object which is moving at a constant
speed is given by multiplying its speed by its time of travel.’
There is one other matter of notation that I have just used, when I
wrote a number as 106, that may not be familiar to all of you. It is
called the exponential (and when used in pocket calculators, the
scientific) notation, and is useful in writing down numbers that are
either very large or very small. For example, the number of atoms in
one gram of a solid is about
1,000,000,000,000,000,000,000.
100,000,000 centimetre.
The actual numbers will depend on what kind of atom it is, since the
atoms of different INTRODUCTION
In chapter IV, I apply the ideas of chapter III to an atom and find that
quantum mechanics tells us that the atom is not at all what we had
imagined it to be, namely a tiny hard ball. The philosopher
Democritus said a long time ago that if one took a piece of matter
and divided it into smaller and still smaller pieces, one would finally
get to a piece that was so small that it could not be divided any
further, and therefore he called it an atom. One might then think that
an atom of, say, silver would be a tiny speck of silver with a silvery
sheen, quite different from a copper atom which would be reddish in
colour, or a glittering atom of gold. The reality turns out to be quite
different, as we shall see in chapter IV.
Suppose I want to talk about the individual ages of the 248 million
inhabitants of the United States. I do not even begin the hopeless
task of listing the age of each person, all 248 million of them. Rather,
I make a list of the numbers of people whose ages lie between
specified limits, say 0—9 years, 10—19 years, and so on. I now
have a compact description of the age distribution of the population,
and have taken the first step into the subject called statistics. When
we use this method in physics, we call it statistical physics. I shall
introduce you to this topic in chapter V. In chapter VI, I replace the
hard-sphere atom of chapter II with the quantum-mechanical atom of
chapter IV, apply the statistical ideas of chapter V, and discover what
happens to the electrons when the atoms find themselves in the
solid. By doing so, I arrive at a picture of what a solid is like. It will
look very different from the orderly stack of spherical atoms which I
had in chapter II. I find that this new picture gives an explanation of
why the solid holds together and does not disintegrate into a pile of
atoms.
be able to answer the questions with which this chapter opened, and
many more. You will have a new way of looking at the world around
you, a way which has its own beauty and which adds to the world-
views of art, music, and literature.
You will find, as you read on, questions interspersed in the text. You
can decide whether or not to try to answer each question first before
looking at my answer which follows. In any event, please do not skip
over the questions; they do form an integral part of the whole story.
II CRYSTALS
formation. They would then form a single orderly pattern of rows and
columns. If on the other hand I divide them into ten units, each of
one hundred soldiers, and then ask them to fall into parade formation
at ten different points on the parade ground, but do not tell them in
which direction they should face, I would end up with a different
result. It is true that each unit will look exactly like every other unit,
but while all the soldiers in a given unit will be facing in the same
direction, this direction is very likely to be different for the different
units. Something like this is what happens when the liquid is cooled
rapidly, leading to a polycrystalline solid. It is possible to assign an
‘orientation’ to each grain in such a solid, somewhat like calling the
direction in which the soldiers of a given unit are facing as the
‘orientation’ of the unit. If we look at the ordered pattern of atoms in
the whole crystal, we find that the order will be different in different
directions, analogous to what is illustrated in fig. I-1. It shows a tree
nursery, with the seedlings planted in a regular pattern. You can see
that the plants are spaced apart by different distances in different
directions, so that it is possible to distinguish one direction from
another.
I can also look in another way at what the difference is which causes
the liquid to end up as a crystal or as a glass: namely in terms of
what the atoms are doing as time goes on in the two cases. If I do
the freezing slowly enough, the atoms have enough time to find their
way to the positions in the solid where they would most like to be, so
to speak, and the result is a crystal. If on the other hand I speed up
the process sufficiently, the atoms just do not have enough time to
find their way to their preferred crystalline arrangement. So they
freeze into the disordered arrangement of a glass. It is known from
actual experience that, given a long enough time, the atoms in the
glassy solid will slowly rearrange themselves eventually into an
ordered crystalline arrangement. One possible crystalline form of
ordinary glass is quartz. The glass in some archaeological objects,
known to be many centuries old, is seen to have transformed to
quartz, which is the crystalline form of glass. This process is known
as devitrification. It appears therefore that the solid, given enough
time, always ends up being crystalline.
12
The first of these follows from the fact that the solid is stable: it does
not spontaneously change its shape. This means that the atoms
continue to stay where they are with respect to one another, and do
not move away from these positions: the net force on each atom is
zero. If I take a rod and pull on its two ends, it will get a bit longer; if I
push, it will get shorter. The actual change in size is perhaps invisible
to the naked eye, but it has been measured with special instruments.
In either case, the rod returns to its original shape when I let go. If I
describe this behaviour in terms of what the atoms are doing, I would
say that when the rod is stretched the atoms move a bit further apart,
and a bit closer together when it is compressed. An ordinary coil
spring becomes shorter when compressed, longer when stretched,
and in each case returns to its original length when released. I can
explain the behaviour of the rod by imagining that there are springs
connecting the atoms to one another. There are of course no real
springs in the solid; it is just that the influence of the atoms on one
another due to being made up of charged nuclei and electrons
produces the same effect, and we might talk about virtual springs
connecting the atoms to one another. This property of a solid, of
recovering its shape after being stretched or compressed, is called
its elasticity.
question But what about a copper wire? When I bend it, it remains
bent and does not return to its original shape. This must surely mean
that at least some of the atoms have moved from their original
positions.
answer The reason the wire remains bent is that the grains in the
wire are not perfect crystals, but have small imperfections which
permit some atoms to move about when the wire is bent. We shall
learn about these imperfections in chapter VII, and see how they can
explain the bending of a copper wire. These imperfections have very
little effect on many of the properties of solids that we shall be
Considering.
The second feature of the crystal comes from the fact that all atoms
of silver are identical. More generally, all atoms of any element are
identical, but different from atoms of any other element. Now
suppose that I place myself in the position of a particular atom in the
crystal, and look at the atoms around me. They will be distributed in
their preferred pattern at various distances in various directions
around me. Now I move to another atom in the crystal. Since all the
atoms are identical, the preferred pattern of atoms around this
second atom should be exactly the same as around the first atom.
Whatever factors led to the preferred arrangement around the first
atom should hold equally for the second atom and lead to the same
arrangement. The crystal, on an atomic scale, must look exactly the
same from the viewpoint of every atom in it. The universe around us
seems to have a similar property. When considered on a large scale
that includes enough stars, galaxies and so on, the universe looks
the same from one point in it as from any other point. Astronomers
call this the cosmologicalprinciple. The atoms in a crystal satisfy their
own version of the cosmological principle. There is nothing mystical
about this principle as it applies to solids. It is merely a consequence
of the experimentally established fact that all atoms of the same
element are identical.
II CRYSTALS
So we now have the two general characteristics of a crystal: the
atoms in it are as if connected to one another by virtual springs, and
their positions satisfy the cosmological principle. When we put atoms
together to meet these two requirements, we shall end up with the
ordered atomic arrangement in a crystal. We shall build the crystal in
three stages, for this makes it easy to satisfy the two requirements.
We shall first build a straight line of identical atoms, then assemble a
number of identical lines of atoms to give a plane of atoms. Finally
we shall stack identical planes of atoms above one another to give a
crystal, always making sure at each stage that the two requirements
are met.
3 The one-dimensional crystal
I now introduce an object that is very useful when trying to
understand nature, even though it does not exist in nature. This is a
one-dimensional crystal, consisting of a long row of atoms sitting
equally spaced in a straight line. A straight line has only length, and
zero width and zero thickness. Suppose that I measure distances
along the line from one of its ends. Then to tell you where any point
on the line is, I need to give you only its distance from that end. That
means that just one number, namely this distance, is enough to tell
you exactly where that point is. A ‘point’ in this context is a thing that
has a position in space, but has vanishingly small length, width and
thickness. Since we need only one number to tell us where a
particular point is on a straight line, we say that it is in one-
dimensional space. We assume that the atom is a point with a mass
equal to the mass of the atom, and that the influence of a
neighbouring atom on a given atom is as if a spring connects the
two.
-o o o-
531
B
FIGURE II-5. Three successive stages, @ q q ||_
atom 2 now has an atom on its left, so must atom 1 and at the same
distance. So I put down atom 3 to the left of 1 at a distance a from it.
Now atom 1 has an atom on each side, and therefore so must atoms
2 and 3, and this is satisfied by atoms 4 and 5 in stage B. Continued
application of the cosmological principle will lead to stage C and so
on, so that we get a long straight line of equally spaced atoms, with
neighbouring atoms connected by virtual springs of length a and no
net force on each atom.
We can now deduce some features of the virtual springs from the
properties of this one-dimensional crystal. We assume that it has its
undisturbed length, i.e. it can be stretched or compressed, but it will
return to its original length afterwards. This clearly means that the
undisturbed length of each spring is a, the spacing between the
atoms.
Since the virtual springs are just a way of expressing the influence of
one atom on another, and as the atoms are identical, all the springs
must also be identical in their strength. I mean by the strength of a
spring how hard it is to press or to stretch it.
question Will the length of the springs be the same for different kinds
of atoms, say silver and copper?
answer No. The length and the strength of the spring will be
determined by the detailed structure (in terms of nuclei and
electrons) of the two atoms it connects, because this is what
determines how one influences the other.
question The lattice must have two ends somewhere, and the atoms
near there will certainly not satisfy the cosmological principle. Is this
not a problem?
In any crystal, there are some atoms at its boundaries. If the crystal
is one-dimensional, they are at its two ends. They are at the edges
of a two-dimensional crystal, and on the surfaces of a three-
dimensional crystal. We shall assume in everything that follows that
we are looking at the properties in a region of a crystal far enough
away from its boundaries so that effects due to the atoms at those
boundaries can be neglected. Such properties are called bulk
properties, to distinguish them from surface properties which are
determined by the atoms at the boundaries.
Further, I make each of the two angles that I have marked on the
diagram equal to 90
degrees. We shall see later that these angles can also be different
from 90 degrees, and then one would get a different lattice.
All the atoms are identical and so one might expect the virtual
springs to have the same length.
answer Remember that the virtual springs stand for the mutual effect
of the electrons and nuclei of the neighbouring atoms on one
another. The electrons are distributed differently in different
directions around the 6 °
2
©
Q® @
P^fl R
90°
3 ‘-
up a two-dimensional lattice,
3 (shaded portion).
#a
lis
il
16
TWO-DIMENSIONAL LATTICES
You can see in the shaded part of fig. II-6 that each atom in lattice 1
has eight near neighbours around it, arranged in exactly the same
pattern. Thus the cosmological principle holds for the atoms in lattice
1, but not for the atoms in lattices 2 and 3 which have only five
nearest neighbours each. However, we know how to solve this
problem, from our experience in building the one-dimensional lattice.
We add pairs of lattices 4 and 5, 6 and 7 (the rest of fig. II-6), and so
on. The arrangement of atoms around each atom is now exactly the
same, and there is no net force on it; we have constructed a two-
dimensional crystal. We can completely specify the crystal by giving
the kind of atom (silver, copper, etc.), the lengths a and b, and the
angle of 90 degrees. We call a and b the lattice parameters. The
numbers, of atoms in each row and of rows, only specify the size of
the lattice, and have nothing to do with how the atoms are arranged.
We use the term lattice when we are interested only in the relative
positions of the atoms, but not in their kind. So the lattice is specified
by the lengths a and b being unequal, and the angle of 90 degrees.
What we have just finished constructing is a rectangular lattice. The
whole lattice can be broken up into a lot of rectangles, each with
edges a and b.
Is the rectangular lattice the only one I can have on a plane, just as
there was only one lattice I could construct on a straight line? No,
because I can change the angle and the ratio of a to b, which
determine the lattice. I denote the angle by the Greek letter cc
(alpha). I can now produce four more lattices, to give me a total of
five. Each of them shows order, but in a way which is different from
the others. I list below the five lattices with their names.
3. Square lattice, a and b axe. equal, and I use the same symbol a
for both of them.
II CRYSTALS
Figure II-7 shows the positions of the atoms in the basic unit for each
of these lattices.
This unit is called the unit cell of the corresponding lattice. It is the
smallest part of the o
o
,90°
(2)
■o
oblique.
o-
,90°
(3)
-o
-6
(5)
18
lattice which completely specifies what the whole lattice looks like. I
can build up the lattice by assembling the unit cells rather like tiles
on a floor. Each atom at the edge or the corner of a unit cell is then
shared with the immediately adjacent unit cell or cells.
answer These five are the only ones possible which satisfy the
cosmological principle.
question What if the added atom is not at the centre of each square?
answer The arrangement of atoms that results does not satisfy the
cosmological principle, because the surroundings of each added
atom are different from those of each atom in the original lattice.
You may amuse yourself by seeing what happens if you add an atom
to the centre of each little triangle in the triangular lattice. By drawing
a picture, you will find that the result is again a triangular lattice, but
with a smaller lattice parameter.
5 Order and symmetry
I have been talking about the order in a crystal or a lattice, implicitly
appealing to your innate sense of what the term means. However,
you might well say that order lies in the eyes of the beholder. To take
an example from the world of hearing rather than of sight, there are
those who find the music of Johann Sebastian Bach highly ordered,
and that of John Cage somewhat less so. But there are others who
detect much order in Cage’s music too. Such differences of
perception contribute to the richness of the world of art, but could
lead to undesirable confusion in the world of physics. We need to
make more precise what we mean by the order in a crystal. We do
so by introducing the idea of symmetry, an idea which plays a central
role in physics.
operations.
before
(b)
after
20
As a final example, consider the vase under the table, with a pair of
identical handles on opposite sides (fig. II-8). This vase has an axis
of rotational symmetry only for half of a full rotation, namely through
an angle of 180 degrees. A rotation through a further 180 degrees
brings the vase back to its starting position. Thus there are two
positions, each reached from the other by the same angle of rotation,
in which the vase looks exactly the same. I describe this property as
follows: A vase with a pair of identical handles on opposite sides has
a two-fold axis of rotational symmetry.
is not.
(a)
©
©6
(b)
.©■’
©■■’
© ©■’
©■’ ©
©•■’ p’
00
00
©” 0
AB
II CRYSTALS
symmetry. © • II ® It © @
©®©®©®©
wAB
©©®®©©©
®•©®©©©
© <§> © ® © ® @
question Are there any other axes of rotational symmetry for the
square lattice?
answer You can verify from the upper picture that if the axis passes
24
These lattices have one other type of symmetry. Consider the lattice
shown in part (b) of fig. 11-11. Draw a straight line through A and B,
think of it as a mirror, and imagine that the portions of the lattice lying
above and below the mirror are reflected by it. Then, for example,
the reflection of atom C will coincide with atom D, and vice versa.
The reflection of the portion of the lattice that is above the mirror will
look exactly like the portion below, and vice versa. The appearance
of the lattice is unchanged by this reflection. The line through AB is a
line of reflection symmetry.
answer A line parallel to the line joining A and B, and sitting midway
between B and C. Now if you imagine this and the other two lines
each rotated through an angle of 90 degrees, the resulting lines are
also lines of reflection symmetry.
question Which of the five lattices we have been looking at has the
least symmetry?
Let us pause and take stock of what we have done. I list the
highlights: 1. The lattice surrounding a lattice site looks exactly the
same as the lattice surrounding any other lattice site. This is the
cosmological principle as applied to a crystal, and is a consequence
of the fact that all the atoms of a given element are identical.
II CRYSTALS
I want now to say something about a more general matter. There are
some physicists who, when they have the time, think and even talk
or write about the beauty of physics.
You might recall the physics you were exposed to in your school
days, with its balls rolling down slopes, electric currents in wires
causing nearby compass needles to move, and so on.
Having got this far, you might be willing to grant that there is some
sort of beauty in these symmetries, but you might well wonder what
on earth one can do with them.
II CRYSTALS
(4) body-centred cubic; (5) face-centred cubic (only the three front
(1)
(2)
i,
1—
(3)
(4)
(5)
(6)
THREE-DIMENSIONAL LATTICES
distance cfrom the first, but with each atom now lying exactly over an
atom of the first lattice (there is a similarity to the way we made a
centred rectangular lattice in two dimensions). I show the resulting
unit cell in fig. II-13, labelled (2). This lattice is different from the
previous one, but still satisfies the cosmological principle. More
three-dimensional lattices can be assembled using each of the other
four two-dimensional lattices.
Even with this constraint, you might think that the possibilities are
practically limitless. There are however only fourteen lattices with
different symmetries that I can build satisfying the cosmological
principle. Figure 11-13 shows the unit cells and the names of some
of these lattices. The circles in these figures represent the lattice
points, and the lines joining them are there to guide the eye in seeing
the shapes of the unit cells.
4. Two lattices are different if they do not share exactly the same set
of symmetries.
29
II CRYSTALS
FIGURE n-14. Part (a) shows the
chloride).
(a)
chlorine atom
8 Crystal structures
The lattices we have been considering are not real objects, because
there are no atoms in them. They are just sets of points, distributed
in space in a way that satisfies the cosmological principle. It is only
after we associate atoms with the lattice points that we get a physical
object. A crystal of silver has the atoms sitting at the points of a face-
centred cubic lattice, whose unit cell is shown in fig. 11-13, labelled
(5). We say that silver crystallizes in the face-centred cubic structure.
Brass is an alloy in which 65 per cent of the 30
CRYSTAL STRUCTURES
atoms are copper and 3 5 per cent are zinc. A crystal of brass also
has a face-centred cubic structure, in which the copper and zinc
atoms occupy the lattice sites at random.
question In such an alloy there are two kinds of atoms, and so the
cosmological principle cannot be satisfied. What happens to the
lattice spacing?
Yet another kind of crystal is made up, not of single atoms sitting at
each lattice site, but rather of a group of two or more atoms
associated with each site. Further, the atoms in the group could all
be the same, or they could be different from one another. We could
also have molecules around each lattice site. I give below examples
of these different possibilities.
The element carbon exists in nature with two different crystalline
forms, as graphite and as diamond. Graphite consists of identical
planes, one of which is shown in part (a) of fig. 11-14. The picture
does not look at all like one of the five two-dimensional lattices.
The crystals in fig. II-15 do not look perfect: they have rounded
edges, and some are broken. This is because of the handling they
have received and their solubility in water.
Note particularly that while their sizes vary, the basic shape is the
same for all crystals of each of the two substances. This is a
beautiful manifestation of the underlying orderly arrangement of the
atoms or molecules in the crystal.
The simplest solid we can think of is just a line of atoms. If this line is
to satisfy the cosmological principle, the atoms will be equally
spaced along a straight line. This is the one-dimensional crystal,
which is discussed in section 3.
33
II CRYSTALS
Section 9 deals very briefly with a glassy solid, and points out that
after a long time of possibly centuries, it will have changed to a
crystalline form, which seems to be the preferred form of atoms in a
solid.
34
l Some preliminaries
We have, until now, thought of atoms as tiny hard spheres that feel a
force of attraction towards one another. We ignored the fact that an
atom consists of a nucleus surrounded by one or more electrons. If
we stopped at this, we would make little progress towards
understanding the properties of solids, other than that the atoms may
form crystals. We must therefore go on and take a look at the
internal structure of the atom itself, and see how this changes when
the atom is surrounded by other atoms in a crystal. We shall then
find that we need a description that is somewhat removed from our
everyday experience. There is a name for this description, quantum
mechanics-, mechanics, because it deals with the motion of objects,
and quantum, because features of the motion, such as energy, are
quantized, meaning that they can only have any one of a set of
distinct values, and no other value in between. Imagine for example
a car which is so constructed that it can travel only at speeds of 10,
or 50, or 100 km per hour, and not at any other speed. Then we
could say that the speed of the car is quantized. This idea of
quantization is something alien to our usual way of thinking. After all,
a car is able to go at any speed we like up to its maximum speed.
We shall find that there is no such freedom of choice for electrons
and nuclei. In this chapter I shall introduce you to the essentials of
quantum mechanics, and then in the next chapter show you the
description of an atom that it provides. I have summarized the
content of this chapter in section 13 at the end (page 61). You may
want to read it now, so that you have a preview of what lies ahead.
2 Motion
As I write this, I take a moment to look out of the window. I notice
that there is a breeze that is causing the leaves on the trees to
flutter. Birds are flying in the air. I see a stream flowing among the
trees, and there are small waves on its surface that are themselves
also in motion. Although I cannot see them, I know that the
molecules in the air around me are in a perpetual state of motion.
The paper clip on the table before me seems to be sitting there,
doing nothing. If I look at what the atoms in it are doing, using the
right sort of microscope, I find that they are all jiggling a little bit
about their regular crystalline positions. It would appear that the
world around us is always in various kinds of motion. At first glance,
there seems to be a daunting variety of such motions, and of moving
things. However, the essential features of each of these motions can
be 35
assigned to one or the other of just two types, namely the motion of
particles, or of waves.
PARTICLES
3 Particles
To begin with, a particle has mass, say M grams. If it is an electron,
then the mass is 9.11Xl028gram. Fora proton, it is 1.67xlO-24gram,
or about 1800 times the mass of the electron. The mass of a particle
is the measure of the amount of matter it contains.
Two particles moving with the same speed but in different directions
have different velocities, as also two particles moving in the same
direction but with different speeds.
In solids we are concerned not only with the motions of the atoms,
but also with how they influence one another. A simple example will
illustrate the point. In air, say, each molecule is whizzing about,
colliding every now and then with another molecule.
After each such collision, the two molecules have velocities that are
different from what they were before. The final velocities will depend
not only on the initial velocities of the two molecules, but also on
their masses. I would certainly notice the difference between a fly
and a cyclist colliding with me as I am walking along a path, even if
they both were travelling at the same velocity. So, in order to
consider collisions (or more generally interactions), between
particles, I need a concept, and a symbol for it, which combines
mass and velocity. I call it linear momentum, and give it the symbol p
(again a vector). Its direction is the direction of the velocity v, and its
magnitude p is given by the mass M multiplied by the speed v. In
short, we can write p = Mv, a vector equation,
and
Note that the second equation gives only the magnitude of the vector
p.
37
In the first case, we say that the body has kinetic energy, and in the
second case, potential energy. The kinetic energy is denoted by the
symbol K, and is given by KM
K~ 2
or equivalently
move when it is no longer held up by the table. You will note that
even when the ball was sitting still on the table, it had the potential to
move; hence the name potential energy. As another example,
consider an electric battery. I connect it to a light bulb with a pair of
wires and a switch. When I close the switch, the battery drives a
stream of electrons — we call it an electric current — through the
bulb causing it to light up. The electrons move because of the
battery. I say that the battery has chemical potential energy.
If I replace the electric battery with a solar battery and shine sunlight
on it, I get the same result. Thus sunlight also has energy that can
be converted to the motion of particles, electrons in this instance.
This energy is commonly called solar energy. A physicist would
probably call it electromagnetic potential energy, because light is
intimately connected with electricity and magnetism, as we shall see
later. Yet another example: if I stretch a rubber band between my
hands and let go, it flies off. The stretched band has elastic potential
energy when it is stretched, and this is converted to kinetic energy
when I let go the band. If it is clear from the context that I am talking
about some particular form of potential energy, then I may simply say
gravitational energy, chemical energy, and so on.
question Can you think of an example that shows that heat is also a
form of energy?
The foregoing does not exhaust all possible forms of energy, but is
meant to show that it comes in many different forms. The examples
also illustrate the fact that it is possible to convert the energy from
one form to another: the locomotive converts thermal energy to the
energy of motion. Now it often happens that one and the same
particle can have energy in more than one form at the same time.
For example, a particle sliding on a tabletop has both kinetic energy
and gravitational potential energy. In such cases, we add up all the
separate energies and call it the total energy E. We have E= kinetic
energy + gravitational potential energy + etc.
I have no problem doing this with the energies, because they are
scalars, just numbers. How do I do this with momenta, which are
vectors with both magnitude and 39
Ill PARTICLES AND WAVES
shown.
answer The canoe is going north in the water at 4 km/hour, and the
water itself is flowing east at 3 km/hour past the person on the river
bank.
The total energy, the toted linear momentum, and the total angular
momentum of a closed system —are constant at all times: they are
said to be conserved.
Let us see how the conservation principles apply to the case of a ball
falling off the edge of a table and coming to rest on the floor. At the
start of the fall, the ball has only gravitational potential energy.
During the fall, the potential energy is continually converted to kinetic
energy of motion, in such a way that the total energy remains
constant. Just before the instant when the ball hits the ground, its
initial potential energy has been completely converted to an equal
amount of kinetic energy, so that the total energy is still the same.
After impact, the ball is at rest on the floor, and both its kinetic and
potential energies are zero. Figure III-3 shows the three successive
stages. You may wonder what happened to energy conservation
when the ball came to rest on the floor, since both its kinetic and
potential energies then are zero. Not to worry: the impact makes a
sound and the ball warms up a bit, because the total energy has now
appeared partly as sound and the rest as heat. If you have ever
hammered a nail into the wall and noticed the nail warming up, you
will have seen the conversion of kinetic energy into thermal energy.
But how about the linear momentum? The ball started from rest with
zero momentum, and continually increased its speed as it fell. Its
momentum certainly did not remain constant. Then we note that the
closed system is not just the ball by itself, but the ball plus the earth
towards which it is falling, as seen by an extraterrestrial creature
somewhere out in space. This creature sees the earth move up to
meet the ball with a momentum which at each instant of time is
exactly equal to and opposite to the momentum of the ball, so that
the total momentum at all times is zero: it is conserved.
WAVES
the direction in which the wave is travelling, I can talk about the
velocity of the wave.
The distance between adjacent crests is always the same, and equal
to the distance between adjacent troughs. This distance is called the
wavelength, and denoted by the Greek letter X (lambda).
Figure III-4 illustrates some of these points. You can see that this
wave has a wavelength of 1 metre. The broken vertical line near the
left end of the picture intersects the wave at the position of the cork
at successive instants of time. It is on a crest at 0 second, has
moved down to a trough at 2 seconds, and is back again on the
following crest at 4
question Can you see that, if the speed of the wave is v, then v is
equal to the product of/and A?
answer I look at the cork, and find that crests are passing it each
second. The crests are spaced a distance A. apart. Therefore the
crest, which one second ago was at a distance times X, is now
passing the cork. So the speed of the wave is/times X centimetres
per second. I can write this as v —fX.
direction of wave
.oo
distance in metres
43
to get the speed. This speed depends on the nature of the wave and
the substance in which it travels. The speed of sound waves in air is
331 metres per second, in water 1480 metres per second. The
speed of light waves in free space is 3 X 10s metres per second, in
water 2.24 X 10s metres per second.
46
particles waves
localized in space and time not localized in space at any time mass,
M cannot imagine mass of a wave
momentum, p wavelength, X
energy, E frequency, /
no interference interference
no diffraction diffraction
7 The electron: particle and wave
We want to understand the properties of matter in terms of the
constituent atoms, which are themselves made up of electrons and
nuclei. To be specific, I talk about electrons here. I describe three
experiments that can tell me whether an electron behaves like a
particle or a wave. The characteristics of electrons that these
experiments will reveal to me are also those of protons and
neutrons, the other fundamental constituents of matter. Similar
experiments are commonly performed in many physics laboratories
all over the world.
In the first experiment, a beam of electrons that are going quite fast
passes into air containing water vapour. Each electron causes
condensation of water along its path and creates a string of tiny
droplets of water marking its trajectory. The gadget, called a cloud
chamber, has been widely used in the study of electrons, protons,
and other basic constituents of matter. A photograph taken with such
a chamber is shown in fig. III-8. The curly track at the top is from an
electron, curly because of the presence of a magnetic field exerting a
force on it. The other tracks are from other particles passing through
the chamber. Clearly the electron is localized, and its momentum
and energy can be measured in various ways. We therefore
conclude that electrons are particles.
47
In the third experiment, the electron waves coming off the nickel
crystal pass through a cloud chamber as in the first experiment. One
would expect them as waves to spread throughout the chamber, but
instead the electrons behave like particles again, making tracks in
the chamber just as in the first experiment. So it seems that,
depending on how they are observed, electrons behave like either
waves or particles. The very same electrons that behaved like waves
when they bounced off the crystal of nickel seem to turn into
particles as they go through the cloud chamber. This situation is
totally alien to everyday experience, and we need a new way of
thinking about it. This way exists, and it is called quantum
mechanics.
I should not leave you with the impression that this is all the
evidence there is for it.
the scattered X-rays. I find that I can describe the result exactly as if
the X-rays consisted of a stream of particles, whose energy and
momentum are related to the frequency and wavelength respectively
of the X-rays, and these particles bounce off the electrons keeping
the total momentum and energy constant. I conclude that these
waves, in this case, behave like particles. These particles of X-rays
are called photons.
and
I put this value of p in the formula for Kznd find that K~ 2m-In words:
the kinetic energy if is equal to the momentum p multiplied by itself
(which I call p-squared) and the result divided by twice the mass m.
Since the momentum itself is equal to the Planck constant h
multiplied by the wave number k, the kinetic energy is got by
multiplying /^-squared by ^-squared and dividing the result by twice
the mass. I think you will agree with me that formulas give us a very
convenient way of expressing what would take a lot of words to say!
shall find the electron there if I look for it’. You will note that this
probability refers to an actual effort, a set of experiments, to find out
where the electron is.
54
distance
If I now look to see where the electron is at any given instant of time,
one thing is certain: I shall not find it outside the box. If there is no
electron, there is no wave. So the amplitude of the electron-wave
outside the box is zero, and it remains zero right up to the two walls.
You must imagine that each wave is vibrating between the positions
shown by the continuous and broken lines somewhat like the
vibrations of a plucked string in a guitar.
There is an important point to note here: the wavelengths can only
have those values which permit integer numbers (1, 2, 3, …) of
troughs and crests to fit inside the box.
At this point, you may have noticed that the waves I have just
illustrated do not look quite like the waves that we encountered
earlier. Those earlier waves were travelling somewhere with some
velocity, while these waves are trapped inside the box, and are not
going anywhere. Such a trapped wave is called a standing wave, to
distinguish it from the other sort which is called a travelling wave. A
standing wave results from adding together two travelling waves of
the same frequency and wavelength, which are travelling in opposite
directions. This is illustrated in figs. Ill-11 and III-12. The two waves
can be thought of as arising from reflections at the ends P and Q_of
the box in fig. III-9. An everyday example of a standing wave is the
way the strings in a musical instrument like a piano or a guitar
vibrate.
of time 0, 1, and 2
distance
i ..-•‘V——O\
itude
Cu 7 X
CO V
>V
—0
—1
-■_•_■ 2
A :-
distance
h~ 2m ■
So we conclude that the energy of the electron can take only certain
discrete values, and none in between. In such a case, we say that
the energy is quantized. We refer to these values of the energy as
the energy levels of the electron in the box. This result is entirely due
to the wave-particle duality of the electron. If the electron were to be
considered purely as a particle in the box, then it could have any
desired value of the momentum and energy; there would be no
quantization. The smallest energy it could have would be zero, when
the particle is just sitting motionless somewhere in the box.
kl= JL
For the wave labelled (3), I see from the figure that there are \
wavelengths in the distance L, so that its wavelength is fl, and its
wave number k3 is question What is the wave number kt of the wave
labelled (4)?
h2k2
2m
57
h2(kn)2
tn~ 2m ■
En~ 2m •
h2n2
E”=8mL2
£l= 8mL2
question We said that the matter waves are also vibrating. What is
the frequency fn of the wave whose energy is £„?
answer Since the energy is given by the Planck constant h multiplied
by the frequency, the frequency fn equals En divided by h.
The lowest energy that the particle can have is called its zero-point
energy, and is a direct consequence of its wave nature. If it were
simply a particle, its lowest energy would be zero and it would be
sitting motionless. This however is not allowed because of its wave
nature: instead, it has a zero-point motion which results in its zero-
point energy.
We shall not do such calculations in this book, but shall look at some
of the results in later chapters, and see that they are what we would
have expected from the understanding of quantum mechanics that
we are now acquiring here.
distance
59
6o
SUMMARY
as you can verify from fig. Ill-10. If this formula looks opaque to you,
just skip it and carry on!
13 Summary
This chapter contains a number of new ideas and ways of looking at
things. I shall give you now a summary of the principal points in the
chapter, section by section. It may be helpful for you to look up the
relevant sections as you go through this summary.
I look in sections 1 and 2 at the world around us, and note that
everything is in a state of motion. I find that I can put the things that
are moving into one of two classes: they are either particles, or they
are waves. A particle is an idealization of our everyday sense of a
moving object: it has mass and occupies a tiny volume. We think that
electrons and nuclei might be good examples of particles, because
they have a mass and they are certainly very small. We shall find
later that this is not the whole story. The idea of a wave is easy to
grasp: we have all seen waves on the surface of water.
We now come to section 8, and find that nature has not yet finished
with its surprises for us. Light as well as other forms of wave motion
shows particle-like or wavelike properties depending on the specific
way in which they are observed. The particles of light are called
photons. Wave-particle duality thus turns out to be a property of all
nature.
62
IV THE ATOM
l Introduction
I should now like to make a small detour, and say a few words that
will give you an insight into some of the ways of physics. Having
been told that an electron has something called charge, you might
be tempted to ask, “Well, what exactly is it? Is it something that
makes the electron glitter, or colours it blue, or something like that?’
The answer is, none of the above. It is just a compact way of saying
that the electron has a property that makes it exert a force, which
has nothing to do with its mass, on another electron. The reasoning
goes as follows:
1. We observe that two electrons repel each other. This is a given
fact of nature.
IV THE ATOM
grams and so on, and I find the mass of a bag of potatoes, say, by
comparing the gravitational force on the bag with the force on these
weights. This last step is essentially what a spring balance does. So
with charge: I can make the charge on the electron my unit of
charge. A superb unit it is too, for it is exactly the same for all
electrons, and never changes with time. Then I can measure any
other charge by finding the electric force on it due to an electron. In
practice, such measurements are made by more indirect methods,
but the underlying principle is as I have described it.
From the discussion of energy in chapter III, I know that the two
electrons had potential energy when they were pinned a distance r
apart, and this energy is converted to kinetic energy when they are
released. Further experiments show me that the larger the distance r
at which the two electrons are pinned, the smaller the potential
energy that they have.
Now, I repeat all these experiments after replacing the two electrons
by two protons. I find that I get exactly the same results for the
potential and kinetic energies as before. Can I conclude that the
electron and the proton must have exactly the same charge? Not
yet! I do a third set of experiments, this time with one electron and
one proton. Almost everything is as with the first two sets of
experiments, but with one important difference: when set free, the
electron and proton move towards each other and not away from
each other. The force is attractive, and not repulsive as it was in the
two previous experiments. It thus appears that the charge on the
proton is similar to, yet essentially different from, the charge on the
electron. Similar because two protons and two electrons have the
same magnitude of force when they are the same distance apart.
Different because two protons (or two electrons) repel each other,
whereas an electron and a proton attract each other.
If you take two objects and stick them together, you get the mass of
the resulting object 64
answer Initially the atom had six protons in the nucleus with a charge
+6f (and also some neutrons with no charge) and six electrons with
total charge -6eand was therefore neutral. The ion has five electrons
with charge—5 c and so its net charge is +6e minus 5e, or just +e.
qiqi
V—
Therefore the potential energy of the electron and the proton held a
distance r apart is —e times +edivided by r, which is equal to —^.
You remember that a negative number multiplied by a positive
number gives a negative number.
The electric charge —e is the same for all electrons, everywhere and
always. The same is true for the charge +eof the proton. The charge
is an intrinsic and unchanging property of the particle, just like its
mass. Any electric charge that we might meet in nature, since it must
be made up of electrons and protons, can only have a value that is
an integer multiplied by e, and nothing in between. I can have a
charge of 5 times e, or 6 times e, but not, say, 5.5 times e. I can say
that electric charge is quantized, and the quantum of charge is e. I
should mention here that protons and neutrons are believed to be
made up of particles called quarks,swhich have positive or negative
charges of magnitudes \eox\e. This internal structure of protons and
neutrons has no effect on those properties of matter which are
discussed in this book, and can therefore be ignored.
Many of you are familiar with a picture of the atom that often appears
in the popular press, and on banners at demonstrations about
nuclear weapons, nuclear power stations, and so on. This picture
looks somewhat like fig. IV-1, with the dark circle in the centre
representing the nucleus, and electrons going around it in elliptical
orbits.
We have to find out what the wave function of the electron looks like
when it is near the proton. Then we can find out how the probability
varies from point to point and what the energy of the atom is. This
we now proceed to do.
0 0.5 1 1.5 2
So, instead of the popular picture of the electron going around the
proton in an orbit, we now have the quantum-mechanical picture of
the electron smeared out, so to speak, in a probability cloud
extending out to a distance of about an angstrom from the proton.
The energy of the proton with this electronic cloud is the ground
state energy of the atom. Using the Schroedinger equation, one
calculates this energy to be about -13.6 electron-volts. A word is in
order here about why the energy is negative. I have assumed that
when the electron and proton are very far apart, their energy is zero.
The negative sign then means that the energy of the ground state is
lower than the energy of the two particles when very far apart. If now
I want to break up the atom by taking the electron very far away from
the proton, I must feed an energy of 13.6 eV to the atom, because
+13.6 added to -13.6 gives me zero. This is an application of the
principle of conservation of energy.
I have introduced here a new unit with which to measure energy, the
electron-volt.
You are familiar with the fact that the electricity in your home is
supplied at about 230
volts, or at about 115 volts if the home is in north America. You can
think of volts as the pressure difference that drives the electrons to
flow in a current, for example through a light bulb. The electron-volt
is the energy that an electron acquires when it moves under a
pressure difference of one volt. The electron-volt, abbreviated to eV,
and the angstrom which I introduced earlier, are commonly used
when talking about things on the atomic scale, for the same reason
that we talk about distances between cities in miles or kilometres
rather than in inches or centimetres: they fit more comfortably the
scale of the things which are being measured.
69
IV THE ATOM
Let us look at fig. IV-2 again. Every point on the sphere is at exactly
the same distance r from the proton. In the ground state of the atom,
the wave function changes only when r changes, and therefore it
must have the same value at every point on the sphere. Now look at
the curve, labelled C, which represents a circle girdling the sphere:
its equator, so to speak. Suppose I start at some point on it, and go
all the way around the equator till I am back to the same point. Since
every point on the equator is at the same distance from the proton,
the wave function in this state must have the same value at all these
points. I have illustrated this in fig. IV-4. You are looking in this
picture at the proton with the equator around it. To help you get a
sense of depth, I have marked the front and back of the equator.
There is an object like a pin at each of several points on the equator.
The height of the pinhead at each point represents the amplitude of
the wave function, whose square is the probability of finding the
electron there. All these heights are the same, and you should
imagine that the probability is the same between the pins also. Again
to help you with the perspective, I have shown the pinheads in the
front as dark and in the back as clear. What we have here is a wave
of constant amplitude around the proton. It is certainly not at all like
the waves we see on water, or, for that matter, the waves associated
with the electron in a one-dimensional box (fig. III-10). It may help to
think of it as a wave with an extremely long wavelength which is
tending to infinity.
The probability in the ground state depends only on how far the
electron is from the proton, and not on which direction. So we
conclude that we find the same probability if we go along any line of
latitude or of longitude, so long as it is at the same distance r from
the proton.
0o
o
Q
back
proton
front
70
THE HYDROGEN ATOM: EXCITED STATES
proton
60
—■ o-
111
front
FIGURE IV-5. The wave function for an excited state in the hydrogen
atom, going round a circle of longitude with the proton at the centre.
Its amplitude is indicated by the height of the pinhead above the
circle. The pinheads in the front are shaded dark. This looks more
like what one expects from a wave: it has two oscillations in going
round the circle.
5 The hydrogen atom: excited states
We now come to the excited states, which are the states with
energies higher than the ground state energy. We find here that the
probability often depends not only on the distance of the electron
from the proton, but also on its direction. When I work out the
Schroedinger equation, I find that the resulting wave functions (and
consequently the probabilities) cannot take any arbitrary form, but
can only be one of a discrete set, each with its value for the energy:
the energy is quantized. These wave functions, unlike the one for the
ground state, have a recognizable wavelike character.
I show in fig. IV-5 how the wave function looks for one of these
states, if I follow it along a line of longitude. Since the wave must join
smoothly on to itself after going full-circle, its wavelength cannot be
any arbitrary value.
question Why must the wave join smoothly on to itself after going
full-circle?
answer Because the probability of finding the electron near any point
in space has a unique value in a given quantum state, and cannot
change abruptly from point to neighbouring point.
question How much energy must I feed into the atom in the ground
state to put it in its first excited state?
question What happens if the atom changes from the second excited
state (-1.5 eV) to the first excited state (-3.4eV)?
answer The atom loses an amount of energy equal to 3.4 minus 1.5
or 1.9 eV. By the conservation of energy, this energy has to appear
somewhere else: it could appear for example as a photon of energy
1.9 eV.
question Can I get the atom in the ground state to absorb a photon
of 10.0 eV?
answer No. If the atom were to absorb an energy of 10.0 eV, then its
final state should have an energy of-3.6 eV. No state with this energy
exists for the atom, which therefore cannot absorb it.
Let us now turn to the left half of fig. IV-6. We see three columns of
numbers, labelled at the top n, I, and E respectively. The numbers
under E (for energy) are the 72
each level.
n
3
0,1,2
0,1
-1.5eV
-3.4eV
y increa
Energ;
-13.6 eV
ionized atom
energy is positive
energy is negative
ground state
IV THE ATOM
We see from the figure that the ground state has quantum numbers
n equal to 1 and equal to zero. Its wave function contains these two
numbers, and I can use a convenient shorthand to write it as \|(« = 1,
/ = 0). This is not the end of the story. You remember that the
electron has not only a charge, it also acts like a tiny magnet that
can point in two opposite directions in the presence of a magnet, and
each of these orientations is a different state of the electron with its
own wave function. So let me introduce a new quantum number with
the symbol s for this property of the electron, and give it the value +\
for pointing one way and —\ for pointing the other Way. The reason
for choosing these peculiar numbers is that it then simplifies other
formulas which the physicist uses. Then we can write down the
corresponding wave functions as and
The energy is the same, —13.6 eV, for these two wavefunctions, and
is set by the value of n alone. Now if you are getting a bit alarmed by
the complicated symbol for the wave function I have just used, I ask
you to think of it just as a shorthand way of reminding yourself that
the wave function for the electron has the values of the quantum
numbers shown inside the parentheses.
Thus we find that there are two different wave functions for the
ground state with exactly the same energy. You may think this a little
peculiar; it would be quite a coincidence to have two wave functions
that look different and yet have exactly the same energy. Consider
this example: a square that is four centimetres on the side looks
quite different from a rectangle that is sixteen centimetres by one
centimetre, but they have exactly the same area. So it is not difficult
to imagine that there can be two, or even more, things that are
different from one another in certain ways and yet have one property
that is exactly the same for all of them. Physicists have a name for
this occurrence of two or more wave functions having the same
energy. They say that the energy level is degenerate. I have a friend,
not a physicist, who giggles each time I use the term in a physics
context. She says it conjures up for her an image of a person with
bloodshot eyes and ragged hair: a degenerate. This is one more
example of the special meaning in physics that we attach to many
words of common usage.
You recall that for the electron in a box we had a picture of what the
wave function looked like: it looked like a wave. Things get more
complicated with atoms and even more so when we get to crystals,
so that it will not be always possible to picture their wave functions.
So we shall sometimes specify a state of the system by saying what
values the quantum numbers have in that state, and not worry about
what the wave func-74
tion looks like. You will now see that this is what I did above when I
wrote down the wave functions for the ground state of the atom.
n= 1,2,3, … ,
/ =0, 1,2, … ,
and s = +1 -i
We have seen that the value of n in the wave function tells me the
energy level to which it belongs: 1 for the ground state, 2 for the first
excited state, and so on. The two possible values of the quantum
number s distinguish between the two opposite orientations of the
electron-magnet. Now what about the quantum number U It contains
information about another property of the atom, namely its angular
momentum. This property, you remember from chapter III, is
connected with rotational motion. Furthermore, the quantum number
/ contains hidden in it another quantum number that I shall call m. I
explain all this by looking now at the first excited state.
This first excited state, as I see from fig. IV-6, has n= 2, while /can be
either 0 or 1
__
_I
—2
_i
Each row in the table is a set of possible quantum numbers for the
first excited state.
Each set has its wave function, and no two sets are identical. We
see that there are eight such wave functions, each of them having
exactly the same energy of—3.4 eV. This state is even more
degenerate than the ground state, which had only two wave
functions.
You may have begun to see a pattern emerging about how the
quantum numbers n, I and m are interrelated. It is convenient to
express this pattern using algebraic symbols, and this I do now. For
a given value of the quantum number n, lean take any of the integer
values between 0 and n-l:
75
2
2
+1
+1
0
0
-1
-1
IV THE ATOM
With each value of/, m takes any of the integer values from +/ to —/:
m=+l, +1-1,…,-/+1,-1.
question How many different wave functions are there for the excited
state with n — 3? And for n — 4?
/ = 0 and m - 0;
When n — 4, can also take the value 3, besides the above values.
You can use the same method as before to work out that there are
14 wave functions with = 3, so that there will be 14+ 18 or 32
different wave functions for n = 4.
I said that the quantum number is related to the atom’s angular
momentum, and we have seen that m is very closely connected with
. If I did an experiment to measure the angular momentum of the
atom along some particular direction, I would find that it can only
have one of the values given by m multiplied by h and divided by
twice 7t, or more compactly expressed,
mh
271
The electron not only has an electric charge — e and behaves like a
tiny magnet, but also has an angular momentum as though it were
spinning like a top. When I measure this angular momentum, I find
that it is also quantized: it can only have one of the two values
2tc
The answer is that the atom will be in its ground state. This is so
because, if it begins in an excited state, any slight disturbance will
cause it to emit a photon (or in some other way lose energy to the
surroundings) and change to the ground state. The energy so lost by
the atom will be exactly equal to the difference in energies of the two
states, because the total energy must be conserved. The ground
state is the stable state of the atom, because it has no state of lower
energy to which it can go.
The next simplest atom must have two electrons with total charge
-2e, and therefore a nucleus with charge +2e, because the atom as
a whole is neutral with zero total charge.
Its energy level picture will be rather different from the one in fig. IV-
6, because now we have to think of the quantum numbers for two
electrons. The energies of the levels will be different from what they
were in the hydrogen atom, because the charge on the nucleus is +2
e instead of+e. What we have now is an atom of helium, which you
know as the gas used to fill balloons on fairgrounds and for
children’s birthday parties. The two electrons go into the ground
state, and their quantum numbers are as shown below: First
electron: n— \,l—Q,s — —\
Second electron: n= 1, /= 0, s= +2
You will note that not all the quantum numbers for the two electrons
are the same. One other point: the nucleus of the helium atom
consists of two protons each with a charge +eand two neutrons with
no charge.
The description of the helium atom which I have given raises two
questions whose answers introduce us to some very important new
ideas. The first question is, since I can balance the charge of—2eof
the two electrons with just two protons each with a charge +e, why
do I want to throw in two neutrons also which, after all, have no
charge? Well, you remember that two like charges repel each other,
so that if I had just two protons in the nucleus, they would fly apart. It
turns out that there is another kind offeree that acts only among
neutrons and protons, and this force attracts them to one another
very strongly. We call this the nuclear force. This attractive nuclear
force, however, is still not strong enough to overcome the repulsive
electric force that is trying to push the protons apart. The two
neutrons increase the attractive nuclear force further, without
affecting the electric force (because they are electrically neutral),
enough to hold the nucleus together. A study of the nuclear force is
at the heart of nuclear physics, which is 77
IV THE ATOM
one of the branches of physics. We shall not dwell further on it, since
what goes on inside the nucleus is not relevant to those properties of
matter that we want to understand in this book.
And now for the second question: When I put the two electrons in the
ground state, why did I not put them both in a state with the same
value for the s quantum number instead of putting one of them in the
s = -2 state and the other in the s = +2 state? The answer lies in a
peculiar property of electrons that I have not mentioned so far. This
is that no two electrons can have all of their quantum numbers
exactly the same. A given state, with certain values for its quantum
numbers, can be occupied by one, and only one, electron. It is as if
an electron, sitting in a given state, excludes any other electron from
getting into the same state. This so-called exclusion principle was
discovered by the physicist Wolfgang Pauli, and is named for him. I
shall say more about it in chapter V. We shall use it now to form a
picture of the wave functions for electrons in a given atom.
I hope you are beginning to see now the picture of an atom given by
quantum mechanics. Let me describe it, putting together the points
made in this chapter. All the atoms of each chemical element are
identical. Each atom has a nucleus with a certain number Z of
protons and a number A/of neutrons. I show the values of Z and
JVfor some elements in table IV-2. Surrounding the nucleus are Z
electrons, so that the atom is electrically neutral; the charge +Ze on
the nucleus is exactly cancelled by the charge ~Ze of the Z
electrons. The electrons are to be found in the successive energy
levels, starting at the ground level. According to the Pauli exclusion
principle, no two electrons will have all the quantum numbers the
same. The value of Z, the number of protons in the nucleus (which is
the same as the number of electrons in the atom), identifies the
element uniquely; each element has a different value of Z.
0
2
146
hydrogen
helium
carbon
uranium
92
78
If you look at fig. IV-3 again, you will be reminded that there is very
little probability of finding the electron at a distance of more than
about an angstrom from the proton in the hydrogen atom. We can
say that the size of that atom is about one angstrom. The electrons
in the other atoms go into successively higher excited states.
I remind you again that no two of these electrons have all of their
quantum numbers the same. You must imagine that there exists a
wave function that describes the probability of finding any one of
these electrons at any specified region around the nucleus, and is
such that the Pauli exclusion principle is obeyed. I am afraid I cannot
draw a picture of this wave function for, say, a carbon atom in the
way I was able to for the much simpler case of a single electron
trapped in a box in chapter III. This is where mathematics comes to
the aid of the physicist. She will say that she does indeed have a
picture of the atom, but it is expressed in mathematical formulas and
not in lines and shadings on a sheet of paper. We shall see as we go
along that we can make much headway even without all that
mathematics.
All the things we see around us are made up of atoms. The only
difference between atoms of different elements is in the number of
electrons around the nucleus and the numbers of protons and
neutrons in the nucleus. This might appear to be too trivial a
difference to account for the amazing variety of properties that things
around us have.
The picture of an atom that we have arrived at is not at all what one
might have thought, namely a tiny hard ball a few angstroms across
in size. We can make this dramatically evident if we imagine that we
have somehow magnified an atom so that the nucleus measures a
centimetre across. Then the electrons, numbering at most about a
hundred, will be distributed in a sphere of diameter about one
kilometre. The cloud of electrons around the nucleus is a very thin
cloud indeed; most of the atom looks like empty space. It would
appear that one atom should be able to pass right through a second
atom and keep going. This is not what happens in reality. When I put
atoms together to form a solid, they behave as if they were hard
impenetrable balls stacked together. To take a more immediate
example, I am sitting comfortably on a chair as I write this; my
atoms, and therefore I, do not just simply pass through the atoms in
the chair. Why is this so?
79
IV THE ATOM
The answer lies in the Pauli exclusion principle, which says that one,
and only one, electron can occupy a given state with a given set of
quantum numbers. A second electron must necessarily go to another
state with a different set of quantum numbers. Now let us see what
happens when I try to bring two atoms of carbon, say, together.
Figure IV-7 shows the two atoms some distance apart. You can think
of it as a kind of snapshot. You see in each atom the nucleus at the
centre and six electrons in the surrounding grey space, which
represents the sphere within which each of the electrons is likely to
be found. Please note that this and the following picture are not
drawn to scale. Each electron is in one of the six possible states of
lowest energy. Now suppose I try, as in fig.
IV-8, to bring the two atoms together so that the two grey spheres
overlap. Then in the overlapping region I am forced to have pairs of
electrons, one originally from each atom, trying to occupy the same
state. This, we know, is prohibited by the Pauli exclusion principle,
which allows only one electron to be in each state. The consequence
is that the atoms cannot penetrate into each other, and my chair
supports me and does not let me pass right through it.
I have said that the number of electrons in an atom uniquely
specifies which element it is: one for hydrogen, six for carbon, and
so on. This is also the number of protons in the nucleus. There are
also some neutrons in the nucleus, which have no electric charge.
So their number is not fixed, but can vary a bit without affecting the
electrical FIGURE IV-7. Two carbon atoms A
Am
80
SUMMARY
IV THE ATOM
82
SUMMARY
Before getting there, there is one other matter we have to deal with.
We have seen so far what the quantum-mechanical picture of single
atoms looks like. In an actual solid/there are very large numbers of
atoms packed rather closely together. Very large numbers indeed: a
cube of solid measuring one centimetre along the edge will have
something like 1021 atoms in it. That means that there will also be
comparably large numbers of electrons in the solid. It is obviously a
hopeless task to try to find out what each electron is doing, because
even the lifetime of the universe would not be enough for the task.
We need to invent a new way of dealing with such large numbers of
objects. This way is called statistical physics, and it is the subject of
the next chapter.
STATISTICAL PHYSICS
l Introduction
Any object that we see around us, even though it may be very small
like a grain of sand or a pin, consists of a very large number of
atoms, about 1021. When the atoms are relatively far apart, as in a
gas, the electronic structure, meaning the wave functions, of each
atom is practically unaffected by the presence of the other atoms.
When the atoms come close together in a solid, the wave functions
become different from what they were for the isolated atoms. In
crystals of metallic elements like silver and copper this change is
such that the electrons in the highest energy state of the isolated
atom are no longer bound to stay near the corresponding nucleus,
but are free to wander around in the whole crystal. This freeing of the
outer electrons happens with each atom in the crystal. So we end up
with a crystal in which a large number of electrons, 1021 or more,
are wandering about. These electrons are responsible for many of
the metallic properties of copper or silver: the shiny appearance and
the ability to carry electric current and to conduct heat easily, for
example. We shall see later how this comes about. For the present
we need to find a way to describe the properties and behaviour of
such large numbers of identical entities like electrons or atoms. We
do this with statistics.
Figure V-l shows some information about the age of the populations
of two countries: the U.S.A. in the year 1989 and India in 1990.
Along the horizontal line at the bottom you see the labels 0 to 9, 10
to 19, and so on, which stand for ages up to 9 years, between 10
and 19 years, …. Above each of these labels are two bars, one for
the U.S.A. and the other for India. The length of each bar represents
the percentage of the population that is in the corresponding age
group. The vertical line on the left of the 84
INTRODUCTION
FIGURE v-l. The population
On
On
r-H
On
(N
-i—»
o
On
CO
-4—’
co
On
-t—»
On
in
in
On
^O
4—*
You will note that this figure, although it tells us something about the
ages of the people, is quite different from a simple, if very long, list of
the age of each person in the two populations. It tells us how the
people are distributed in different age groups. For this reason we say
that fig. V-l shows the distribution of the populations of the U.S.A.
question What are the percentages for the age group from 0 to 9
years in the U.S.A. and India?
answer It is about 15 per cent in the U.S.A, and 25 per cent in India.
There were 248 million people in the U.S.A. and 827 million in India
when these counts were made. You can imagine the impossibility of
trying to make a list showing the age of each person. It suffices for
many purposes just to know what fractions of the population are up
to 9 years old, 10 to 19 years old, and so on. Such information would
be needed in making plans for education, health care, social
security, and other such matters. The picture of the distribution offers
something more: it presents in a visually striking manner something
about the populations that would be invisible in a bare table showing
the age of each person, and not easy to see even in table V-1, which
shows the numbers that were used in drawing the figure. I point out
the following features of the way the percentage of people in a given
age group varies as the age of the group increases:
V STATISTICAL PHYSICS
Table V-1. Distribution by age groups of populations of the U.S.A.
and India age group
per cent
(U.S.A.)
per cent
(India)
14.9
24.8
14.0
21.8
16.3
17.7
16.8
12.7
12.2
9.5
8.9
7.0
8.5
4.2
1. For the U.S.A., this percentage slowly increases to a maximum
value at the 30 to 39 year age group, and then begins to decrease.
2. For India, this percentage starts at a relatively high level for the
youngest age group, and then steadily declines as the age
increases. The pattern is strikingly different from that of the U.S.A.
One can make sense of these two features if one thinks about the
relative affluence of the two countries and the consequence for infant
mortality, health care, and nutrition, and also recalls that for some
years in the 1950s, the so-called baby-boom years, the U.S.A. had
an unusually high birth rate. I shall not dwell further on this, for this
book is about physics and not demographics. I do note three
features of this way of presenting the information. The first is that we
know something about the distribution of the population among
different age groups, but nothing about the age of any given
individual. The second is that the pattern of the age distribution
implicitly says something about the population: its affluence, state of
health, and even history. Thirdly, the information in this age
distribution can be used for other purposes such as the planning of
health care, social services, and so on. There are analogies to these
three features in the way we describe a collection of a large number
of particles in physics. This description is called statistical physics.
Before I get to that, I must explain to you what temperature means in
terms of the atomic structure of matter. As you will see soon,
temperature plays a crucial role in statistical physics.
2 Temperature
Boiling water is hot, and ice is cold. I specify how hot or cold
something is by giving its temperature. Water boils in New York at a
temperature of 212 degrees Fahrenheit, and freezes to ice at 32
degrees Fahrenheit. In Rome, on the other hand, water boils at 100
TEMPERATURE
shall write ‘twenty degrees celsius’ as ‘20 °C\ There is yet another
way of expressing temperature, by adding the number 273 to the
degrees Celsius and calling the result as so many kelvin. For
example, 20 °C would be the same as 293 kelvin, which we shall
write as 293 K. You will see in a while that there is some point to
expressing temperatures this way.
We get our idea of something being hot or cold through our sense of
touch. I now want to connect this idea with what the atoms are doing
in a piece of matter. Suppose I take a container filled with helium
gas, say, which is at a temperature of 20 °C. The helium atoms are
not just standing still, but are whizzing about in all directions at all
possible speeds, bumping into one another and with the walls of the
container. When two atoms collide, the speed of each will in general
be different from what it was before. You can see that it is hopeless
to expect to follow the motion of each atom as time goes on.
Fortunately, we do not need to do this. Numerous experiments have
shown that the number of atoms whose speeds lie in any given
range will remain unchanged, so long as the temperature remains
unchanged. The atoms whose speeds change to outside this range
because of collisions are replaced by an equal number of atoms
whose speeds move into this range because of collisions. Now
suppose that I heat the container so that it and the gas inside it are
at a temperature of 300 °C, say.
The atoms will still be moving about at all speeds, but their
distribution in different ranges of speeds will be changed by this
change of temperature.
o
00
(N
00
CNJ
(N
00
(N
cn
00
CO
CM
^o
rt
00
OO^Ht-hcSICSCSCOCO
speed, km/second
V STATISTICAL PHYSICS
The horizontal axis carries labels showing speeds within the ranges
0 to 0.4, 0.4 to 0.8, … kilometres per second. I should mention that a
speed of 1 km/sec, which is the same as 2250 miles per hour, is
somewhat more than what we normally meet; the atoms are indeed
speeding about. Above each of these numbers are two bars, one for
each of the two temperatures. The length of a bar, measured on the
vertical scale at the left, gives the percentage of atoms that have
speeds in the corresponding range and at that temperature. For
example, I see from the figure that about 30 per cent of the atoms in
the gas at 20 °C have speeds between 1.2 and 1.6km/s, and this
drops to 20 per cent when the temperature is raised to 300 °C.
answer The question is similar to ‘If four apples cost one dollar each,
and six other apples cost two dollars each, what is the average cost
of an apple?’ To answer this, we work out the total cost of the ten
apples as four times one, plus six times two, or 16 dollars, so that
the average cost per apple is 16 divided by 10 or 1.60 dollars. We
can work out the average speed in the same way, substituting atoms
for apples and speed for cost.
Suppose we have 100 atoms in the gas. The actual number will not
matter, because we are working only with percentages. We find from
the figure that about 27 per cent, or 27 of them, have speeds
between 0.8 and 1.2 km/s. We say approximately that each has a
speed of 1.0 km/s, and we multiply 27 by 1.0 to give us the number
27.0. We do the same thing with all the speed ranges, add up the
resulting numbers, and divide by 100, which is the total number of
atoms, to get the average speed per atom. When you work it out,
you will find that the average speed is 1.2 km/s. We could have
expected the answer to be about this, because we see from the
figure that a large fraction of the atoms have speeds around 1.2
km/s. If more than half the apples I bought cost a dollar each, and
there was not too much spread in the prices of the rest, then I would
expect the average cost per apple to be about a dollar.
there are atoms moving at all speeds from practically zero to more
than 4.4 km/s, the percentage reaches a maximum value for speeds
in the range between 1.2 and 1.6 km/s, and drops off at lower as well
as higher speeds. There are some atoms moving at speeds much
greater than 4.4 km/s, but they are such a small percentage of the
total number of atoms that they do not show up on the scale of this
picture.
You can see that the distribution, although it has the same qualitative
shape, is quantitatively different. Raising the temperature leaves a
smaller fraction of atoms at the lower speeds and increases the
fraction of atoms at higher speeds. An appreciable percentage of
atoms is now moving at high speeds at which there were practically
no atoms at the lower temperature. The range of speeds at which
the maximum percentage of atoms occurs is now at 1.6 to 2.0 km/s,
while it was at 1.2 to 1.6 km/s at 20 °C. We can calculate the
average speed of an atom for the distribution at 300 °C, and find that
it is 1.7 km/s, distinctly higher than the value of 1.2 km/s at 20 °C. It
would appear that the average speed is a possible candidate for the
single number we are looking for to relate to the temperature. It turns
out that there is an even better candidate, namely the average
energy of an atom in the gas.
3 Temperature and energy
When the container of helium gas was heated from 20 °C to 300 °C,
we see from fig.
V-2 that the only changes were that there were fewer slower atoms
and more faster atoms, and the range of speeds in which the highest
percentage of atoms found themselves was higher. As a
consequence, the average speed increased from 1.2 km/s at 293 K
to 1.7 km/s at 573 K. Notice that I have now changed to kelvin from
degrees Celsius, and for good reason too, as you will soon see.
From the foregoing observations it seems reasonable to conclude
that the temperature must somehow be connected with the state of
motion of the atoms, which was the only thing that changed when
the gas was heated up. So let us try to find out what this connection
might be. To do this, we need to recall some things that we learnt
about the motion of particles in chapter III, section 3 (page 37).
We learnt there that a moving atom with mass M and velocity v has a
momentum p and kinetic energy if which are given by the formulas
p = Mv
and
„ _ Mv2
K — -,
Things are different with the kinetic energy of the gas. We see from
the formula for K that increasing the speed increases K even faster:
doubling the speed quadruples the kinetic energy. So the raising of
the temperature, resulting in a speeding up of the atoms, causes the
total energy (which we shall call E) of the gas to increase. It is
tempting to say that the temperature is somehow related to the total
energy of the gas. But not so fast! If I took two identical containers of
the gas at the same temperature and put them together, the resulting
object would have twice the energy of the separate containers, but
the temperature would be unchanged.
and
The letter M stands for the mass of the helium atom expressed in
grams. The average speed and the average energy depend only on
the temperature of the container of gas, and not on how many atoms
there are.
The average speed was 1.2 km/s at 293 K and 1.7 km/s at 593 K.
The speed does increase with an increase in temperature, but not
exactly in proportion. A doubling of the temperature causes a much
smaller increase of the average speed. However, when the
temperature is almost doubled from 293 K to 593 K, the average
energy is also doubled. There exists a simple proportionality
between temperature and average energy. You can see now why I
chose to measure the temperature in kelvin. It is only then that this
simple proportionality appears. We have verified this proportionality
for helium gas at two temperatures now. Much work with many
gases at many temperatures has confirmed this proportionality of the
temperature to the average energy of the atoms or molecules in the
gas.
The factor \ in the formula for K is there because of the value chosen
for the constant kB. We see that doubling the temperature doubles
the kinetic energy. If we go in the other direction and halve the
temperature, then the kinetic energy is also halved and the speeds
of the atoms are lower. As we keep reducing the temperature, K gets
smaller and smaller. Is there going to be an end to this? Yes,
because the lowest energy an atom can have is its zero-point
energy, which is a consequence of the wave aspect of the atom
inside a container, just as the particle in a one-dimensional box
(chapter III) has a finite zero-point energy. This must mean, from the
formula above, that the energy of motion (called the thermal energy)
which is to be attributed to temperature is zero at a temperature of
zero K, and the atoms are left with just their zero-point energy. We
shall call this the absolute zero of temperature, 0 K. From the way
we defined the Kelvin temperature scale (sometimes also called the
absolute temperature scale), we see that the absolute zero is at -273
°C. This is the lowest possible temperature that could exist anywhere
at all, corresponding to all the atoms being in their lowest energy
state.
5 The temperature of crystals
In the preceding section, I talked glibly about cooling the gas until it
reaches the absolute zero, where the atoms have no thermal energy.
Well, this does not quite agree with reality.
If I take a container of some gas and cool it to lower and yet lower
temperatures, then at some temperature (called the boiling point) it
will condense into a liquid. On further cooling, it reaches a
temperature (the freezing point) at which it becomes a solid. An
example: I V STATISTICAL PHYSICS
The inside of a hot oven is filled not only with hot air but also with
infrared radiation which is after all a form of light and is composed of
photons. It is as if the oven is filled with a gas of photons, all moving
at the speed of light and with an energy distribution that will
correspond to the temperature of the oven. We shall see presently
that this distribution is different from the one for atoms of helium gas.
E=hf.
Now suppose I take two particles labelled A and B that are the same
kind: they both are electrons, or hydrogen atoms, for example. The
wave function of this system of two particles will certainly be different
from that of one particle. It will depend on their 94
answer Alas, not so easily. For one thing, we do not have enough
information about the particles, such as whether they are in a box
and if so what size and shape of box, etc. We do not even know if
both of them are electrons or protons or photons or whatever. If I had
all this needed information, I could in principle calculate the wave
function <|). I would then still have difficulty in making a picture of it,
for its value will depend on the positions of each of the two particles,
and there is no simple way I could show all of this on a flat sheet of
paper.
There are two ways of satisfying this condition. One is if <t>(?A. 9b)
= <K?b, 1a)>
and the other (remembering that a negative number multiplied by
itself gives a positive number) is if
We shall be interested in systems not with just two particles, but with
a very large number JV, typically 1021, of particles. The wave
function for such a system would have the form §(qlt q2, …, gw),
where each q stands for the position and quantum numbers of one of
the particles. Using the same reasoning as I did for two particles, I
see that interchanging the two particles at q1 and q2 and leaving the
others where they were will give me either the same wave function
or merely change its sign: either 95
V STATISTICAL PHYSICS
or
You will note that in all the foregoing, I have said nothing about the
actual form of the wave function, other than that it describes a
system of Af identical particles. The actual form will depend, for
example, on whether the system is the collection of 92 electrons in
an atom of uranium or the 1021 or so atoms in a container of helium.
In any case, the wave function must be one or the other of the two
kinds described above. What the particles are - electrons, photons,
helium atoms, etc. - determines which of the two kinds is
appropriate.
96
Table V-2.
electron
nI
10
32
number
m
Now consider the special case where the two sets of quantum
numbers for the two fermions A and B are identical. I denote this
single set by the symbol a. Then the wave function depends on (X
and the position vectors rA and rB of A and B, and can be written as
\|/(rA, rB, a). Interchanging A and B leads to the result \|/(rB, rA, a) =
-\|/(rA,rB, a).
V STATISTICAL PHYSICS
where kn is the magnitude of the vector kn. The wave function for
each of these wave vectors will also depend on the spin quantum
number s, and I can write it as \|/(kn, s). If for example the particles
are electrons, then for a given kn there are two wave functions,
vj/(kn,+2) and \y(kn,-2) for the two possible values of the spin
quantum number for the same value of kK. How the particles are
distributed among these wave functions depends on whether they
are fermions or bosons, as we shall see in the next section.
9 Fermion and boson distributions
I shall first consider the case of fermions. I imagine a large number,
like 1021, of fermions in the box, and see how they are distributed
among the different states 98
\|/(kn, s). I assume to begin with that the box is at the absolute zero
of temperature, 0 K, and then see what happens as I heat the box
and its contents to higher temperatures.
by an assembly of fermions, at
states behind.
M.
(a)
(c)
(d)
99
V STATISTICAL PHYSICS
question Does this mean that no two fermions can have the same
energy?
answer No. The Pauli principle says that no two fermions can be in
the same quantum state. However, quantum states can have
degeneracy, meaning that more than one such state can have the
same energy. Thus many fermions can have the same energy even
though each is in a different quantum state. In a cubic box, for
example, there are twelve degenerate states all of the same lowest
possible energy.
Now suppose I heat the box and its contents to some temperature
above the absolute zero. This means that I add some energy to the
fermions, so that they move to nearby states of higher energy. Note
that fermions sitting in states well below the Fermi energy cannot go
to states of somewhat higher energy, because these states are
already fully occupied and will not accept any more fermions, thanks
to the Pauli principle.
Only those fermions that are at or near the Fermi energy can find
empty states at higher energies into which they can move. Fermions
have room only at the top. This state of affairs is illustrated in fig. V-
4, stacks labelled (b), (c), and (d) for successively higher
temperatures. A change in the temperature affects only those
fermions with energies near the Fermi energy, which form only a
very small fraction of the total number, and leaves the vast majority
of the fermions in the same states they occupied at the absolute
zero. Electrons are fermions, and many of the properties of solids
that we shall be looking at involve changing the energy of the
electron system. Thus these properties are mainly determined by
just the relatively few electrons with energies near the Fermi energy,
while the rest of them are, so to speak, passive spectators.
The situation is quite different with bosons, as illustrated in fig. V-5.
The part labelled (a) shows the case for 0 K, and (b), (c) and (d) for
successively higher temperatures. At each temperature, the box
labelled (1) represents the state of lowest energy, and (2), (3) and (4)
are three higher energy states. Being bosons, the particles can all
get into the lowest energy state, and this is exactly what happens at
0 K. With increasing 100
SUMMARY
(d)
(c)
(b)
(a)
»•#••••••••®m»»»Bmm®mm
•••••••
••••••••••••••••a**
(4)
(3)
(2)
CD
(4)
(3)
(2)
(1)
(4)
(3)
(2)
(1)
(4)
(3)
(2)
(1)
temperature, more and more of them acquire thermal energy and get
into the higher energy states, as shown in parts (b), (c), and (d) of
the picture. You can see that the resulting distributions are totally
different from those for fermions. It is therefore to be expected that
the properties of boson systems will be quite different from the
properties of fermion systems.
10 Summary
I introduced in section 1 the idea of a statistical distribution as a
convenient way of handling the individual characteristics of the
members of a very large assembly. I illustrated it with the example of
the distribution by age of the population of a country.
V STATISTICAL PHYSICS
IO2
I remind you that the typical size of an atom is a few angstroms (one
angstrom =
103
FROM ENERGY LEVELS TO ENERGY BANDS
+47e, is very small compared to the size of the atom. The differently
shaded regions show where electrons with different quantum
numbers n are most likely to be, and the number of these electrons
is shown for each region. Thus there are two electrons with n = 1
very close to the nucleus, and one electron with n = 5 near the
surface of the atom, and the remaining electrons are distributed in
between.
A
toO
c3
S-H
105
I show the resulting picture of the energy levels in fig. VI-2 for the
quantum number n equal to 1, 2, and 3. For each of these values of
n, there is a certain number of levels that are clustered together, and
there are gaps in energy between adjacent clusters. Each level
corresponds to a different wave function with its set of quantum
numbers n, I, m and 5. We know from the Pauli principle that there
can be at most one electron described by each of these wave
functions with the corresponding energy.
atoms. This does not violate the Pauli exclusion principle, because
each of these electrons has a wave function belonging to a particular
atom with vanishing probability of being near any other atom. This is
equivalent to saying that the wave functions of the different atoms do
not overlap.
The atoms are now gradually brought closer and closer together,
until they finally end up in the positions they have in the actual
crystal. The atoms are now crowded close together, and we can no
longer neglect the mutual influence of electrons from nearby atoms
on one another. Because of this, each of the AT electrons that was in
a given energy level when the atoms were far apart will now have a
different energy in the crystal. Each discrete energy level in the
isolated atom now becomes a band of energy levels in the crystal,
lying very close to each other and numbering as many as the
number of atoms in the crystal. This state of affairs is illustrated in
fig. VI-3.1 show the bands as just grey patches in the picture. If you
imagine it hugely magnified, you will see the distinct levels, each
capable of accommodating one electron (the Pauli exclusion
principle). The electrons in the crystal are accommodated in these
energy bands, starting at the lowest energy level in the lowest band,
and going progressively higher in energy. You see that the basic idea
is the same as what we had with the atom, starting with the nucleus
and progressively adding the electrons to the different energy levels.
The difference is that the atomic energy levels arise from a single
isolated atom, whereas the energy bands for the crystal are due to
all the atoms of the crystal taken together.
You notice that the different bands in fig. VI-3 are separated by
energy gaps. It happens in some materials that a pair of adjacent (in
energy) bands overlap, instead of having a gap between them. The
net effect is still that of bands of allowed energies separated by
gaps, as in fig. VI-3. An electron can only have an energy that lies in
one of these bands but not an energy in any of the gaps. An
analogy: it is as if the rules of travel for cars on a highway were such
that they could only move at speeds between, say, zero and 20
km/hour, or between 30 and 40 km/hour, and so on. A speed 106
INSULATORS AND METALS
the crystal.
energy levels in
isolated atom
energy bands in
crystal
.a
occupation by electrons, in an
states left.
insulator
metal
energy gap J
ID
occupied states
empty states
or a rise in temperature on an
circles).
insulator
energy J gap
metal
energyjgap
occupied states
empty states
electrons each with a charge —e, so that its total charge is zero. The
outermost electron has wandered off in the metal, leaving behind an
ion with a total charge of+47e-46e, or just +e.
109
ELECTRON WAVES IN A CRYSTAL
What do the waves look like for the electrons that were in the outer
shells of the atoms in a crystal? The answer is illustrated in fig. VI-
10. The continuous curve shows a typical wave for such an electron
in a piece of crystalline solid as one goes along a line of atoms in
some region inside it. The broken curve shows a wave for a free
electron in the same region of a box that has the same shape and
size as the solid. The wave for the electron in the crystal has a
pattern of wiggles at each atom, but its overall amplitude changes
from one atom to the next like that for a free electron wave. One
would expect this kind of picture, if one recalls that the wave
amplitude at any point tells us about the probability of finding the
electron in the neighbourhood of that point. The probability of finding
the conduction electron inside the ion is strongly influenced by the
other electrons and the nucleus of the ion itself and leads to those
wiggles. Since the electron can move from atom to atom, it also has
the aspect of a free electron wave. It is rather like some folk dance in
which there are a number of identical groups of dancers at different
places on the floor. These are the electrons in the ions. The outer
electrons are like individual dancers who spend a little time with each
group before moving on to the next group. These waves of the
electrons in a crystal are called Bloch waves, after the physicist Felix
Bloch who first thought of them. Such electrons are referred to as
Bloch electrons, to distinguish them from free electrons in a box.
For the free electron in a box, we saw in chapter III that there is a
relation between the energy E and the wave number k,
b2k2
E=-
2m ■
how this variation, for fixed direction and changing magnitude of the
vector k, might look for two adjacent energy bands with a gap
between them. As the wave number increases, VI THE QUANTUM
MECHANICAL CRYSTAL
the energy in the lower band increases and the energy in the upper
band decreases. In this figure, each of the little squares, like the one
labelled P, represents a possible state with its specified wave
number and energy. For the state P, I can denote the values of wave
number and energy by £[P] and E[P] respectively, as read offon the
two scales. There can be a maximum of two electrons, with opposite
spins, in each of these states (Pauli exclusion principle). I show a
small number of states in each band to avoid cluttering the picture.
The actual number is comparable to the number of atoms in the
crystal, about 1021.
Figure VI-11 gives us more information about the energy bands than
is shown in fig. VI-3. We see that each band is made up of closely
spaced states, each with its wave function, wave number k, and
energy E. There is no state with energy lying in the gap between the
two bands. If a given electron changes from one state to another, the
FIGURE VI-11. Two neighbouring
the axes.
114
ELECTRON WAVES IN A CRYSTAL
figure tells us how much the corresponding changes are in the wave
number and energy of the electron. An electron can go from one
state to another within a band with very small change of energy, but
to go from one band to the other must involve a change of energy
equal at least to the energy gap.
dispersion curves.
I show in fig. VI-12 a few of the possible shapes for bands. In this
figure, I have not shown the little squares representing the separate
states in each band as in fig. VI-11, but have joined them as smooth
lines. For an actual crystal, there is one of these graphs of energy £
versus wave number k for each direction that the wave vector points
in the crystal, looking like so many strands of spaghetti in a bowl.
Such curves are called dispersion curves. A collection of all such
pictures for a given solid is called the electronic FIGURE vi-13.
Energy bands in an
n6
The electrons in the crystal occupy the states in the energy bands,
starting with the lowest energy, until all the electrons have been
accommodated. Each quantum state of given energy takes two
electrons with opposite spin. Then, depending on how many
electrons there are altogether, the last band containing electrons will
be either exactly full, or still have unoccupied states. These two
possibilities are shown in figs. VI-13 and VI-14 respectively.
The lower band is completely full, the upper band is quite empty, and
there is an energy gap between the two. To set up an electric current
means that some electrons have to change to states of slightly
higher energy. They cannot do this since the lower band has no
empty states for the electrons to move into, and the empty states in
the upper band are too far away in energy to be accessible to the
electrons. As a result, applying FIGURE VI-14. Energy bands in
energy.
increasing energ
noon
4—
“X,
♦ occupied
states
° empty states,
upper band
« empty states,
lower band
~ Fermi energy
“‘1™*-
Figure VI-14 shows how the bands are occupied in a metal. The
highest band that has electrons still has unoccupied states left. All
the lower bands (which are not shown in the picture) are completely
occupied by electrons, and the next higher band (shown) is empty.
Now if an electric voltage is applied to this crystal, electrons in states
below the one that is marked ‘Fermi wave number’ (more about this
in a moment) have empty states to move to, and so an electric
current results.
VI-14 and think about it in terms of wave vectors, then I would say
that all states with wave vectors up to a certain maximum length in
that direction are occupied by electrons, and states with larger wave
vectors are empty. This largest wave vector is called the Fermi wave
vector, after the physicist Enrico Fermi. There is such a vector for
each direction in the crystal, though their lengths may not all be the
same; after all, the crystal looks different in different directions, and
so do the electron waves. The magnitude of this vector is the Fermi
wave number.
question Why should the energy for all the Fermi wave vectors be
the same?
Fermi surface
empty states
n8
answer Suppose they all had different energies. Then there would be
a shifting of electrons from state to state until they all had the same
energy.
question What does the Fermi surface look like for electrons filling a
box?
A metal like silver has tens of electrons per atom. Most of these
electrons in the metal are in completely filled energy bands, and play
no part in carrying a current.
question Can one talk about a Fermi surface for an insulator also?
119
SEMICONDUCTORS
current.
o
■
♦ occupied,
lower band
° empty,
upper band
- occupied,
upper band
”’^
♦♦**
Figure VI-18 shows what this state of affairs looks like in terms of the
Fermi surface, which I have taken to be the same as the one in fig.
VI-15. Some electrons that were close to the Fermi surface acquire
thermal energy and move to states of higher energy just outside the
surface. This leaves an equal number of empty states inside the
Fermi surface, which behave like holes. You notice that only a thin
layer of states at the Fermi surface is affected by temperature. It
turns out that our interest is in knowing how much increase in energy
takes place with an increase in temperature, rather than the actual
energy at any given temperature. This means that we can think of
the thermal energy as being used to create electron-hole pairs, an
example of which is shown in fig. VI-18 on the right. An electron in
state A inside the Fermi surface gains thermal energy and changes
to state C outside the surface, leaving an empty state at A. One can
divide this gain in energy into two parts: the energy in going from A
to B (at the Fermi surface) is 123
empty states
c9
energy o:
Feraii
‘ electron
surface
\ / ‘V
\ / * electron
FIGURE VI-18. The same Fermi surface as in fig. VI-15 but now at
some temperature above the absolute zero (shown on left), and an
electron-hole pair (on right). The thermal energy is very small
compared to the Fermi energy at all temperatures of interest. This
means that only a few electrons very near the Fermi surface have
empty states of slightly higher energy to which they can go after
gaining thermal energy (dark circles). These electrons leave behind
empty states (open circles), equivalent to holes. The distribution of
electrons in all the other states is unaffected by temperature; these
electrons have nowhere to go, because all nearby states are already
occupied by other electrons. The thermal energy is divided between
the electron and the hole as shown on the right and described in the
text.
thought of as the energy of the hole, and the energy in going from B
to C as the energy of the electron. The thermal energy can create
these electron-hole pairs from only a small fraction of the total
number of electrons, namely those with energies close to the Fermi
energy. This means, as we shall see in chapter VII, that only a small
part of the thermal energy of metals is carried by the electrons and
holes. The greater part is in the vibrations of the ions.
PHONONS
~2L—
fundamental
positions of atoms
125
VI THE QUANTUM MECHANICAL CRYSTAL
or
f2 = 2f,
3v
Now look at the wave labelled ‘top harmonic’ in fig VI-19. The
neighbouring atoms are displaced in opposite directions from where
they would be if there were no wave. The wavelength here is just
twice the interatomic distance: I cannot have a wave that has an
even shorter wavelength. So I conclude that there is a minimum
wavelength for the lattice vibrations and a corresponding maximum
frequency of vibration. Since the wave number is just the reciprocal
of the wavelength, I can equivalently say that there is a maximum
wave number that I shall call &□ and a maximum frequency^) for the
waves. I use the subscript D here to recognise Peter Debye, the
physicist who first thought of the thermal vibrations of atoms in solids
in terms of sound waves.
126
PHONONS
g-
4567
harmonics
10
£„(«) = nhfu
CH intermediate
temperature
H high temperature
energy ranges
Let me pause now and take stock of what we have done. We began
with a crystal whose atoms were vibrating because of the thermal
energy of the crystal. These vibrations were equivalent to sound
waves, whose energies were quantized as phonons. The phonons
belong to the class of particles called bosons. The state of the crystal
at some temperature can be described, as far as the atomic
vibrations are concerned, by saying that the volume of the crystal is
filled with a gas of particles called phonons. There is a kind of
symmetry here with the quantum description of electrons confined in
a box. In that case, we started with objects that we think of as
particles (electrons), and ended by describing them as waves filling
the whole box. Here, with atomic vibrations in a crystal, we began
with sound waves filling the crystal, and ended up with particles
(phonons) moving about in the space occupied by the crystal,
128
property
mass
electric charge
statistics
total number in
crystal
electrons
yes
yes
fermion
kind of atoms
phonons
no
no
boson
varies with
temperature
8 The quantum mechanical crystal
We have come a long way from the crystal lattice of chapter II to the
quantum mechanical crystal of this chapter. The only property that
we gave the crystal lattice was its translational symmetry, from which
emerged other symmetries like those of rotation and reflection. We
then put atoms at the lattice points, let the crystal be at some
temperature, and used the results of the quantum mechanics of
atoms (chapters III and IV) and statistical physics (chapter V) to
describe the resulting situation. We find that the atomic nuclei sit at
the lattice points. The wave functions of the few outermost electrons
are significantly altered from what they were in the isolated atom,
and the remaining electrons are nearly unaffected by being in the
crystal. It is the outer electrons that contribute to those properties of
crystals that are of interest to us. Their energies lie in bands that are
separated by energy gaps. Within each band is a characteristic
relation between the electron’s energy and the wave number of the
electron wave. The electrons are fermions, and so a given quantum
state can hold at most two electrons with spins up and down
respectively. At the absolute zero of temperature the electrons are in
successive energy states in the bands, starting with the state of
lowest energy. The highest energy band with any electrons at all in it
will either have all its states occupied (an insulator), or will have
some fraction of the states empty (a metal). As the temperature of
the crystal is raised, a few electrons which are in the states of
highest energies acquire additional thermal energy, cross the energy
gap and go into empty states in the upper band if it is an insulator or
semiconductor. If it is a metal, electrons move to nearby empty
states in the same energy band. Empty states, equivalent to holes,
are left behind in either case. Most of the electrons in both cases do
not change their quantum states when the temperature of the crystal
is changed.
features.
AA
HI 1
AA1
o
AA ^
AA <
AA
AA r
w f\f\
AA t
AA |
AA
A/
AA
AA
1 f\f\
o
AA
AA
AA
electron AA phonon
Ohole
phonons
electrons, holes
Fermi surface
130
SUMMARY
132
VII COPPER WIRES AND GLASS RODS
I take a copper wire and a glass rod, both of the same size. If I grasp
the ends of the copper wire and bend it just a little and let go, it
springs back to its original shape. If I bend it more than a certain
amount and then let go, it stays bent. The same experiment with the
glass rod ends differently. The rod will bend a bit initially and recover
when I let go, but the attempt to bend it further will cause it to snap
into two or more pieces. I find yet another result if I try the
experiment with a wire of the kind of steel used to make springs. I
can bend this wire a fair amount, but it will spring back to its original
shape when I let go. I go back to the permanently bent copper wire,
straighten it out and again bend it and so on, repeating the cycle a
number of times. I find that it gets harder to change the shape of the
wire with each successive cycle: the copper wire begins to resemble
more and more the steel spring. Finally, I take the springy copper
wire from this last experiment and heat it up to 500 °C, say, and hold
it there for some time. This kind of heat treatment is called
annealing. I test the wire after it has cooled down, and find that it has
lost its springiness.
All these materials are made up of atoms, and they are held together
by electric forces derived from the charges of the electrons and
nuclei. Nevertheless, these experiments show that they respond in
different ways to a mechanical force that tries to change their shape.
These different types of behaviour can be understood when one
looks closely at how the atoms are arranged in real solids, as distinct
from the idealised perfect lattices that we have assumed so far.
Before I get to that, I summarize the behaviour of a solid when
subjected to forces as mentioned above.
sample, and note the resulting change in its shape. For a small
enough force, the change is proportional to the force, and is
reversible, i.e. the change is always the same for a given strength of
force. In this regime the material is said to show elasticity. This
prevails up to some maximum force, beyond which the deformation
of the material does not disappear if the force is taken away. At this
point the material has reached its elastic limit, and its size and shape
will have changed by a few parts in ten thousand. Beyond the elastic
limit, the material will either break (brittkness), or remain
permanently deformed (plastic deformation). I can summarize the
task now as an attempt to understand the elastic, plastic and brittle
behaviour of matter in terms of its atomic structure.
2 The springs
The elastic behaviour of solids is as though there were springs
connecting pairs of neighbouring atoms. The electric force between
the electrons and the nuclei in a solid is the only mechanism we
have to produce the effect of these springs. We saw in chapter VI
three different ways that the outer electrons of an atom arranged
themselves in solids. They were the following:
question Why does the electron have a lower energy when it stays
closer to the nucleus?
134
THE SPRINGS
(b)
+e
-2e
+e
answer Since there are both positive and negative charges in the
atoms, there must be both attractive and repulsive forces at play.
Each atom is sitting still in the crystal, apart from its zero-point and
thermal vibrations.
There is therefore no net force on it. This means that the attractive
and repulsive forces must cancel each other out. Stretching the
crystal changes the distribution of charges in such a way that the
attractive force dominates, whereas compressing the crystal causes
the repulsive force to be the larger.
3 Dislocations
I start with a perfect crystal of some element. I take it to be two-
dimensional, so that I can draw pictures that are easier to follow than
if it were three-dimensional. I show in part (a) of fig. VII-2 such a
crystal. The springs between neighbouring atoms in the FIGURE VII-
2. A crystal is depicted by
(b)
136
DISLOCATIONS
(a)
00
o
o
o
o
(b)
ooooooo
ooooooo
oo o”o 6 6
oooooo
oo^ooo
OO-OGO
J37
left.
(50
./
n
\
v.-‘ ^ ; Kj \
;i
) ‘■ )
> r\ r-\
) i. j (a) {J ( j ; ; ( j
CD
/■■ “\ .— “n
l :■•
i■;I
138
DISLOCATIONS
answer Think of trying to pull apart a piece of solid into two pieces;
this also requires a similar breaking of the bonds between all the
atoms on either side of the fracture.
o o ooooo
:■ * o o o o
o©ooo
o•ooo
oo•ooo
o o•oo o
(b)
ooooooo
oosoooo
oo©ooo
jOOOOO
OGOOOO
oo oooo
139
FRACTURE
than when
momo©o©
o #C0 o ©
©o•ooo
The basic reason for such hardening is that the atoms of the alloying
element along with some of the atoms of the host element take up
such positions that they act as traps for dislocations and inhibit them
from moving, as illustrated in fig. VII-7. This effect of alloying is
known as solution hardening or precipitate hardening, depending on
whether the alloying atoms are distributed uniformly or lie in small
clusters. By a judicious combination of the hardening methods, one
can produce for example the steels that stand large amounts of
elastic deformation and are used to make springs.
4 Fracture
There are materials like glass, which deform elastically a little bit, but
suffer fracture upon further deformation. We know that the atoms in
glass are not in regularly ordered positions as in a crystal. It is
therefore not possible to imagine in glass a dislocation such as is
illustrated for a crystal in fig. VII-4. It would seem that the only way
for glass to deform 141
(a)
142
SUMMARY
Although its surface may look smooth to the eye, there are always
present some microscopically small notches on the surface. When
the rod is bent, the concentration of force at one of these notches
reaches a level at which it begins to grow, and the rod fractures
there.
The force is then mostly concentrated at this scratch, and the rod will
break there. This method is used by glassblowers to break glass
rods and tubes to a desired length.
5 Summary
A given solid responds to a mechanical force applied to it in different
ways: it deforms slightly and reversibly, or permanently, or it breaks. I
describe these different behaviours, called elastic, plastic, or brittle,
in section 1.
144
VIII
l Some preliminaries
When I stir hot coffee with a silver spoon, I find that the handle
becomes warm very quickly. If on the other hand I use a plastic
spoon, its handle stays cool. I conclude that silver conducts heat
better than plastic. I find from experience that I can divide materials
into two classes: those that conduct heat well, and others that do
not. We know that metals like silver, copper or aluminium are good
conductors of heat and also of electric currents, while materials like
glass or plastics are poor conductors of both. So we may expect that
there is something common to the ways in which materials conduct
electricity and heat, and we shall see later that this is the case.
hot
(low of heat —^
bar of material
cold
145
VIII SILVER SPOONS AND PLASTIC SPOONS
the spoon.
1’ ■■
\\
\ ‘■
\\
\■
:\
\\
\
\
\ \ 4\
\\
\\
\%
\\
SPECIFIC HEAT
ture, as in fig. VIII-2, of how the temperature along the spoon varies
at different instants of time. The curves labelled 1 to 4 show the
temperatures at the initial moment of heating and at three later
successive times respectively. The sharp rise of temperature which
occurred at one end travels slowly down the spoon.
to 21 °C. This quantity is called the specific heat per gram of the
material. Since we know from other experiments the number of
atoms in one gram of each material, we can also calculate the
specific heat per atom, and this is shown in the second column of the
table. The table shows that the specific heat per gram is different for
different materials, whereas the specific heat per atom has about the
same value for the first three, and is lower for diamond. I shall return
to this point presently, but first we look at what happens to the
electrons and phonons in a solid when some thermal energy is
added to it.
copper
silver
lead
diamond
water
specific heat per gram
0.39
0.24
0.13
0.51
4.2
4.1
4.3
4.48
1.05
147
Thus the specific heat per atom should have this value for all
materials. The table above shows that this is approximately true for
copper, silver and lead, but not for diamond.
The reason for this deviation for diamond is that itsj^, is much larger
than thej^ for the other three materials, and one must go to much
higher temperatures to get kBTto be greater than hfo. When this
condition is not met, the statistical physics of phonons tells us that
the specific heat is smaller, and in fact heads towards zero as the
temperature approaches the absolute zero. All these predictions are
borne out by experiments.
Looking at the table again, we see that the specific heat per atom for
the three metals is in fact slightly larger than the value of 3kB which
comes from the phonons. This excess is partly due to the presence
of conduction electrons in the metals, which can receive thermal
energy and move to neighbouring quantum states that were
previously unoccupied, and thus add to the specific heat. We now
estimate this electronic contribution.
CL = 3A*b.
The subscript L reminds me that this comes from the /attice. Now
suppose that each atom sets free one electron; these electrons then
form a Fermi surface. At OK, all quantum states are occupied by
electrons up to a maximum energy called the Fermi energy £F, which
is about a thousand times the thermal energy per atom at room
temperature. At a temperature T, the additional energy which an
electron can take on as thermal energy is about k^T, which is a very
small fraction of the Fermi energy. As shown in fig. VI-18 (p. 124),
this means that only those electrons whose energies are less than
the Fermi energy by about ks T can move to empty states of higher
energy leaving empty states, equivalent to holes, behind. The
number of such electrons and 148
M:B2T2/£F.
If now I warm the crystal by one degree from Tto T+ 1, the energy of
the electrons and holes increases to
NkB2(T+l)2/EF.
The difference between these two energies is the energy per cc that
must be fed to the electrons and holes in the crystal in order to raise
the temperature by one degree, and is therefore the electronic
specific heat, which I denote by the symbol Cg (the subscript E
because this comes from the electrons and the associated holes):
CE = NkB2(T+ 1)2/£F minus NkB2T2/EF,
I have used a little bit of algebra to get to the last step, and it is easy
to verify it by putting in some numbers. We took Tto be 293 degrees
and so we must subtract 293
and therefore they must be responsible for its transport down the
bar.
answer The higher the temperature, the more thermal energy there
is in a given section of the bar. This energy is contained in the
phonons, and so there must be more phonons in this section. You
can see this in fig. VI-21
(p. 128).
Consider a phonon towards the hot end of the bar in the region
marked A. It will travel a certain distance before it collides with
another phonon. Suppose that this collision takes place in the region
marked B. The collision changes the phonons to other phonons
whose total energy is equal to that of the colliding phonons because
the total energy is conserved. Now there is more energy, and so
there are more phonons, in the region of the collision than is
prescribed by the temperature there. These surplus phonons
continue on until the same thing happens to them further down the
bar. There is at the same time a flow of thermal energy in the
opposite direction at B, due to phonons arriving there from a region
like C which lies towards the cooler end of the bar. The net flow of
energy is then the difference between these two.
quartz 100
silicon 430
common salt 70
copper 850
silver 1140
iron 440
How many cars do I see going past in one hour? The answer is 10
cars/km multiplied by 100 km/hour, or 1000 cars per hour. By a
similar reasoning, the thermal energy flowing to the right per second
at the surface B is given by 2 £(7^) times the area S times the speed
v of the phonons.
Note that the speed of the phonons is the same as the speed of
sound, which in most solids is a few thousand metres per second.
There are also phonons at B which are travelling to the left. These
phonons arrived there from C where they had a thermal energy
E(T2) per cubic cm, and they bring a thermal energy per second
flowing to the left equal to
So the net thermal energy flowing to the right per second, which I
denote by Ql> is equal to the difference of the above two quantities:
I now do a little bit of arithmetic to get the above formula for <2l into
a form which shows me how to relate Ql t0 the specific heat and the
mean free path, which are properties of the phonons in the material.
I multiply gL by two numbers which are each equal to one, so that
the result is stil
along the bar. I can then write the amount of thermal energy going
down the bar per second as
QL = vCLlSR.
I have assumed so far that all the phonons are travelling either to the
left or to the right along the bar. In fact, of course, they are travelling
in all directions. When I take this also into account, the formula for
Q^ is modified to <2L = jvClISR,
QL = KtSR.
and 20 °C respectively.
\
o
21 C
A
\
O
20 C
B
\
Table VIII-3. Flow of heat per second for the cube shown in fig. VIII-5
material of cube
quartz
glass
silicon
common salt
silver
copper
stainless steel
brass
typical plastic
flow 0
0.07
0.01
1.5
0.064
4.2
4.0
0.14
0.8
0.1
The way that thermal energy is conducted by the electrons and holes
is exactly the same as the way by phonons. I get the same formula
for the thermal conductivity as in the case of the phonons, except
that I must use in it the corresponding quantities for the electrons
and holes: the specific heat CE, the Fermi velocity vp, and the
electron’s mean free path E. Calling this electronic thermal
conductivity CE, I get So the thermal energy carried per second by
the electrons and holes is 2E = KESR,
Q = Qe + Ql—
In an alloy, we have more than one kind of atom in the lattice: copper
and zinc atoms in brass, for example. This means that we have lost
the translational symmetry of a pure metal in which every lattice
point is occupied by the same kind of atom. This departure from
perfection acts as a disturbance to the particle (electron, hole or VIII
SILVER SPOONS AND PLASTIC SPOONS
You will notice from table VIII-3 (p. 154) that quartz is a much better
conductor of heat than glass, even though they both consist of
molecules of silicon dioxide. So the mean free path for phonons in
glass must be smaller than in quartz. The difference between the two
materials is that quartz is a crystal and the molecules are arranged in
it with translational symmetry, whereas in glass there is no such
symmetry. So we come to the general conclusion that loss of
translational symmetry, whether caused by alloying or by conversion
into a glass, causes the mean free paths of phonons and electrons
to decrease. As a result of this the thermal conductivity becomes
less.
6 Thermal expansion
When a metal cap on a glass jar has been screwed on so tightly that
it cannot be unscrewed by hand, an old kitchen trick is to heat the
cap, for example by running hot water over it. This works in some
cases, and the cap comes off easily. This could be because a bit of
whatever was in the jar had gone hard in the space between cap and
jar, and the heating simply softened it. But the method can also work
when everything is clean. The reason then is that the heating causes
the cap to expand in size so that it is no longer a tight fit on the jar.
Such a change in size of a material when its temperature is changed
is called its thermal expansion. In practically all situations that we are
likely to encounter, it is extremely small and would not be seen by
the unaided eye.
copper 17 x 10”4
iron 12 X 10”4
aluminium 24xlO~4
glass 5xlO”4
You notice that different materials expand by different amounts for
the same rise in temperature, and that glass expands less than the
listed metals. One sees now why a stuck bottle cap can be freed by
heating it, even if the glass gets a bit warm.
156
THERMAL EXPANSION
cool
gas-filled balloon
warm
phonons
electrons
phonons
electrons
cool
warm
a solid
Now imagine, as shown in the lower part of fig. VIII-6, the balloon
replaced by the surface of the solid, and the gas molecules by the
phonons in the solid, and the free electrons also if the solid is a
metal. These phonons and electrons exert a pressure on the
surface, and the pressure increases with increasing temperature.
This results in the thermal expansion of the solid.
on the wall.
paths of molecules^
a small portion
of balloon wall
that are hitting it and bouncing back all the time. As each molecule
bounces back, it causes a recoil of the wall directed to the right in the
figure. What is at work here is the conservation of momentum. The
momentum of the molecule changes as a result of the collision with
the wall. This change must be compensated by a change of
momentum of the wall, such that the total momentum remains
unaltered.
The cumulative effect of such recoils due to all the molecules hitting
the wall is that the gas pressure on that side of the wall tends to
move it to the right. At the same time, the wall suffers similar
collisions from molecules on its other side too, and these give it an
impulse which tends to move it to the left. The impulse, and therefore
the gas pressure, increases when the speeds of the molecules
increase, which is what happens when the temperature of the gas
rises. Thus we see why the balloon expands when the gas inside it
gets warmer.
question Where does the elasticity of the balloon come into all this?
answer The excess pressure of the warm gas inside the balloon over
that of the cool gas outside is balanced by an effective pressure of
the balloon acting inwards, resulting from the fact that it has
increased in size. This is a bit like a stretched rubber band trying to
return to its original size.
158
The free electrons in a metal, which are largely responsible for its
thermal conductivity, are moving about inside the metal, but cannot
get out through its surface. The attraction between the negatively
charged electrons and the positively charged ions in the metal holds
the electrons trapped inside the metal. The effect of this attraction is
as if there were an impermeable but elastic sheath all over the
surface of a piece of metal. The electrons are confined inside this
sheath, with which they are constantly colliding and recoiling, in
analogy with the gas molecules and the balloon wall. There is a
resulting pressure due to the electrons on the surface of the metal.
The average speed of the electrons increases when the temperature
rises, thus causing this pressure to increase and the metal to
expand. We call this the electronic contribution to the thermal
expansion.
Now how about the phonons? The electrons have mass, and they
are moving about at various speeds. It is easy to imagine that they
transfer impulse to the wall when they collide with it. The picture is a
little different with phonons. They represent the quanta of energy of
the vibrating atoms, and there is no obvious way in which we can
associate a mass with a phonon. A phonon carries no momentum in
the sense in which a molecule in a gas does. You can see this when
you recall that the phonon is the quantum aspect of a sound wave in
the solid. For each atom in the sound wave with a given speed in
some direction at any given instant, there is another atom with the
same speed but in the opposite direction, so that the total
momentum of the wave will be zero. We cannot talk about a phonon
colliding with the surface of the solid and transferring momentum to
it.
159
While eating a meal, you may have on occasion put a potato in your
mouth which was much hotter than you expected. You may then
have noticed how even a sip of water is enough to cool off the
potato. This is because the specific heat of water is much greater
than that of the potato. Even though the amount of water is small, it
undergoes only a modest rise in temperature as it absorbs the
thermal energy coming out of the potato as it cools. The specific heat
of water is one of the highest for any material we know, as is evident
in table VIII-1. This is why, for example, if a pan of water and a
garden tool are left outside in sunlight, the tool may become too hot
to touch whereas the water may be barely warm. It is this ability of
water to absorb large amounts of thermal energy with only modest
changes in its temperature which is an important factor in
determining the earth’s climate.
degrees. Thus the heating power now needed will be 13/17, which is
72
The copper bottom spreads the heat from the stove uniformly and
therefore helps to prevent hot spots with their deleterious effect on
what is in the pot.
160
SUMMARY
5°C
outside
22°C
inside
161
162
l Some preliminaries
2 What is light?
other materials like aluminium foil. We have seen in chapter III that
light has the property of duality: it behaves like waves showing
interference and diffraction, and like particles (which we call photons)
in the effect discovered by Arthur H. Compton when a beam of X-
rays is scattered by electrons. The momentum p and the energy £ of
a photon are related to the wave vector k and the frequency/of the
corresponding light wave by the rules of quantum mechanics:
p = bk
and
E=hf
The other colours have wavelengths between these two values. The
frequency is given by the speed of light, 3 X 1010 cm/sec, divided by
the wavelength.
IO24
IO22
1020
10
18
£ 10
16
I10‘4
lio12
10
in
10
106
104
102
1
gamma rays
-A
X-rays
ultraviolet light
ImsIBIeillgEt”!
infrared light
microwaves
television, FM and
AM radio
long wavelength
waves
10
,-14
10
10
10
,-10
-6
10
-2 M
IO4
IO6
IO8
10
10
also the speed of visible light. You can verify from the figure that this
is the speed of all the waves in the entire spectrum.
3 Electric and magnetic fields
In this section, I describe what it is that is waving in an
electromagnetic wave. In order to do this, I must first explain what is
meant by the term field as used in physics. This is 165
■••■■7
/\r
*p
/
a term which we use here with a special meaning which is different
from what it has in everyday speech. I begin with a simple example:
a kitchen oven which has been turned on. I show the oven in fig. IX-2
as a hollow box. The air inside the oven is steadily getting hotter. At
some instant of time, which I denote by the symbol t, I measure the
temperature T at every point inside the oven. I can specify any point
P by noting its position vector r from one corner of the oven, as
shown in the figure. Changing the direction and/or magnitude of r will
let me place P anywhere in the oven. What I have done, at the
chosen instant of time, is to make a snapshot picture, so to speak, of
the temperature everywhere inside the oven. I represent this picture
by the symbol T(r,t).
This symbol stands for the temperature at every point in the oven at
all different times, and is therefore a very convenient shorthand for a
lot of information. In words, I say that there is a temperature field
inside the oven, which depends on position r and time t, and is
denoted by T(r,t).
question Imagine that the air inside the oven is not still, but moving
about as suggested by the curved arrows in fig. IX-2. Can you think
of a field associated with this motion?
166
answer Yes. At each point P the air is moving with a certain speed in
some direction, both of which may change with time and with the
position of P. I can therefore think of a velocity field which I can
denote by v(r, t) for the air in the oven.
The temperature field is a quantity which has a magnitude (so many
degrees Celsius, for example), but no direction. We call such a
quantity a scalar. The temperature inside the oven forms a scalar
field. The velocity of the regions of air inside the oven forms a vector
field. What is common to both fields is that, over a given volume of
space and span of time, a certain quantity which can be a scalar
(e.g. temperature) or a vector (e.g.
I can write down a formula for the force which the unit positive
charge feels due to the charge of the electron. I just have to multiply
the charges together and divide by the square of the distance apart:
<■!
S■i
due
oq
in
■ -i
! JltlllllTllTlTllTOlirinTirnrii
167
resemble those of a large bar magnet lying along its axis of rotation.
I illustrate in part (a) of fig. IX-4 the influence of a bar magnet on
compass needles placed in various positions in its neighbourhood.
At each position, there is a definite direction in which the needle
points. To turn the needle away from this direction requires a twisting
torque which varies from point to point, becoming smaller the further
it is from the magnet. Thus the influence of the bar magnet upon the
compass needle has two aspects: a direction (the direction of the
needle) and a magnitude (proportional to the torque needed to
disturb the needle). Further, it is an influence which pervades the
space surrounding the magnet. I can include all of this in the
statement that a bar magnet creates a vector field in the space
around it, called a magnetic field and denoted by the symbol B. As
with the electric field E, the magnetic field B may in general vary with
time and position. The direction of the magnetic field vector
determines the direction of the compass needle, and its magnitude
tells us how strongly the needle sticks to this direction.
A bar magnet is not the only source for a magnetic field. An electric
current flowing in a wire will also produce a magnetic field in the
surrounding space, as illustrated in part (b) of fig. IX-4. It shows a
short section of the current-carrying wire, and the orientation of
compass needles placed at different points around the wire. The
orientation shows the direction of the magnetic field, whose strength
decreases with increasing distance from the wire.
169
GENERATION OF LIGHT
it. The wave consists of electric and magnetic fields, and to sense
them I must place something suitable at P. To see what the electric
field is doing, I imagine an electron at P. It senses the oscillating
electric field of the wave moving past it. It feels an increasing force
which reaches a maximum when the crest of the wave (point C in the
figure) reaches it. The force then begins to decrease to zero and
reaches a maximum in the opposite direction when the point D of the
wave reaches P, and so on. Thus the electron feels an oscillating
force at a frequency f, the frequency of the wave. The electron also
has a magnetic moment, and so it senses the magnetic field too.
get the result shown in fig. IX-6. The shape of this distribution results
from the fact that photons are bosons, and is the same whatever the
temperature of the body: 3000 K
for the filament in an electric bulb, 6000 K for the surface of the sun,
and so on. You can see that the largest number of photons are at
energies which are about three times the thermal energy. As the
frequency of the photon is proportional to its energy, this means that,
for example, the largest numbers of photons in sunlight are at
frequencies which are twice those in the light from an electric bulb.
When I put the actual numbers in, I find that most photons from the
sun are in the red part of the spectrum with a wavelength of about 8
X10”5 cm, while they are in the infrared (wavelengths around 1.6
X10”4 cm) in an electric lamp.
172
ABSORPTION OF LIGHT
question As you read this, you and your surroundings are also giving
off a continuous spectrum of electromagnetic radiation. What
frequencies do most of these photons have relative to those from an
electric lamp?
One can see from fig. IX-6 that the lower the temperature, the lower
the range of photon energies at which the largest numbers of
photons are found. The universe is filled with electromagnetic
radiation which corresponds to a temperature of about 3 K; this
means that the largest number of photons in this radiation have
wavelengths in the vicinity of a millimetre. This radiation is a residue
of the big bang which was the start of our universe 1O10 years ago.
5 Absorption of light
All solids consist of particles, electrons and nuclei, which have two
properties which make them feel the influence of the electric and
magnetic fields in an electromagnetic wave: they are electrically
charged, and they are permanent magnets. In addition, quantum
mechanics tells us that the energies of the electrons are to be
described in terms of energy bands, and the energy of the
electromagnetic wave is quantized in terms of photons. Using this
information, we now seek answers to questions like why glass lets
light through and aluminium foil does not, and why copper is red and
sapphire is blue.
173
The light also behaves like photons, quanta of energy. A photon can
be absorbed by an electron, whose energy is then increased by the
energy of the absorbed photon. For this absorption to be possible,
there must be a quantum state of this higher energy into which the
electron can move. If there is no such state, then the photon cannot
be absorbed and will therefore continue on its way. Now we begin to
see why some materials are transparent and some are opaque.
Now consider photons of energy larger than the gap energy, CD for
example. An electron can absorb such a photon and go to a
previously empty state in the upper band.
174
ABSORPTION OF LIGHT
absorbed.
photon erergy
empty
energy gap
full
surface of metal
©1
<• ©2 >
<■■•• © •■•>
<• © >
until at some depth below the surface the light wave and the electron
oscillations have both died out. This depth in most metals is very
small: a few tens of angstroms (recall that one angstrom is 10”8 cm).
So we can see now why even a very thin sheet of metal lets no light
through.
You might wonder where the light which is reflected from the surface
of the metal comes from, since it looks as if all the light falling on the
surface got absorbed in setting up the oscillations of the electrons.
You recall that electromagnetic waves are generated by oscillating
electric charges. The surface layers of electrons are oscillating at the
same frequency as the incoming light. These oscillations produce
electromagnetic waves of the same frequency, and these waves are
the reflected light.
What I have described so far seems all right for a metal like
aluminium, which reflects daylight without changing the colour of the
reflected light. But what about metals like copper and gold, which
look red and yellow respectively in ordinary daylight which has all the
colours of the visible spectrum? The answer lies in a special feature
of the electronic band structure of these metals, which I illustrate in
fig. IX-9. The conduction band, labelled band 1 in the figure, has all
its quantum states occupied by electrons up to the Fermi energy,
and the states of higher energy are unoccupied. This part of the
band 176
ABSORPTION OF LIGHT
empty states
B
A
occupied stales
Z8l?r.:.:.-:..;.i:
Hi
On the other hand, photons with energy equal to or greater than the
amount corresponding to CD can be swallowed up by electrons in
band 2 which then move to previously empty states above the Fermi
energy. The reflected light will therefore be missing all these photons
and will be different from the light which hit the surface of the metal.
177
answer Yes, the electron will go back to C; but the photons emitted
in this process come out in all directions, and so the number in the
direction of the reflected beam will be much smaller than that in the
incoming beam of light.
question How does one explain the red colour of ruby or the blue of
sapphire, which are insulators and therefore have no free electrons?
178
ABSORPTION OF LIGHT
absorption.
.2 0.00001
a.
On
On
0.001
iooo—
increasing frequency
179
that are present in all foodstuffs, thus raising their energy and
temperature. The plastic or glass of which the container is made
does not have those energy levels which would permit the
absorption of these photons, and therefore remains relatively cool.
The conduction electrons in a metal, on the other hand, can absorb
these photons and thus heat the metal, as anyone who has
mistakenly left a metal spoon in the oven will have noticed.
wave had gone on through the crystal without being affected by it,
then it would be like the broken line in the right half of the figure, with
the same wavelength X(0). But in reality the wave is affected by the
crystal, because it produces displacements of the electrons and
nuclei which make up the atoms. The magnitude and direction of this
atomic polarization varies from atom to atom, and I have indicated
this by a small square against each atomic position at a distance
proportional to the polarization. For example, the polarization of atom
A is about equal in magnitude but opposite in direction to that of
atom B. The electric field is tied to the polarization at each atom, and
so we have an oscillating electric field with the same wavelength as
the oscillating polarization, and this is then the light wave in the
crystal. Note that the light wave in the crystal is associated with the
movement of things with mass, namely electrons and nuclei. This is
in contrast to the same light wave in free space, where there is no
mass to be moved. It is this difference which leads to the wavelength
of light being shorter in matter than in free space. Since speed is
equal to frequency multiplied by wavelength, the speed of light in any
material medium is less than its speed in free space.
The fact that the speed of light is less in a solid than it is in free
space causes a beam of light to change its direction as it goes from
one to the other. This effect, known as the refraction of light, is what
makes possible a host of optical equipment: eyeglasses, cameras,
telescopes, and microscopes, to name a few. I illustrate how
refraction comes about in fig. IX-12. A beam of light comes from the
left and enters a transparent solid.
181
space.
empty space
crystal
This last condition can be satisfied only if the beam in the solid is in a
specific direction, and none other, relative to the beam in empty
space.
answer Light travels the distance AC in free space in the same time
that it travels the distance BD in the solid. Therefore the ratio of vto
cis equal to the ratio of the length BD to the length AC.
answer Yes. Imagine the solid replaced by water, and the arrows
showing the direction of the beam of light to be reversed. So now the
light is coming from the bottom of the pool, say, and its direction is
changed as shown when it emerges from the water. A person
looking into the water will therefore see the bottom in the direction
shown by the broken lines.
The bottom will thus appear to be nearer to the surface of the water
than it actually is. In fact, the actual bending of the beam of light as if
emerges from water to air is such that a pool which is two metres
deep will seem to be only a metre and a half deep: inexperienced
swimmers, beware!
7 Dispersion of light
The speed of an electromagnetic wave in empty space, whatever its
frequency, is the same,“and we denote it by the symbol c, with a
value of 3 X 1010 cm per second. When the wave travels through a
material medium, not only is its speed reduced, but the amount of
reduction depends on both the frequency and the medium. This
effect, 182
DISPERSION OF LIGHT
sunlight
violet
rainbow ^ red
called dispersion, is greater the more densely packed are the atoms
in the medium. For example, the dispersion of light in air is negligibly
small, so that light of any colour has practically the same speed, c, in
air. In water or in glass, on the other hand, there is a significant
decrease in speed which is different in the two substances and for
different colours. The amount of bending of light as it enters (or
leaves) water or glass from air therefore depends on the colour. This
is the origin of the rainbow, and the spectrum of colours produced
when sunlight passes through a glass prism, to name two examples
of dispersion. I illustrate these effects in figs. IX-13 and IX-14.
But this does not happen. The rules which govern reflection and
refraction of light combine with the geometrical properties of the
sphere (which is the shape of the raindrop) to make the light coming
out in a particular direction (which changes slightly with colour) to be
more bright than in all other directions. The result is the rainbow.
The fact that the speed in water of blue light is lower than that of red
light is what ultimately causes blue to be on the inside and red on the
outside of the rainbow. Only a part of the light emerges from the
raindrop after the first reflection. The rest is reflected once more at
the raindrop’s surface, and a part of this emerges and gives rise to
the second rainbow which can sometimes be seen. The order of
colours in this rainbow is the reverse of that in the first. This fact of
nature has been overlooked by some artists famous for their
landscape paintings. The Neue Pinakothek in Munich for example
has two landscapes with double rainbows, painted by Jan Wildens
and by Josef Anton Koch, where the colours are not in the order that
nature says they should be.
183
IX GLASS PANES AND ALUMINIUM FOILS
sunlight
spectrum
E=hf.
Thus the light coming out has a spectrum of colours with certain
frequencies which are characteristic of the atoms producing it.
Knowing the spectrum, one can deduce which type of atom it comes
from. The rainbow is the spectrum of the thermal radiation from the
sun produced by a spectroscope that occurs in nature, namely rain.
8 Solar batteries and LEDs
A solar battery is a semiconductor device which converts some of
the energy it receives from sunlight into electrical energy, just as an
ordinary battery converts chemical energy into electrical energy. The
LED (acronym for /ight-emitting rfiode) is a semiconductor device
which does the reverse: it uses electrical energy to produce light. A
solar battery is a practical application of what we have learnt about
the energy bands in semiconductors.
Figure IX-15 shows how a solar battery works. In part (a), an n-type
and a p-type semiconductor are close together but not in contact.
The n-type has electrons, and the p-type has holes, moving about. In
part (b), the two semiconductors are brought into contact to form
what is called a p-n junction. The electrons and holes at the junction
184
n-type
n-type
(a)
p-type
(b)
n-type
sunlight
G+i-R © e
©+©e
e e Jbs © © ©
©I!
©© ©I!-©©©©
e©e | e g
holes, current.
gadget
p-type (c)
electrons
© electron
© hole
- negative ion
+ positive ion
move across and neutralize each other, leaving positive ions on one
side and negative ions on the other.
question Why does the process not continue, thus neutralizing all the
moving charges on both sides?
answer An electron from the left which tries to move through the
junction meets the phalanx of negative ions on the right and is
repelled; it cannot get through. The positive holes on the right are
similarly stopped by the positive ions on the left. Remember that like
charges repel one another.
185
question What happens to the electrons in the n-type and the holes
in the p-type which are left behind?
answer: They find their way to the region of the junction, where they
neutralize each other.
9 Summary
This chapter deals with the nature of light, and how it interacts with
material bodies.
186
SUMMARY
In section 7,1 find that the reduction in the speed of light is different
for different wavelengths and substances. This is called the
dispersion of light. This is the origin of the rainbow, and the principle
is used in an instrument called a spectroscope which analyses light
from a source into its constituent wavelengths.
187
CABLES
l Preliminaries
transformers
switches
meters
etc.
appliance
\ household wiring
230 V
or
115 V,
30 W
water pump
pipes
pipes
turbine
188
PRELIMINARIES
space inside pipe which carries metal (copper, aluminium) wire the
water which carries the electricity
pressure voltage
turbine
water current (amount of water electric current (amount of per sec)
electric charge per sec)
I shall say a few words about volts and amperes, which are
respectively the units for measuring voltage and electric current. If
the voltage of a battery is two volts, for example, it means that the
difference in potential energy of an electron in moving from one
terminal of the battery to the other is the charge multiplied by the
voltage, or 2e, where eis the magnitude of the electron’s charge. It is
this energy which gets converted to other forms of energy when the
electron goes through the appliance. The electric field and the
voltage are closely related. In a battery, for example, there is a
voltage between the terminals, and there is an electric field
everywhere in the space between the terminals.
to prevent it, the speed would continue to increase, and so would the
current. But with a long and thin enough wire, the current quickly
reaches a constant value and stays there. I find that the size /of this
current is proportional to the voltage V. So the voltage divided by the
current is a constant:
?=*
germanium 1
190
106i
a
4=1
co
CO
0>
10
-6
10
rl2
semiconductor
metal
100
temperature
200.
metal
191
X ELECTRIC BULBS AND INSULATED CABLES
Nothing about this picture should change if I cool the wire to a lower
temperature, and I would therefore expect the resistance to be
unchanged. But, as fig. X-2 shows, this is not what happens: the
resistance drops dramatically as the temperature drops.
I said that we could forget about the ions. Well, not quite: it is only
the static ions that we can forget. There are still the vibrations of the
ions about their mean positions.
SCATTERING OF ELECTRONS
in the solid. It is the scattering of the electrons when they collide with
the phonons which limits the electric current, and leads to a
resistance.
Electrons are fermions, and this means that each quantum state with
a specified k vector can have no more than two electrons with
opposite spins. I described in chapter VI (p. 118) how this leads to
the picture of a Fermi surface inside which lie the wave vectors of
these electrons. Remembering that the momentum is equal to the
wave vector multiplied by the Planck constant, and also to the
velocity multiplied by the mass, I can think of the Fermi surface as
giving me a picture of the velocities of the electrons. Figure X-3
taken together.
without current
The current is the charge passing a given point per second, and is
therefore proportional to v, which is called the drift velocity. The time
X which elapses before the electron is scattered is called the
scattering time. A longer scattering time means a larger drift velocity
and therefore a larger current too. Furthermore, the current is also
proportional to the number per unit volume (called the density) of
electrons which are there to carry it. So we should look to the
scattering time and the density of electrons for an explanation of the
different behaviours of the electrical resistance of metals, alloys and
semiconductors shown in fig. X-2.
We take the case of a pure metal first. When its temperature goes
up, so does the amount of thermal energy in it. This energy consists
of the vibrations of its atoms about their average positions, and this
motion is quantized into phonons. The number of phonons increases
with an increase in temperature, and so the frequency of collisions
between electrons and phonons also increases. This means that the
scattering time (which is the time between collisions) and therefore
the conductivity decrease as the temperature rises. In other words,
the electrical resistance should decrease with decreasing
temperature in a metal, and this is what happens. The electron
density may vary from metal to metal, but does not change with
temperature in a given metal. We thus have a basic explanation of
the different resistivities of different metals, and how they change
with temperature.
In an alloy, we have more than one kind of atom in the lattice: copper
and zinc atoms in brass, for example. We saw in chapter VIII (p.
155) that this leads to an additional scattering of the electrons, due
to the loss of translational symmetry. This in turn 194
SCATTERING OF ELECTRONS
How about the scattering time, which also influences the resistivity of
the semiconductor? The scattering is still by phonons, and is
stronger at higher temperatures just as with metals. But its effect in
changing the resistivity is weak in comparison with the effect of the
change in number of electrons and holes, and the net result is a drop
in resistivity with increasing temperature, as is actually observed.
conduction band
valence band
• occupied state
O empty state
An insulator, such as the material which covers the copper wire used
in households for carrying electric currents, has essentially the same
band structure as a semiconductor, but with one important qualitative
difference. The energy gap between the filled band and empty band
above it is so large, that at ordinary temperatures there are very few
electrons with enough thermal energy to raise them from the filled
band to the empty band. This means that negligible current is
produced when a voltage is applied: we have an insulator.
196
SEMICONDUCTORS AT WORK
current between the source and the drain, and there we have the
basic functioning of a transistor. There is a similarity to a rubber hose
through which water is flowing at a certain rate because of a
pressure difference between its ends. This flow rate can be varied by
pinching the hose at some point so that its cross-section changes
there.
When an electric heater is switched on, current flows through it, and
heat is given off. Suppose the current is amperes, and the voltage is
V volts. A current ofamperes means that / coulombs of charge, in the
form of electrons, are flowing through the heater per second. The
potential energy of the electrons is just the charge multiplied by
oxide
accordingly.
source
drain
198
the voltage. Since the voltage along the heater drops by F volts, the
potential energy of the electrons leaving the heater per second is 7x
V joules less than that of the electrons entering the heater. This
energy which is lost by the electrons in their passage through the
heater appears as thermal energy, at a rate of IX Fjoules per second,
or /x Fwatts.
force also.
current—
199
between the flow of electrons and the flow of water. I introduce the
units used in measuring electric currents.
2OO
SUMMARY
201
XI
MAGNETS
l What is a magnet?
(a) When I bring two magnets together so that the north pole of one
approaches the FIGURE XI-1. Basic features of a
(a)
attract IN!
magnet
(P)
repel
(c)
JED attract1
1 attractc
iron
iron
(d)
nothing C
glass
202
ELEMENTARY MAGNETS
south pole of the other, the magnets attract each other: there is an
attractive force between unlike poles.
(b) If two poles of the same kind approach each other, there is a
force of repulsion between them.
(d) If the rod is of almost any other material (copper, glass, plastic,
etc.), it feels a negligible force from the magnet. Such materials are
not ferromagnetic.
We take the properties (a) and (b) as given by nature, rather like the
forces between electric charges. In fact, it is tempting to push the
analogy and to imagine that there are separate north and south
poles, which could be called magnetic monopoles. The analogy here
is with electrons and positrons which can be called electric
monopoles. A magnet would then be a concentration of north poles
at one end and south poles at the other end of the bar. However,
many experimenters have tried without success to see if such
magnetic monopoles actually exist. Our experience so far is that any
magnet, whether of the kitchen variety or an electron or proton or
neutron, always comes with both a north pole and a south pole. The
two are inseparable: breaking a magnet in two just gives two smaller
magnets, and not separated north and south poles. So it seems that
an atomic description of magnetic materials must be based on the
magnetism of the constituent particles as an immutable fact of
nature, like for example the charge of the electron.
2 Elementary magnets
All matter is made up of electrons and nuclei, and just as with the
other properties, we want to understand the magnetic properties in
terms of these fundamental constituents of matter. I have said earlier
that an electron, and most nuclei, behave as if they were spinning
tops as well as permanent magnets. I go into this now in a little more
detail.
I begin with the electron. It is a magnet, but a very weak one indeed,
compared to an ordinary permanent magnet. I would need
something like 1021 of these electron-magnets all pointing the same
way and crowded into a volume equal to that of an ordinary bar
magnet, in order to produce the same magnetic field as the magnet.
203
XI MAGNETS
question Why must the electron-magnets all point the same way for
this effect to happen?
The proton and the neutron are also magnets, but much weaker than
the electron. It would take a couple of thousand of them to equal the
magnetic moment of an electron.
An atomic nucleus is made up of protons and neutrons, and so in
general would also have a magnetic moment which is much smaller
than a Bohr magneton. We shall see later that even these small
moments are important in some cases.
FERROMAGNETS
cancelled pairwise. This will happen, for example, if all the protons
and neutrons are separately paired up with the spins in each pair
pointing in opposite directions, thus: t J-. Then the total magnetic
moment of the nucleus will be zero.
I said that about 1021 electrons crowded together with all the
moments pointing the same way would have the strength of an
ordinary magnet, say a compass needle. This number is in fact
comparable to the number of electrons in a piece of solid of that size.
This spontaneous lining up of the electron moments takes place only
in a few materials, among them iron, nickel and cobalt. Furthermore,
we do not have to worry about the magnetism of the nuclei in this
case, since it is thousands of times weaker than that of the electrons.
How does this lining up of the electron moments take place? The
answer is in the next section.
3 Ferromagnets
In an atom, most or all of the electron moments cancel out one
another in pairs, with the two moments in each pair pointing in
opposite directions. Then there may remain a few electrons whose
moments have not got so cancelled out. Atoms with such unpaired
electrons are potential candidates for forming ferromagnets. We
know that there are forces of attraction and repulsion between
magnets. So it is tempting to think that somehow these magnetic
forces between the unpaired electrons on neighbouring atoms may
lead to all the moments lining up, giving a ferromagnet. But when
these forces are worked out in detail, they turn out to be too weak to
give rise to the known ferromagnets like iron and nickel. We must
look elsewhere for the origin of this property.
XI MAGNETS
nent magnet comes already equipped with its magnetism, unlike the
piece of iron? I shall answer these questions next.
One might think that one would get around this problem by using a
piece of iron which is entirely one crystal. There being no grains in it,
the piece should have the electron moments all pointing in one easy
direction, and therefore become a magnet. But this does not happen.
The easy direction for the moments to point in iron is along an edge
of the unit cube in the crystal: but there are six such directions in a
cube, as you can see in fig. XI-3, part (a). The crystal gets divided
into regions, and the moments in each of these regions will point in
one of the easy directions, but they will be different in the different
regions. Each of these regions is called a magnetic domain. I show
in fig. XI-3, part (b), a possible arrangement of domains in a crystal
shaped like a rectangular bar. The final result is the same as with the
lump of iron: the single crystal bar will not show strong magnetism,
because the effects of the different domains cancel one another out.
shows no magnetism.
206
XI MAGNETS
One can also use the magnetic field produced by an electric current
flowing through a spool of copper wire wrapped around a rod of iron
to make it magnetic, and then one has an electromagnet. This is the
kind of magnet used in many applications where magnetic fields are
needed.
**«
®$*$$$
208
FERROMAGNETS
material
iron
nickel
cobalt
gadolinium
dysprosium
1043
631
1403
289
105
Figure XI-5 shows how the strength of the magnetism varies as the
temperature changes from near the absolute zero to the Curie
temperature. It has the highest value at the lowest temperature,
gradually drops as the temperature is raised, and becomes zero at
the Curie temperature.
question Can this effect be because the electron gradually loses its
magnetic moment as the temperature goes up?
ferromagnetic substance.
zero
temperature
Curie
209
XI MAGNETS
The reason for the decrease in the magnetic strength is that with
rising temperature more and more of the electron moments begin to
point in the opposite direction, until at the Curie temperature there
are as many moments pointing one way as there are the opposite
way, and the net magnetic strength then becomes zero. This loss of
ferromagnetism on going through the Curie temperature is an
example of what is called a phase transformation. One gets the
ferromagnetic phase below the Curie temperature, and it transforms
into a non-magnetic phase on being heated above that temperature.
2. There is more order in the phase below To than in the one above
7^,. The actual type of order can be different in different cases:
crystalline order in melting, the ordered orientation of electron
magnetic moments in ferromagnetism.
hydrogen(l) +2.792743 i
carbon(13) +0.702381 i
oxygen(17) -1.89370 \
phosphorus(31) +1.13162 \
copper(63) +2.22664 \
XI MAGNETS
E=hf.
is doubled.
(a)
zero field
some field
S= 111
“Z
field doubled
212
oxygen
hydrogen
carbon
/ frequency
-4—»
(b)
magnetic field
213
XI MAGNETS
identical nuclei will set up its own distinctive group of energy levels,
different from that of any other nucleus. If I now shine photons of a
single frequency on the material and increase the magnetic field
gradually, I get a strong absorption from each type of nucleus at that
value of the field for which its energy spacing is equal to the photon
energy. I show in fig. XI-8, part (b), what the absorption might look
like for a mixture of two different types of atoms at fixed photon
frequency and changing magnetic field.
When they are all added up, most of them cancel one another out,
and what is finally left is a very weak net magnetic moment, much
smaller than what we find in a ferromagnet. All materials which are
not ferromagnetic show this kind of weak magnetism.
214
SUMMARY
7 Summary
I consider in this chapter the consequences of the magnetic
properties of electrons and nuclei. The one which is closest to
everyday experience is ferromagnetism, as for example in iron. I
describe in section 1 the basic properties of ferromagnets.
215
XII SUPERCONDUCTORS
l What is a superconductor?
If I now warm up the lead wire from the superconducting state, I find
that its resistance jumps back to its original value at the same
temperature at which the resistance had disappeared when cooling
the wire. I call this temperature the superconducting transition
temperature, and use the symbol 7^ (read T-nought) to denote it.
WHAT IS A SUPERCONDUCTOR?
io-lof—
10
-15 _
10
-20
11
resistance (ohms)
10
temperature (K)
100 200
entrance of the teacher. So the task is to find out what plays the role
of the teacher in superconductivity.
XII SUPERCONDUCTORS
tungsten 0.012
tin 3.7
lead 7.2
218
00
magnetic
600—
400-
2000-
^-
^™
^~
rr
0 12 3 4 5 6
temperature (kelvin)
219
XII SUPERCONDUCTORS
in gauss
tungsten
tin
lead
1
300
800
120,000
300,000
1,000,000
N
\
magnet S
superconductor –,,
(b)
tabletop
\\
\“
(c)
magnet
iron—
tabletop
22O
gauss. I shall refer to them as the low-field (LF) and the high-field
(HF) superconductors respectively. The reason for the different
behaviours lies in the details of how the superconductor reacts to a
magnetic field, as we shall see now.
question Can you identify one essential detail which is missing in fig.
XII-3?
XII SUPERCONDUCTORS
superconducting, N = normal.
(b)
large field
superconducting, M = mixed, N =
normal.
A
small field
(fl)
medium field
(P)
results made sense only if the gas consisted not of single oxygen
atoms, but rather of pairs of atoms bound together.
XII SUPERCONDUCTORS
million times smaller than in its normal state. The metal, although
electrically a superconductor, becomes thermally an insulator. Since
heat in metals is conducted primarily by electrons, it would appear
that though the electrons in a superconductor carry an electric
current with no resistance at all, they are strangely reluctant to carry
a heat current. The specific heat in the superconducting state also is
less than in the normal state.
3 Some clues
The explanation of how superconductivity comes about is one of the
most impressive achievements of physics. It represents a beautiful
bringing together of several aspects of the physics of solids that I
have described earlier in this book: the quantum mechanics of
electrons in a solid, phonons, the nature of electrical conduction,
magnetic effects.
3. The average energies of phonons are not too far from the
energies corresponding to the transition temperatures. One might
therefore wonder whether phonons should be somehow brought into
the picture. This thought is reinforced by the fact that the transition
temperatures of different isotopes of the same substance differ
slightly, and we know that the isotopes differ slightly in their phonons,
but not in their electronic structure.
4. The fact that electron pairs are the units which carry the
supercurrent suggests that we should look for something which will
be the glue, as it were, binding a pair of electrons together.
SOME CLUES
question With reference to the third point above, why should the
isotopes differ slightly in their phonons?
The phonons are the quanta of the vibrational energies of the atoms
in the solid, and these energies are different for different atomic
masses.
figure xn-6. Relative numbers of
energies.
tele
o-
<u
“1.
fl
<u
relati
l_
10
energy levels
15
20
15
20
energy levels
225
XII SUPERCONDUCTORS
4 Cooper pairs and BCS
The physicist Leon Cooper showed how a certain kind of interaction
between electrons can lead to the formation of electron pairs
composed of opposite momenta and spin, which have therefore
come to be called Cooper pairs. A Cooper pair is a peculiar object. It
is not two electrons somehow bound to each other so that they stay
close together and move about as a unit: an electronic molecule
consisting of just two electrons, so to speak, just as an oxygen
molecule consists of two closely bound oxygen atoms moving about.
In fact, it is not possible for two electrons by themselves to be bound
together: the electric force of repulsion between their charges would
make them fly apart. One needs a mediator between them to hold
them together in spite of this repulsion. An example of such a
mediator is the nucleus of the helium atom, which holds two
electrons bound close to each other. This is because each electron
is attracted to the positive charge of the nucleus, and this attraction
more than overcomes the repulsion between the electrons. I can talk
about an electron pair in the helium atom bound to each other, and it
takes a finite energy to break up the pair: this would be the energy
needed to remove one electron and leave behind a helium ion.
The mediator that holds the two electrons of a Cooper pair together
is a phonon, which is the quantum of energy of the thermal vibrations
of the atoms in the solid. I illustrate how this comes about in fig. XII-
8. Two electrons A and B approach each other travelling in opposite
directions with the Fermi velocity. Electron A emits a phonon and
moves off in a different direction. Electron B absorbs the phonon and
changes its direction of motion by the same amount as A, so that the
two electrons are still moving in opposite directions. The pair
remains intact, bound together, so to speak, by the phonon. This is a
Cooper pair. This process happens to all pairs of electrons moving
with opposite velocities all over the Fermi surface, and the result is
the superconducting state. Remembering that momentum is mass
times velocity, we can say that a Cooper pair is two electrons of
opposite momenta (and therefore of zero net momentum) bound
together by the emission and absorption of a phonon. The two
electrons also have opposite spins: this is established through
experiments and predicted by the theory.
We now have a picture of how Cooper pairs are formed through the
mediation of phonons. But there is more to it than just the formation
of pairs: it turns out that the 226
energy of the electrons in such a pair is less than the energy that the
electrons would have if they were moving independently. When
electron A creates and emits a phonon, what is happening is that it is
pulling towards itself the positive ions in its neighbourhood because
of the electric force between its negative charge and their positive
charges: you remember that a phonon represents a compressional
wave in the solid. The potential energy of the electron is therefore
lower than it would be if the ions were in their undisturbed positions.
Electron B, which arrives there half a cycle of the phonon later, sees
these ions being further from it than would otherwise be the case,
and therefore its energy is raised. However, this increase is smaller
than the decrease in energy of electron A.
The net energy of the Cooper pair is thus lower than that of the two
electrons when they are moving independently. The difference
between the two is the binding energy of the Cooper pair, and must
be supplied to it in order to break up the pair.
I have now described how one Cooper pair is formed. All the
electrons in the vicinity of the Fermi surface are paired up in this
manner at the absolute zero of 227
XII SUPERCONDUCTORS
The current is carried by the condensed Cooper pairs which are all
moving at the same velocity. Electrical resistance is caused by the
scattering of electrons moving independently of one another. Such
electrons can only be pulled out of the Cooper pair condensate when
at least a finite minimum amount of energy is supplied to it. This 228
there is a current.
normal metal
XII SUPERCONDUCTORS
number of electrons in a metal can influence one another through
the mediation of phonons and thereby form the superconducting
state. But its significance is far broader, because it showed a way of
looking in general at fermions which are interacting with one another.
It has been used, for example, to explain properties of atomic nuclei,
where the fermions are protons and neutrons, and of certain
astronomical objects called neutron stars, composed of very densely
packed neutrons.
5 Josephson oscillations
The physicist Brian Josephson used the BCS theory to predict an
effect one should see in superconductors. The effect was indeed
found experimentally, and is named after him. Two superconducting
pieces A and B are brought together with a thin insulating layer C
interposed between them as shown schematically in fig. XII-10, thus
forming a junction. A battery is connected to A and B, so that there is
a voltage of F volts between FIGURE XII-10. The Josephson effect.
photon 2eVlh
230
JOSEPHSON OSCILLATIONS
them. The layer C is thin enough that some current, even though
small, can flow through it. If the rest of the electrical circuitry (not
shown in the figure) is suitably set up, then photons (electromagnetic
waves) of frequency/= leVlh, where e is the electronic charge and h
is the Planck constant, are emitted by the junction. Given the
voltage, the frequency /depends only on two universal constants,
and not on which particular superconductor is used: a remarkable
result indeed. The effect follows purely from the existence of Cooper
pairs (hence the factor 2e in the formula for the frequency) and the
rules of quantum mechanics (and therefore the factor h).
E=hf.
wave
time
phase
angle
90
distance
180 270
231
XII SUPERCONDUCTORS
Now we look at what these properties mean for the junction shown in
fig. XII-10.
f= h ■
SUMMARY
233
XII SUPERCONDUCTORS
234
XIII CONCLUSION
CONCLUSION
find different numbers of heads and tails each time. There will
however be many more times with five heads and five tails than any
other possible combination. The larger the number of coins, the
more likely will be the result with equal numbers of heads and tails.
With the 1021 or so particles that we are concerned with in solids,
the most probable distribution becomes a practical certainty.
Physics explains why things are the way they are, in the realm of
condensed matter.
This however does not mean the end of condensed matter physics
as an active area of study and research; indeed, far from it. It is like
cooking; perhaps not too far-fetched an analogy, for an
experimenter’s laboratory is somewhat like a kitchen. One often
encounters physicists exchanging, or being secretive about, their
recipes, and it is not something to eat that they have in mind. The
analogy: the basic principles and techniques of cooking are well
established, but this by no means spells the end of the art (or
science) of cooking. Quite the contrary: there is no end in sight of
possible variations on the well-known or the invention of the new,
though the basic knowledge and techniques remain the same. So it
is in condensed matter physics too. The basic principles, as I have
described them in the preceding chapters, are well established, and
they give us a good explanation of material properties. But what is
not now, and may never be, possible is to predict exactly all the
properties of a given material. For example, you might give me a
new material, tell me its composition and crystal structure, and ask
me if it is going to be superconducting. My answer will be ‘Perhaps;
but then again, perhaps not.’ The reason for such indecision is that
one has a very large number of particles interacting with one
another, and one can make a theory of its properties only after
making approximations. The problem otherwise will be too
complicated to solve, even though we know how to formulate it. It
could be that a particular approximation causes something to be
ignored which was precisely what would have made your new
material become a superconductor. One of the skills a physicist
needs in this field, as she approximates, is to know what to keep and
how 236
CONCLUSION
much of it, and what to throw away — rather like a cook in his
kitchen. A related question is, can we design a material so that it has
specified properties, for example a superconducting transition at 300
K? The answer at best can only be ‘Not yet.’
237
GLOSSARY
238
GLOSSARY
Bloch electron
Bohr magneton
Boltzmann constant
boson
Brillouin zone
brittleness
bulk properties
c.g.s. units
cc
charge
chip
conduction electrons
conservation of energy
conservation of momentum
The name for particles which obey Bose-Einstein statistics; the wave
function for a system of bosons is symmetric, i.e. it is unchanged
when two particles are interchanged. Examples are helium atoms,
photons and phonons.
respectively.
Abbreviation of ‘cubic centimetre’; also called a millilitre.
The physics of solids and liquids, where the atoms are so close to
one another that they cannot be treated as independent particles.
A fundamental law of nature: the total linear momentum and the total
angular momentum of a system, isolated from the outside, each
remains the same for all time.
239
GLOSSARY
continuous spectrum
Cooper pair
cosmological principle
coulomb
critical current
crystal
crystal structure
Curie temperature
current
current density
degeneracy
density
diffraction
dimensions of space
diode
dislocation
dispersion
Light has such a spectrum if all frequencies within some range are
present in it; sunlight has a continuous spectrum.
The unit for measuring electric charge, equal to the total charge of
about 6 X1018 electrons.
The maximum electric current that a superconductor can carry and
still remain a superconductor with zero electrical resistance.
The density of something is the amount of that thing per unit volume,
usually taken to be one cc. Thus the density of electrons is the
number of electrons per cc. One also says the density of silver to
mean the mass of silver per cc of volume. The context makes clear
which sense is meant.
The fact that the speed of light in a substance is different for different
frequencies.
240
GLOSSARY
dispersion curve
doped semiconductor
elastic limit
elasticity
electric charge
electric field
electric neutrality
electromagnet
electromagnetic radiation
electron
electron-volt
element
elementary particle physics
energy
energy level
equation
erg
Same as charge.
241
GLOSSARY
excited state
exponential notation
Fermi energy
Fermi momentum
Fermi surface
fermion
ferromagnetic material
ferromagnetism
field
force
formula
fracture
frequency
function
242
GLOSSARY
gauss
glass fibres
glassy solid
grain
gravitation
ground state
hertz
hole
infinity
insulator
interaction
interference
Josephson oscillations
joule
kinetic energy
lattice
lattice parameters
examples are the mass and electric charge of the electron, the
speed of light in free space, the Planck constant, the Boltzmann
constant, the Bohr magneton.
Most solids are made up of tiny crystals called grains which are
stuck together.
superconducting elements like tin or lead can stand before they lose
their superconductivity.
The influence which one body exerts on another. The origin of ■the
interaction could be their masses, in which case the interaction is
gravitational. If it is their electric charges and magnetic moments,
then the interaction is called
electromagnetic.
243
GLOSSARY
lattice site
lattice spacing
line spectrum
linear momentum
macroscopic
magnetic domain
magnetic field
magnetic moment
magnetic monopole
mass
matter
metal
microscopic
mode of vibration
molecule
momentum
MOSFET
MRI
A point in a lattice.
The distance between two adjacent points in a lattice.
Light with a line spectrum has only certain frequencies in it, and not
others. Light from a neon tube is an example.
Stuff which has mass; the difference between matter and energy has
become a bit blurred since Einstein’s E = mc2.
On an atomic scale.
244
GLOSSARY
n-type semiconductor
neutrino
neutron
node
normal state
nuclear fission
nuclear magnetic resonance
nuclear magneton
nucleus
ohm
operation
order
p-n junction
p-type semiconductor
particle
persistent current
What is left of an atom if all its electrons are stripped off. The
nucleus is made up of protons and neutrons, contains most of the
atom’s mass, and is about 10~I3cm across.
245
GLOSSARY
phase transformation
phonon
photon
Planck constant
plastic deformation
polarization
polarized light
polarizer
polycrystal
position vector
positron
potential energy
precipitate hardening
probability
A wave function oscillates in time as well as in space. The oscillating
amplitude at any instant and point can be related to the angle swept
by a body rotating at the same rate. This angle is the phase of the
wave function.
A beam of light in which the electric field points in the same direction
everywhere.
This is the usual form of most solids, consisting of tiny crystals called
grains which are held together by the forces between atoms from
adjacent grains..
A particle which is exactly like the electron in all respects except one:
its electric charge is positive and equal in magnitude to the charge of
the electron.
GLOSSARY
proton
quantization
quantum mechanics
quantum number
quantum state
quark
radiation
radioactivity
reflection symmetry
refraction
resistance
rotation symmetry
scalar
scattering time
Schroedinger equation
scientific notation
Quantities like energy and momentum for an object can take only
certain discrete values, and nothing in between. This is called the
quantization of energy, momentum, etc.
Any one of the possible stable states in which an object can find
itself, with an energy, a momentum, etc which are specific to that
state.
The name of the particles which make up protons and neutrons.
It is used to calculate the wave function, the energy levels and other
quantities of interest.
GLOSSARY
second sound
semiconductor
sintering
solar battery
solution hardening
specific heat
spectroscope
standing wave
state
statistical physics
statistics
superconducting transition
temperature
superconductivity
superconductor
surface properties
The passage of a temperature pulse along a piece of matter with a
speed somewhat less than the speed of sound; can occur when the
mean free path of the phonons is very long.
A solid with a modest energy gap between the highest fully occupied
band (the valence band) and the lowest empty band (the conduction
band), so that the effective numbers of electrons and holes can be
sensitively controlled through temperature and doping.
temperature.
248
GLOSSARY
symmetry
thermal conduction
thermal conductivity
thermal energy
thermal expansion
thermal radiation
transistor
translation
translation symmetry
travelling wave
unit cell
unpolarized light
valence band
variable
vector
velocity
volt
watt
The flow of energy in the form of heat from the hot part to the cold
part of a piece of matter.
The amount of thermal energy flowing per second from one face to
the opposite face of a one-cm cube, when the temperature
difference between the two is a degree Celsius. It is a characteristic
property of a given substance.
This is the symmetry an object has if it looks the same after being
moved from here to there through a specified distance without
rotation.
The usual wave, in which one sees the crests and troughs actually
going somewhere.
The band of highest energy levels for electrons in a solid in which all
the states are occupied by electrons. The next higher band is either
empty or only partially occupied by electrons, and is called the
conduction band.
The unit for measuring power, which is the rate at which energy is
used or produced; equal to one joule per second.
249
GLOSSARY
deformation.
wavelength
Wigner-Seitz cell
250
INDEX
absolute zero, 91
angular momentum, 38
atom, 75
conservation, 41
nucleus, 211
quantized, 76
annealing, 133
atoms
chlorine, no
electronic structure, 77
impenetrability, 79—80
isotopes, 81
misleading representation, 66
silver, 104-106,109
size, 79
sodium, no
glass, 174
semiconductors, 195
Boltzmann constant, 91
boson, 96
distribution, 100
spin, 96
brittleness, 134
bulk properties, 15
chip, 2,197
close-packed direction, 23
conservation principles
energy, 41
momentum, 41
cosmological principle, 13
one-dimensional lattice, 15
three-dimensional lattice, 29
two-dimensional lattice, 17
crystal
one-dimensional, 14—15
shape, 23, 29
silver, 8
electrons, 108—112
three-dimensional, 27—29
two-dimensional, 16—19
crystal structures
diamond, 112
face-centred cubic, 30
graphite, 31
chloride), 31
silicon, 112
diffraction, 46
dislocations, 136-141
dispersion curve
electrons, 116
phonons, 126,132
prism, 184
rainbow, 183
distribution
atomic speeds in helium gas, 87
bosons, 100
fermions, 98
phonons, 128
populations by age, 84
electric charge, 63
electron, 64
potential energy, 65
proton, 64
electrical resistance
definition, 190
effect of alloying, 192
insulators, 196
metals, 194
resistivity, 190
semiconductors, 195
electromagnet, 208
electron
mass, 37
spin, 36, 76
energy
and frequency, 51
definition, 38
kinetic, 38
potential, 38, 65
251
INDEX
energy (cont.)
total, 39
zero-point, 58, 92
energy bands
crystal, 106,114—117
insulators, 107-108,117
metals, 107-108,118
semiconductors, 120—123
energy levels
excited state, 71
exponential notation, 3-4
fermion, 96
distribution, 98
Pauli principle, 97
spin, 96
ferromagnets, 205—210
domains, 206—208
field, 165-167
electric, 167-168
magnetic, 168-169
scalar, 167
temperature, 166
vector, 167
velocity, 167
fracture, 141-143
fundamental constants, 52
glassy solid, 4, 32
fracture, 141-143
preparation, 10,12
grains
in polycrystalline solid, 10
ofsand, 2
gravitation, 2
ground state, 67
hole, 120,123
hydrogen atom
energy levels, 72
excited states, 71
quantum numbersv72-77
stability, 66
insulators, 107
interference, 45
isotopes, 81
lattice, 15
parameters, 17
light
absorption, 173-180
fat, 179
glass, 174
metals, 175-176
water, 179—180
dispersion, 182-184
generation, 171—173
refraction, 180-182
linear momentum, 37
conservation, 41
wave vector, 51
magnetic moment
electron, 204
nuclei, 211
magnets
permanent, 207—208
phonons, 150
MOSFET, 197
motion, 35
MRI, 214
in a MOSFET, 197
in a MOSFET, 197
particles
comparison to waves, 46
definition, 36
duality, 49, 50
energy, 38
light, 49
momentum, 37
properties, 37-40
the electron, 47
electrons in atoms, 78
phonon, 124—128
distribution, 128
252
INDEX
energy, 127
photon, 50
. distribution in energy, 94
energy, 51
momentum, 51
Planck constant, 52
polarization
atoms, 180
light, 171
polycrystalline solid, 10
probability
electron, 76
nuclei, 211—214
quantization of energy
phonons, 127
photons, 51
quantum mechanics, 49
some formulas, 51
rainbow, 183
reflection symmetry, 25
scalar, 37
scattering
Schroedinger equation, 59
scientific notation, 3-4
semiconductor
doped, 121—122
n-type, 123
p-type, 123
sintering, 208
specific heat
definition, 147
superconductor, 225
spectroscope, 184
standing waves
spin
electron, 36, 76
metal, 134
silicon, 135
superconductors, 216—234
magnets, 232
Meissner effect, 221
surface properties, 15
symmetry, 19
reflection, 25
rotational, 23-24
translational, 21-22, 29
temperature, 86
absolute zero, 91
energy, 89
thermal conductivity
alloys, 155
insulators, 151—153
metals, 155
thermal energy, 9t
transistor, 2,197
translation, 21
unit cell
three-dimensional lattice, 28
two-dimensional lattice, 18
vector, 37
addition, 40
velocity, 37
virtual springs, 13
waves, 36, 42
amplitude, 43, 53
comparison to particles, 46
diffraction, 46
duality, 49, 50
electromagnetic, 170-171
thermal radiation
frequency, 43
interference, 45
matter, 58
sound, 125
standing, 55
travelling, 55
253