You are on page 1of 173

ATOMIC STRUCTURE

OF MATTER
JANUARY 2014

Copyright

2014
Copyright is held by Prof Martin Dove, Queen Mary University of
London
i

Acknowledgements

In the end this book will be published with full accreditation to all involved. At this time, these people are Anthony Phillips, Serena Maugeri
and Min Gao from Queen Mary University of London, and David Palmer and Mike Conley of CrystalMaker Software Ltd. The movies and
crystal structure images were produced using a development version of
the CrystalMaker visualisation software. Funding to support some of the
work was provided by a grant from Queen Mary University of London.

ii

Preface

This book is an attempt to show the world of materials down at the


scale of the atom, exploiting the capabilities of digital media.
Databases of crystal structures contain hundreds of thousands of examples of different materials, both inorganic and organic, and new crystal
structures are being added to the databases daily through the ingenuity
of organic chemists and crystal engineers.
We need to know about the atomic structure of materials because
much of modern life depends on the successful exploitation of the properties of materials. Modern societies depend on a number of key areas,
including information and digital technologies, energy generation and
storage, agriculture, medicine and pharmaceutics, and finance and retail. Materials underpin many of these.
If we take the simple example of mobile IT devices, we find advanced
materials throughout: the electrodes and the ionic conducting electrolyte of the battery; the silicon of the microprocessor (the chip) and
the flash memory, and the ceramics used within the components of the
electronics; the optoelectronic materials used as the sensor in the camera and in the electronic camera flash; the transducers used in the microphone and small loudspeakers; the light-emitting diode materials of

the screen (which can be inorganic or organic materials); the transparent conductors used to give the screen its response to touch; the thin
strong glass front cover; and even the lightweight alloys used for the
case.

iii

We have had the techniques to understand the atomic structure of materials since the pioneering work of Henry and Lawrence Bragg in developing the x-ray diffraction method in 1913. It might of course be argued that mankind developed a detailed knowledge of metalwork
through the bronze and iron ages and past the industrial revolution
without any knowledge of the atomic structure of metals, and there are
other examples of materials that were successfully exploited without
any understanding of their structure (one example is the active component of polaroid materials, which was only understood very recently
and 150 years since their first use), but we now ask so much more of
modern materials in terms of the microscopic size of components, the
speeds of which we expect responses, and the robustness required to
see millions of components work together reliable for long periods of
time. And it is because we ask so much of our materials that we need
to have good materials design, and good design needs a good knowledge of material structures.
This book is about the atomic structure of materials, giving the reader
the opportunity to observe and experiment.

iv

CHAPTER 1

Welcome

Forward to chapter

This chapter is our Hello to you.


The book is about the atomic structure of materials. Most of the materials we are concerned with are crystalline materials, and in this chapter
we explore some of the reasons why we believe crystal structures have
an underlying atomic structure. We look at both macroscopic observations as well as some of the microscopic observations. We are not in a
position to actually see atoms, but there is a lot we can learn in a general overview.
In this chapter we also deal with length and time scales. These are well
outside our human experience that they can just appear to be numbers, but it is worth trying to gain an appreciation for these scales, for
example by comparison with other quantities.

vi

Introduction
Preamble
This book is concerned with the structure of matter at the atomic
level. Nature has given us 84 stable (or virtually stable) atom types,
with a range of size, valence, and shape of electronic orbitals, giving rise to different ways of bonding with other atoms, which leads
to the rich diversity of material structures and properties that are either found naturally or can be synthesised. Understanding the
atomic structure of matter, and how its properties are determined by
the structure, is as much as a visual task as it is a matter of writing
equations. This book has been created specifically to help address
7

INTERACTIVE 1.1 Example of a crystalline solid, GaAs. This shows


the crystal structure as an example of what we will be exploring in
this book. You can rotate into different orientations to see how the
atomic structure is repeated along different directions. The s
MOVIE 1.1 Animation of atoms in a fluid. The key point is that the
atoms are moving in two ways. First the atoms move back and forth
within cages formed by the atoms in the local environment. Second
over a longer period of time the atoms are able to diffuse
throughout the fluid sample. One atom has been coloured red to
help it be followed by eye it as it moves around.
this issue of the visual understanding of the atomic structure of matter.

In this book you will find animations and images that are impossible to
reproduce within the traditional book format, and usually only possible
to access with specialised software or via web tools. These will give an
unparalleled perspective on the atomic structure of matter.

The states of matter


At school we learn that there are three states of matter: solid, liquid
and gas. The chief characteristics are to do with size and shape, in that
8

a solid has both size and shape, a liquid has size but no fixed shape (a
liquid flows (Movie 1.1 shows the atomic motions within a liquid), and
a gas has neither shape nor size (a gas expands to fill the space it is in).
Inevitably, this picture is rather a gross simplification, but it does at
least open up a basic characterisation.
We will mostly be concerned with the solid state. We will mostly be
concerned with solids that are crystalline in form, but we will also discuss glasses. We should quickly dispel some myths regarding the solid
state that might be implied by this basic characterisation, namely that
solids are static ordered objects. We will see in this book that some solids are indeed well ordered, but even in these cases the atoms are vibrating back and forth around their average positions, with their kinetic
energy corresponding to the temperature of the material. In some cases
the motions can be quite large, leading to some degree of structural
disorder. Molecules can tumble, some atoms can form on a fluid subphase within the structure, and glasses form with significant topological disorder. Moreover, even in the most ordered structures, imperfections known as defects can arise through the crystal growth phase,
as a result of subsequent handling, and indeed spontaneously due to
the atomic motions.
But the foundation to understanding the solid state is indeed the perfectly ordered crystalline phase, where atoms are arranged in neat patterns that repeat almost indefinitely throughout three-dimensional
space. To understand crystal structures requires an appreciation of symmetry and chemical bonding, both of which have a significant visual
component.

GALLERY 1.1 Examples of crystal shapes


Crystals of iron pyrites (otherwise known as Fools Gold), FeS2

Clues from crystal shapes


Long before the existence of atoms was accepted by the scientific community, it was understood that the regular shapes of atoms that exist independent of the size of the crystal probably implied some fundamental microscopic building block that is repeated within the crystal from
face to face. Indeed, this idea dates back at least as far as Indian and
Greek philosophers of the 6th and 5th centuries BCE respectively, and
was revived in the West in the 17th century. Examples of crystal shapes
9

GALLERY 1.2 Photographs of snowflakes


Snowflake photograph, taken by Alexey Kljatov and downloaded
from http://www.flickr.com/photos/chaoticmind75/.

are shown in Gallery 1.1. These crystals were all produced naturally
within the Earth's crust, either as a result of slowly growing from solutions or from the molten state.
The mysterious shapes of snowflakes, which always have a hexagonal
symmetry, must arise from some internal structure of the ice crystals in
exactly the same way that the more regular shapes of naturally grown
crystals reflect an inner structure.
Iron-rich meteorites have a composition that is primarily iron (majority
phase) and nickel. As the meteorite was cooled, the metal phase sepa-

FIGURE 1.1 Widmansttten pattern of an iron-rich meteorite, formed


by the exsolution of two iron-nickel phases of different crystal
structure.
rated into two other phases, one that is low in nickel content called
kamacite, and one that is richer in nickel called taenite. This is because
intermediate composition are unstable. These two phases have different
crystal structures; kamacite has a structure called body-centred cubic,
and taenite has one called cubic close packed. We will meet both structures in the next chapter. These structures match well only with a welldefined relative orientation, and this gives rise to characteristic patterns
seen on a polished surface called Widmansttten patterns as
shown in Figure 1.1. We will explain the origin of these patterns in the
next chapter.
10

FIGURE 1.2 Double refraction of a calcite crystal.

Some simple experimental observations


Calcite: double refraction
We can perform simple experiments to understand more about the internal structure of crystals before we get towards the atomic state. Let
us take a crystal of calcite (chemical formula CaCO3; calcite is the primary component of limestone and marble, and nature produces some
lovely clear crystals for us to play with). The most striking thing about
this crystal, aside from its perfect shape, is that when we look through
the crystal at an image the other side, we actually see two copies of the
image. This is seen in Figure 1.2. When we rotate the crystal, we find
that the double images rotate about each other.
This observation suggests that there is a set of internal axes that rotate
as you rotate the crystal, in a clear relationship with the faces of the
crystal.

INTERACTIVE 1.2 Crystal structure of the polarising material


herapathite, the active material often used in polarising sheets. The
key component is the chain of iodine atoms (purple), embedded
within the organic component (black is carbon, red is oxygen, blue
is nitrogen). The molecule has been known for over 150 years, but
its structure was not solved until 2009 (Kahr et al, Science, 324,
1407, 2009)

Optical birefringence
A second experiment on this crystal of calcite or indeed on many
crystals can be performed using two sheets of polaroid film. Light
propagates as vibrating electrical and magnetic fields, and the polarisation (that is, the direction in which the vibrating electrical field is generated) can lie in any direction perpendicular to the direction in which
the light wave is travelling. A polaroid sheet is typically made from
11

FIGURE 1.4 Generation of slip planes within a crystal under tensile


strain.
FIGURE 1.3 Slips lines in a single crystal wide of cadmium.
needle-shaped crystals that polarise light, held within a polymer matrix, as seen in Interactive 1.2. The polaroid sheet will transmit light in
which the its polarisation is exactly parallel to the alignment of the polarising molecule. Any polarisation vector lying at an angle to the polarisation direction can be be attenuated by a factor of sin2 . Thus the
light is extinguished when = 90 , but the parallel component can
propagate to some extent when 0 . Placing a second polaroid sheet
with a perpendicular polarisation above the sheet will extinguish all
light. But when we place our crystal of calcite between the sheets, we
find that generally light is able to pass through the crystal, except for
certain alignments. It appears that the crystal is able to rotate the polarisation of light passing through the first polaroid sheet so that it is not
extinguished by the second sheet, except for special cases in which it

appears as if some internal axis of the crystal is aligned with the polarising direction of the polaroid sheets.

Stretching cadmium single crystal wires


It is possible to grow single crystals of cadmium in the form of wires.
An experiment in which this wire is stretched has an interesting result.
It is possible to stretch the wire to more than double its original length
before it breaks. And when it stretches, its shape transforms from a wire
with a circular cross section to a thin ribbon, with the ribbon width of
more-or-less the same as the diameter of the wire. A view of a ribbon
seen in an optical microscope is shown in Figure 1.3. It can seen that
the surface now consists of stripes, that appear to represent disks cut
from out of the original wide and now placed on their side. The obvious representation of this process is shown in Figure 1.4. Here we have

12

represented the wide as consisting of platelets, which slide over each


other and rotate as the response to the tensile stress.
What this observation does is point to the same sort of internal stsructures that are suggested by the other experiments and observations. Indeed, all three experiments are consistent with the inference from the
observation of regular shapes that crystals consist of an internal fundamental repeating substructure, with the repeat distances being related
to the axes of the crystal that are seen in the shape and also in the experiments with polaroid sheets. And at this stage we have not yet
thought about the atomic structure.

13

Scale
Length scales
We operate on the length scale of the metre. The metre is a bit
longer than our walking stride, but we can stretch our arms a little
bit wider than one metre. It might be a walking journey of about
one thousand metres (1 km) to visit the local shops, and a car or
train journey of up to a few hundred km to visit our friends and families. The subjects of geography and geology take us into the realm of
thousands of km, and astronomy takes us way beyond such distances. Going downwards in size we start to struggle with objects of
the size of a millimetre or so (such as the screw that hold my spectacles together). The thickness of paper or the width of a human hair
14

takes us down to some hundredths of a millimetre, which is now expressed in terms of tens of micrometers (one micrometre, written as 1
m, is equal to 1 millionth of a metre, 106 m). Fine powders have
grain sizes that go below 1 m, but by now we have reached the limits
of the human eye. Optical microscopes allows us to see down to a fraction of a m, but we then touch the limits of visible light, with wavelength of order 0.5 m.
We often hear about nanotechnology and nanoparticles. The nanometre is one thousandth of a 1 m, or 109 m. Viruses have sizes of upwards of 20 nm (rather smaller than bacteria). Inorganic nanoparticles
(examples such as gold, germanium and silicon dioxide) can be grown
to sizes down to 10 nm. DNA molecules are about 2.5 nm wide. The
development of fast computers depends on being able to produce silicon circuits that are as thin as 25 nm.
Clearly the length scale of the atom must be smaller than the
nanoscale. In fact it is down by one order of magnitude, namely about
0.1 m or 1010 m. This length scale is sufficiently special that it has its
own unit, the ngstrom, namely 1 = 1010 m, and this unit is permitted under the SI system. It is the unit that is routinely used in the science of atomic structures of materials, and is the unit that will be used
here.
For further reading, Wikipedia has a nice article comparing length
scales from the very large to the very small.

Time scales
Our natural time scale is 1 second. This is the unit we count down the
start of a race, for example, and is about the time it takes us to stride 1
m. A slow heart will beat at a rate of once per second. The human

body however performs a number of tasks that are much faster. We


take about 0.1 s to blink, we can respond to an external stimulus after
about 0.02 s, and we our brain neurons fire over a time of around one
thousandth of a second, namely 1 millisecond or ms. A sound wave in
air will travel 1m in 3 ms. We hear sound that oscillates once ever 50
ms down to once every 0.05 ms.
But we tend not to think of wave in terms of the time taken to oscillate
once, but in terms of the inverse of time, namely the frequency. Thus 1
Hz, defined as one oscillation per second, is the standard unit of frequency, and thus typically we hear sound between 20 Hz and 20 kHz
(1 kHz = 103 Hz).
Radio waves have frequencies in the range 30300 kHz (long wave),
300 kHz to 3 MHz (medium wave or AM radio), and 30300 MHz (FM
radio). Microwaves have frequencies up to 300 GHz (1 GHz = 109
Hz), infrared radiation has frequencies from 300 GHz up to 450 THz (1
THz = 1012 Hz), and visible light though to x-rays through to gamma
rays takes us to higher frequencies. This range takes us through and byond the time scales associated with the dynamics of atoms inside materials and beyond, and is the frequency scale associated with the radiations that are used to measure the time scales of atomic dynamics.
The key frequency ranges for the vibrations of atoms are up to around
40 THz, which is the frequency of vibration of the SiO bond in a silicate material such as quartz. CH stretching frequencies are of order
90 THz and OH stretching frequencies are of order 108 THz. But
most vibrations in materials extend to a few THz, and frequently the
interesting vibrational dynamics are around 1 THz. The corresponding time scale is 1012 s, defined as 1 picosecond (1 ps).

15

Vibrations are not the only dynamics of interest. Materials will have
missing atoms and impurity atoms. The time scale for atoms to hop into
vacant sites or for impurity atoms to hop between neighbouring sites is
typically of the order of 1 s, or faster or slower by powers of 10 depending on temperature and the height of the potential energy barriers.
Thus we can see dynamical processes associated with large-scale
atomic motions diffusion on the frequency range from kHz to GHz.
Again, Wikipedia has a useful article on time scales from the very slow
to the very fast.

The importance of appreciating length and


time scales
In this book we will show you three-dimensional images of atoms
within crystals that you can manipulate, and movies showing atomic
vibrations. You will see these on the lengths scales of centimetres and
the time scales of seconds. But in you mind you must transport yourself
to the true length scales of 1 , and to the true time scales of 1 ps,
even though these are well outside human experience.

16

Seeing atoms
It should be clear from the preceding section that the world of the
atom is outside what we might expect to see with any optic device.
The wavelength of visible light is so much larger than the size of the
atom that there is no scope to build a microscope that will allow us
to focus down to the atom. Howev

Interaction of x-rays and crystals


At the time of writing, we are approaching the centanary anniversary of discoveries that have transformed modern life through enabling us to deduce the atomic structure of matter. In 1912 two
teams, namely those of Max von Laue in Munich and the English/
17

Australian father and son team William (Leeds) and Lawrence (Cambridge) Bragg showed that beams of x-rays can be scattered by crystals
through a process known as diffraction, because the wavelength of xrays (a form of electromagnetic radiation like light, but with much
shorter wavelength) is similar to the distances between bonded atoms.
For example, a common source of x-rays in the laboratory is from a
copper anode, giving x-ray beams with a wavelength of 1.54 . The
first experiments with x-rays used photographic film to record the xrays scattered by the crystal, finding that the scattered beams were only
found in well-defined directions. These researchers deduced the mathematical relationships that showed how the scattering in these special
directions is related to the internal structure of the crystal, thereby proving for the first time that crystals do have an internal periodic structure.
von Laue was awarded the Nobel prize for Physics in 1914, and the
Braggs were awarded the prize the following year (see the Appendix).
X-ray diffraction, as the technique is now called, was quickly refined.
The theory was developed to enable the intensities of reflections to be
related to the positions of the atoms, and methods established to allow
researchers to deduce the atomic structure of crystals from the completed diffraction data. Nowadays this method is routine for most materials indeed, it is now operated as a service rather than as an intellectual effort for most materials and materials as complex as crystals of
proteins and virus can be tackled using the method. Indeed, x-ray diffraction was the crucial technique for deducing the atomic structure of
DNA, and that was only 40 years after the discoveries of von Laue and
the Braggs.
Figure 1.4 shows an example of the diffraction of x-rays by a crystal as
recorded on a x-ray photographic film. The primary feature (admittedly
enabled through a particular way of recording the photograph that re-

FIGURE 1.5 Photograph of the pattern of diffraction of x-rays from a


crystal. The individual points represent reflection of the x-rays of a
single wavelength from distinct layers of atoms; the lattice
arrangement of these points reflects the underlying periodic
symmetry of the arrange of atoms in the crystal. The streaks arise
from a continuous spectrum of x-rays wavelengths in the beam.
moves all distortions from the geometry of the experiment) is of an array of peaks that reflects an underlying lattice structure within the crystal: the same lattice structure that is reflected in the shapes of crystals.
18

X-rays, however, do not see atoms. The way they are scattered by a crystal is directly determined by the arrangement of atoms within a crystal,
which allows us to deduce the atomic structure of a crystal from x-ray
diffraction data but only through subsequent analysis.

Electron microscope
The electron microscope exploits the wave nature of electrons. Electrons accelerated to energies of 10 keV will have a wavelength of
0.12 , and shorter wavelengths correspond to higher energies than
can be generated. Because electrons are charged particles they will interact with the electrons and the electric fields within the crystal,
which gives rise to the diffraction but also leads to loss of energy of the
beam. Thus experiments with electron beams are performed on very
thin samples.
The electron beam is diffracted in the same way that x-rays are diffracted, but because the wavelength is so small the angle of diffraction
is much smaller. This allows the use of electromagnetic fields acting as
lenses to recombine the scattered beams to create an image, performing the mechanical analogue of the analysis of x-ray diffraction. Thus it
is possible to directly record an image of the atoms in the crystal. Well,
it is not the sort of image we might see with light, because atoms will
scatter x-rays in complicated ways that means that features in the image will not necessarily correspond to actual atoms. One example of
when we can associate atoms with features is shown in Figure 1.5,
which shows an image of crystalline silicon. Here the blobs represent
actual atoms, as is confirmed by the inset showing the crystal structure
modelling by software.

FIGURE 1.6 Transmission electron diffraction of the atoms in


crystalline silicon. The top right shows the corresponding view of
the crystal structure of silicon drawn from software.

19

CHAPTER 2

Crystal structures of the elements

Forward to chapter

In this chapter we will introduce the crystal structures of the elements.


We will start with the metals because in many cases their crystal structures are the simplest. Through a discussion of the crystal structures of
elements we will introduce a number of formal concepts that we need
to understand. These are primarily concerned with quantifying the symmetry of crystals.
The crystal structures of many metals are based on a close packing of
spherical atoms. We therefore begin by considering the packing of
spheres in two dimensions, and from this we introduce some of the key
concepts in symmetry. These two-dimensional packings of spheres are
then stacked together to form the three-dimensional structures found in
nature.

xxi

Two-dimensional packing of atomic


spheres
Close packing arrangement of spheres
The simpler crystal structures are best described as the packing of
spherical atoms. Some elements have crystal structures that are the
tightest packing possible what is called close packing and we
will explore these cases shortly. But we will quickly find that we
need to introduce some formal concepts, so we start with the simpler example of a two-dimensional crystal of spherical atoms. The
two-dimensional close packing of spheres is shown in interactive
graphic 2.1, where we see that this packing is accomplished
through the formation of a highly-symmetric arrangement of the atoms.
22

INTERACTIVE 2.1 Close-packed arrangement of atoms within a


single two-dimensional layer.

Symmetry of the close-packed two


dimensional system
The immediate impression that is given by this image is the symmetry
of the arrangement, which can be described by defining two types of
symmetry.
The first type of symmetry concerns that of the environment of any
point in the structure; this is called, for obvious reasons, point symme-

GALLERY 2.1 Symmetry operations for the 2d close-packed structure


Two types of mirror planes, shown in different colours. We highlight
the rotational and translational symmetry copies. Reflection of the
structure in any of these planes brings the structure into coincidence
with itself.

try. Gallery 2.1 shows how it is possible to describe some of the point
symmetry elements:
Considering an axis coming out of the plane of atoms and passing
through an atom, it is clear that rotation of the plane by (n /6) 360,
23

where n is any integer, will bring the whole structure into coincidence with itself. This type of symmetry is called rotational symmetry, and the fact that we can rotate by 1/6th of a revolution means
that this this is called 6-fold rotational symmetry.
There are other points in the plane where there is a different type of
rotational symmetry. The point between three atoms defines the position of the axis of 3-fold rotational symmetry, where we can rotate
the structure by any (n /3) 360 to bring it back into coincidence
with itself. The point where two atoms touch defines the position of
the axis of 2-fold rotational symmetry, where now the rotation is
(n /2) 360.
If we draw an infinite line that passes through the centres of two connected atoms, we see that this line defines a mirror such that the arrangement of atoms to the left of the line is a mirror image of the atoms to the right; this is called mirror symmetry. Identical lines can be
drawn at 60 rotations, reflecting the 6-fold rotational symmetry.
There is an additional set of mirror planes at right angles to this initial
set, passing through one atom and also through the point where two
neighbouring atoms touch.
More information on point symmetry is provided in Appendix A.

Translational symmetry and the crystal


lattice
The second type of symmetry concerns the way in which the structure
can be brought into coincidence with itself by a translation of the
whole structure. This is called, again for obvious reasons, translational
symmetry.

FIGURE 2.1 Definition of the lattice vectors of the two-dimensional


close-packed structure. In this structure with hexagonal symmetry,
the angle between vectors is constrained to equal 120.
Translational symmetry allows us to define the concept called the crystal lattice. This is defined as the infinite set of vectors by which you can
translate the crystal and bring it back into coincidence with itself. It is a
regular array of points. For a two-dimensional lattice, the lattice is defined completely by two vectors, and all other points in the lattice can
be obtained as linear combinations of these two vectors. It is conventional, however, not to explicitly define these vectors, in part because
24

we need to allow for the lattice to be both rotated and translated without losing the definition of the lattice. Thus the lattice is more usually
described in terms of the lengths of these two vectors and the angle between them. This is illustrated in Figure 2.1.
The two vectors of the lattice actually define a repeating volume (or
area in two dimensions) of the crystal, as shown in Gallery 2.2. This volume is called the unit cell, and is the fundamental repeating entity. In
our two-dimensional example, the unit cell contains just one atom, but
in most cases the unit cell will contain several atoms; when we generalise to three dimensions shortly, our first example will contain 2 atoms
in the unit cell.
The dimensions of the unit cell, namely the lengths of the edges of the
unit cell (that is, the lengths of the vectors that define the lattice) and
the angle(s) between the vectors, are collectively called the lattice parameters. In three dimensions, the three unit cell lengths are given the
symbols a, b and c, and the three angles have the symbols , and .
The relationship between these parameters, and particularly how the
angles are defined in terms of the three vectors (e.g. is the angle between the vectors of length b and c), is shown in Figure 2.1. The vectors that define the lattice are denoted as a, b and c (but see Box 2.1).
The symmetry of the unit cell defined for the two-dimensional array of
spheres is reflected in the values of the lattice parameters, namely that
we have a = b and = 120 (by convention the obtuse angle is chosen).
Sometimes it is possible to define a unit cell that is twice as large as the
smallest possible unit cell but which has unit cell angles that are 90. It
is possible to define such a unit cell for our two-dimensional arrangement, as seen in on the imagines in Gallery 2.2. In this case, we have

GALLERY 2.2 Definition of the unit cell


The simple parallelogram defines the area of the repeating unit, and
here we show four stacked together to illustrate how this unit can be
repeated in space. The area encompasses the space occupied by a single
atom, and is defined such that the atom lies at the origin and on all
vertices.

a=

3b. There is now a lattice point in the centre of the unit cell;
25

Box 2.1
The use of the lattice vectors a, b and c follows the standard
practice of people who study crystal structures, but other communities prefer a convention that might appear more logical,
namely the set of vectors a1, a 2 and a3. However, the logic
doesnt survive the attempt to name the angles between the vectors. Thus this convention is often used within Physics text
books where much of the discussion concerns simple crystals of
metals, where it is easy to avoid having to think about the unit
cell angles.

when this approach is taken in three dimensions, the additional lattice


points (there may be 1, 2 or 3 additional lattice points) may lie within
the unit cell or on one or more faces of the unit cell. The smallest unit
cell that can be defines is called the primitive unit cell, and the larger
one chosen to have lattice parameters that reflect the symmetry is
called the conventional unit cell.

26

Three-dimensional close-packed
structures
Going from two to three dimensions
The easiest way to generate a three-dimensional close-packed structure is to stack close-packed layers on top of each other. The best
way to do this is not to merely place one atom above another, but
instead to translate the second layer such that the atoms in this
plane snug into the dimples between atoms within the first plane.
These dimples lie at the middle of the triangle defined by three
touching atoms. The interesting thing is that there are two times as
many of these dimples as there are atoms, separated by distance
a / 2 as seen in Figure 2.2. Thus when we come to add a third
plane, it has a choice of which set of dimples to use that has a sig27

INTERACTIVE 2.2 Hexagonal close-packed structure. We have


coloured the atoms in the two distinct close-packed layers
differently.

FIGURE 2.2 Two sets of dimple sites (red and yellow) in the 2d closepacked layer into which atoms in an equivalent layer above
achieve three-dimensional close packing.
nificance that was not important when we added the second layer. If
we denote the positions of the atoms as the set A, the positions of one
set of dimples as B and of the other set of dimples as C (see Figure 2.2),
we can consider the original plane of atoms as lying in set A and the
second set of atoms as lying in positions defined by B. We then need to
define the positions of layers 3 and beyond, and decide how often the
sequence of layers repeats in this new third direction. It is found that

nature tends to chose two options, namely the sequences ABAB and
ABCABC, which we will now consider separately.

ABAB stacking of planes: the hexagonal


close-packed structure
The ABAB stacking of planes is illustrated in Interactive 2.2. If we define the distance between atoms as (given as twice the radius of the
atoms), we have the lattice parameters within the plane defined as
a = b = , with simple geometry giving us the repeat distance between

28

the A layers as c = 8/3 =


= = 90 and = 120.

8/3a. The unit cell angles are

This new structure no longer retains the 6-fold rotational symmetry of


the single plane, but instead this is reduced to a 3-fold rotation axis
that is found passing through the dimples in the plane. However, the
complete symmetry of this structure can be said to reflect the 6-fold rotational symmetry when combined with a rotation. These symmetry elements are called screw axes. Because of the 6-fold rotation axis, this
lattice type is called hexagonal, and the specific structure formed by
combining closed-packed planes of spheres in the ABAB sequence is
called hexagonal close packed, abbreviated as hcp.
The unit cell of the hcp structure actually contains 2 atoms, a possibility that we encountered in the previous section. In terms of vectors, we
can define the positions of the two atoms as one at the origin of the
unit cell vector position (0,0,0) and the other at vector
a /3 + 2b/3 + c/2. This is not always a particularly helpful way to define
the positions of atoms. The more common method is to define the positions as xa + yb + zc, where x, y and z are conventionally have values
between 0 and 1. They are given the name fractional coordinates, and
clearly in the hcp structure the fractional coordinates of the two atomic
sites are
0,0,0

1 2 1
, ,
3 3 2

In the hcp structure, each atom has 6 atomic neighbours in the closepacked plane, and 3 atomic neighbours in each of the planes above
and below. This gives a total of 12 neighbours, arranged in the shape of
polyhedron shown in Interactive 2.3.

INTERACTIVE 2.3 The polyhedron formed by the nearest-neighbours


in the hexagonal close-packed structure, where each vertex
represents the site of the neighbouring atom and where the origin
atom is located at the centre of the polyhedron.
Examples of metals with the hcp structure are given in Table 2.1, together with the c/a ratio which, in the ideal case we previously discussed, should equal 1.633. Most materials have a c/a ratio that agrees
with the ideal value to within 4% or less, but two examples differ from
ideal value by more than 10%.

29

TABLE 2.1 Lattice parameters, c/a ratio, and percentage difference


from the ideal value of 1.633, for some hcp metals.
Element

a ()

c ()

c/a

Difference

Be

2.286

3.584

1.568

4.0%

Cd

2.979

5.619

1.886

+15.5%

Co

2.507

4.070

1.623

0.6%

Hf

3.196

5.051

1.580

3.2%

Lu

3.503

5.551

1.585

2.9%

Mg

3.209

5.211

1.624

0.6%

Os

2.734

4.317

1.579

3.3%

Re

2.761

4.456

1.614

1.2%

Ru

2.706

4.282

1.582

3.1%

Sc

3.309

5.273

1.594

2.4%

Tc

2.625

4.388

1.672

+2.4%

Ti

2.951

4.686

1.588

2.8%

3.647

5.731

1.571

3.8%

Zn

2.665

4.947

1.856

+13.7%

Zr

3.232

5.147

1.593

2.4%

ABCABC stacking of planes: the cubic


close-packed structure
If we follow the approach we took to generate the hcp structure, the
ABCABC stacking of planes willl lead to a unit cell containing 3 atoms
at positions with fractional coordinates
0,0,0

1 2 1
, ,
3 3 3

2 1 2
, ,
3 3 3

INTERACTIVE 2.4 Relationship between the hexagonal and cubic


settings of the cubic closed packed structure
and with c = 6a. However, this particular structure can also be described in terms of a cube with atoms in the middle of each face, as
seen in Interactive 2.4.
If we again specify the distance between atoms as , the lattice parameters of this cubic arrangement are a = b = c = 2, with the three unit
cell angles all equation to 90. The cubic unit cell has four atoms, at positions
0,0,0

1 1
, ,0
2 2

1 1
,0,
2 2

1 1
0, ,
2 2

But what we have defined is not a primitive cell, but instead a conventional cell where each atom position also represents a lattice point. The
30

INTERACTIVE 2.5 The relationship between the conventional unit cell


of the ccp structure (red thin lines), containing 4 lattice points, and
the primitive unit cell (thick lines).
primitive unit cell is compared with the cubic unit cell in Interactive
2.5. It can be shown that the primitive unit cell has a = b = c = and
all three unit cell angles equal to 60.
This type of conventional cell, with lattice points in the centre of each
face as well as at the origin, is called face-centred, and the lattice type
for the ccp structure is called face-centred cubic.
The stacking of close packed layers in the ABC sequence is further illustrated in Interactive 2.6, which shows the close-packed layers stacked
upon each other.

INTERACTIVE 2.6 Representation of the cubic close-packed structure


where different close-packed layers are shaded in slightly different
colours, with the ABC stacking sequence described in the text.
As in the hcp structure, each atom has 12 nearest neighbour atoms.
These are arranged to form a slightly different polyhedron, as shown in
Interactive 2.7.
The cubic close-packed structure actually represents a very efficient filling of space. Assuming the atoms can be described as touching spheres
of radius r, the lattice parameter is defined as a = 2 2r, giving a unit

cell volume per atom of Vcell = a 3 /4 = 4 2r 3. The volume actually occupied by one atom is Vatom = 4r 3 /3. Thus the fraction of space occupied by the spherical atoms is
f = Vatom /Vcell = (4r 3 /3)/(4 2r 3) = / 18 = 0.74.
31

TABLE 2.2 Lattice parameters for some ccp elements.


Element

a ()

Element

a ()

Element

a ()

Ag

4.085

Ir

3.839

Pt

3.924

Al

4.050

Kr

5.656

Rh

3.803

Ar

5.313

Ne

4.446

Sr

6.085

Au

4.078

Ni

3.524

Xe

6.129

Ca

5.588

Pb

4.951

Cu

3.615

Pd

3.891

Dynamics
The impression given by this discussion is that these structures consist
of hard spheres that are tightly packed together. In reality, as we will
see later, the atoms are not very hard, and there is a degree of flexibility with regard to how close the atoms can approach each other.
INTERACTIVE 2.7 The polyhedron formed by the nearest-neighbours
in the cubic close-packed structure, where each vertex represents
the site of the neighbouring atom and where the origin atom is
located at the centre of the polyhedron.
The hcp structure will have the same packing efficiency as the ccp
structure, because it uses the same closed-packed planes and stacking
of pairs of planes.
Some examples of ccp metals with their lattice parameters are given in
Table 2.2.

We will learn later that the atoms moved due to a large number of
waves travelling through the crystal. There are so many waves that the
accumulated action on any individual atom appears chaotic, but in reality the atomic motions are following well-defined vibrations.
Movie 2.1 shows the dynamics of atoms in a ccp structure at a temperature close to the melting point. The key impression to gain from this animation is the extent to which atoms are able to vibrate back and forth
in spite of the impression we have earlier that atoms can be represented as hard spheres. Atoms appear able to vibration with an amplitude as large as 1530% of the interatomic separation at temperatures
close to the melting point.

32

MOVIE 2.1 Dynamics of the atoms in the base plane of a ccp


structure close to the melting temperature. The animation shows a
large degree of thermal motion, yet clearly no atoms are leaving
the vicinity of their lattice sites.

33

Other simple monatomic crystal


structures
Simple cubic structure
We can imagine a crystal structure whose unit cell is a simple cube
and with only one atom in the unit cell. This lattice type is called
primitive cubic, and the unit cell length is equal to the spacing between atoms. This structure is illustrated in Interactive 2.8. Each atoms has six nearest neighbours.
In fact this structure for most materials will be unstable with respect
to an shear stress, unless there are strong directional forces. As a result, there is only one element, polonium, that is found with the simple cubic structure, with lattice parameter 3.359 .
34

INTERACTIVE 2.8 Crystal structure of polonium, with a simple


primitive cubic lattice.
The packing fraction of the simple cubic structure is obtained as before
as
f = (4r 3 /3)/8r 3 = /6 = 0.52
This is clearly not a very efficient packing.

Body-centred cubic
Many crystal structures of the metallic elements adopt a different simple crystal structure, which is defined as a cubic unit cell with an atom
at the centre of the cube, so that the fractional coordinates of the two
atoms are

INTERACTIVE 2.9 Body-centred cubic structure, where we colour the


atom at the centre of the unit cell slightly differently to facilitate
visualisation. Here we show 2 2 2 unit cells.
0,0,0

1 1 1
, ,
2 2 2

This structure is shown in Interactive 2.9.


In this structure each atom also represents a lattice point, as can be deduced by noting that the environment around the atom at the origin is
exactly the same as the environment around the atom at the centre of
the unit cell. The coordination number is 8. This lattice type is called,
for unsurprising reasons, body-centred cubic, abbreviated as bcc.

35

TABLE 2.3 Lattice parameters of some bcc metal crystal structures


(data from Webelements).
Element

a ()

Element

a ()

Element

a ()

Ba

5.028

Li

3.510

Rb

5.585

Cr

2.910

Mo

3.147

Ta

3.301

Cs

6.141

Na

4.291

3.030

Fe

2.867

Nb

3.300

3.165

5.328

Ra

5.148

The bcc monatomic structure has a packing fraction defined in the


same way as for the ccp structure, but this time with a = 4r / 3. Thus
Vcell = a 3 /2 = 32r 3 / 27, and hence
f = (4r 3 /3)/(32r 3 / 27) =

3 /8 = 0.68.

This is clearly slightly less efficient than the ccp structure.


The bcc structure is found as the crystal structure of a large number of
elements, particularly the alkaline metals (e.g. lithium and sodium), the
alkali earth metal barium, and several transition metals (e.g.iron and
chromium). Examples with lattice parameters are given in Table 2.3.

Widmansttten patterns revisited


In Chapter 1 we showed an image of the Widmansttten patterns seen
within iron-nickel meteorites. There we argued that the patterns suggested an internal microscopic structure; here we can now use our
new knowledge to explain this pattern.
The patterns arise because the iron-nickel solid solution at certain compositions is unstable, and separates into two phases, one of which is
relatively deficient in nickel (typically around only 510% of Ni) and

INTERACTIVE 2.10 The interface between the iron-rich kamacite


phase (green atoms) and the taenite phase (gold atoms) which has
a much higher nickel content.
the other of which has a much high nickel content (between 30 to
70%). These two phases, called kamacite and taenite respectively, have
different crystal structures. Kamacite has the bcc structure favoured by
pure iron, and taenite has the ccp structure favoured by pure nickel.
When these phases grow, they will develop interfaces between them,
and the overall energy is lowered if these faces of the two phases at the
interface match most closely. Thus turns out to be possible, as shown in
Interactive 2.10.
The bcc structure of taenite can be easily seen to be such in Interactive
2.10, because it is viewed down the z axis, which we will denote as
36

[001]. The ccp structure of the kamacite phase is harder to see because
we are viewing it down the x-y axis, the direction we will call [110].
The interface is then cut at an angle to the main axes in both cases. We
are not in a position to describe crystal faces or planes of atoms, but
for future reference, the interface is described as the (110) plane of
taenite and the (111) plane of kamacite. From knowing how we have
defined the vector directions, you can probably see that these plane
designations make some sense.
For a perfect match, the repeat distance going down the interface in
each direction needs to be very similar. In the case of this interface, the
two vectors are the [110] vector in kamacite and half of the [112] direction in taenite. For this interface to be relatively free of strain requires
the following relationship between the unit cell parameters:
3/2ataenite =

2akamacite

If we approximate taenite as the pure nickel ccp structure and kamacite as the pure iron bcc structure, using the data from the previous tables we find that the left hand term equals 4.316 and the right hand
side equals 4.055 . This matches to an agreement of 6%. This is not
perfect, but the difference can be accommodated.

INTERACTIVE 2.11 Interstitial sites in the ccp structure, where the


atoms are represented by blue spheres, the red spheres represent
sites of octahedral symmetry, and the teal-coloured sites represent
sites of tetrahedral symmetry.

Diamond structure

If we place additional atoms of the same type in the octahedral sites


we generate the simple cubic structure. For now this is not particularly
interesting (but these sites will be interesting when we consider binary
compounds).

If we look again at the ccp structure, it can be seen that there are potential sites for additional atoms in the spaces between the atoms of the
ccp structure. These are shown in Interactive 2.11, which shows sites
that have 6 atomic neighbours in the shape of a regular octahedron,
and sites with 4 atomic neighbours in the shape of a regular tetrahedron. We will consider these spaces in more detail when we consider
the crystal structures of some simple compounds.

The case of the tetrahedral sites is much more interesting here, because
the elements carbon (diamond), silicon, germanium and tin (the phase
of time called grey tin) crystallise in a structure in which half of the tetrahedral sites are occupied. The structure is shown in Interactive X.
The atomic positions are
37

It is worth re-iterating that each atom in the diamond structure has four
neighbouring atoms. This is unlikely to lead to an efficient packing of
spheres (see below), but instead elements forming the diamond structure have strong directional covalent bonding that favours tetrahedral
arrangements of atoms.
The packing fraction of the diamond structure is easily calculated. The
interatomic distance is one quarter of the cube diagonal, and so the lattice parameter is equal to a = 8r / 3. With 8 atoms in the cubic unit
cell, the space occupied by one atom is a 3 /8 = 64r 3 /3 3. The packing
fraction is thus
f = (4r 3 /3)/(64r 3 / 27) =

3 /16 = 0.34

which is exactly one half of the packing fraction of the bcc structure.

INTERACTIVE 2.12 Diamond structure, showing the outline of the


unit cell. The key feature is the tetrahedral coordination of the
atoms
0,0,0

1 1
, ,0
2 2

1 1 1
, ,
4 4 4

3 3 1
, ,
4 4 4

1 1
,0,
2 2
;

3 1 3
, ,
4 4 4

1 1
0, ,
2 2
;

1 3 3
, ,
4 4 4

Or described in conventional terms, the diamond structure has a facecentred cubic lattice, with two atoms in the entity called the motif or
basis that are then duplicated by the lattice vectors. The motif of basis
is described as having atoms at the two positions
0,0,0

1 1 1
, ,
4 4 4
38

What we have learned about crystal


structures so far
In defining the crystal structures of the elements thus far we have
had to introduce a number of formal concepts. We avoided the
temptation to develop the idea formally, but instead introduced
them as we needed them. In this section we will review some of
these, and provide hyperlinks to Appendices where these will be
dealt with in a more formal way.
We will soon need to care about the formal descriptions of these
concepts, because they give us a language to easily describe the key
features of crystal structures. Thus now is an appropriate point to
pause to look at these concepts.

39

Crystal symmetry
We have noted that crystals are defined by the existence of symmetry;
without symmetry a solid would be a glass rather than a crystal. We
identified two types of symmetry. One is the point symmetry, defined
as the set of symmetry operations that cause the environment around a
point to be transformed so as to generate an identical copy. The other is
the translational symmetry, by which translation of the entire crystal by
a lattice vector will give a new structure that is identical to the original
structure.

Translational symmetry: the crystal lattice


Translational symmetry is defined by the three lattice vectors, which
we have called a, b and c (but which some communities denote as a1,
a 2 and a3) . Translation of the crystal by any linear combination of the
lattice vectors will bring the structure back into coincidence with itself.
This is equivalent to saying that the crystal structure is periodically repeating along the three lattice vectors. In this context we defined three
terms:
the lattice is the complete set of points defined by all linear combinations of the three lattice vectors, and is thus a set of points that is periodic in space.
the unit cell is the parallelepiped volume encompassed by the three
lattice vectors. Whilst the crystal periodicity is what it is on a caseby-case basis, we do have a choice of how we describe it, and there
are cases where we can chose a definition of lattice vectors to give a
shape of the unit cell that reflects the symmetry of the lattice. Thus
where possible and when sensible lattice vectors are defined to be at
right angles to each other, even if that means defining a unit cell with

a volume that is some multiple of the primitive unit cell. The unit cell
edge lengths are denoted by the scalar quantities a, b and c, and the
angles between the three lattice vectors have symbols , and .
the motif or basis is the set of atoms associated with each lattice
point. Some of our examples so far only had one atom in the motif,
but the hcp and diamond structures both have two.

Centred lattices
We have met two descriptions of the unit cell (indeed, of the lattice) in
which it is conventional to define the unit cell to be a multiple of that
of the primitive lattice. This is always done in order to be able to define
unit cells with right-angles where possible. The possible centred lattices
are
Face-centred unit cell, where there is a lattice point at the origin of
the unit cell and at the centres of each face of the unit cell. A face
centred lattice is given the symbol F.
Body-centred unit cell, where there is a lattice point at the origin and
centre of the unit cell. A body-centred lattice is given the symbol I.
Single-face centred lattice, where there is a lattice point at the origin
and at the centre of one face of the unit cell. Depending on which
face is chosen, a single-face-centred lattice is given the symbol A, B
or C.
Rhombohedral lattice, that is too much to discuss here but should be
mentioned. This has symbol R.

40

Point symmetry
We have been introduced to point symmetry in terms of two symmetry
operations, namely rotational and mirror symmetries:
rotational symmetry exists when the crystal is rotated by 360 /n and
comes back into coincidence with itself. For a molecule the integer n
can have any value, but the existence of the crystal lattice is only
commensurate with values for n of 1, 2, 3, 4 and 6. Translational symmetry is inconsistent with n = 5, although this symmetry appears to
be present in some materials called quasi-crystals. We have explicitly
met the cases n = 3 and n = 6 so far. This symmetry is denoted by the
number n.
mirror symmetry describes the situation when a crystal can be reflected in a plane to generate an exact copy of itself.
There is technically one other type of symmetry we need to be concerned with, although it has one important case that is often treated
separately; this is the rotoinversion symmetry, in which a rotation
about an axis by an angle 360 /n is then combined with changing the
sign of all three coordinate values, called the inversion operation. Rotoinversion symmetry is denoted by the symbol n.
The special case of the simplest rotoinversion symmetry, 1, is called the
centre of symmetry. Many crystals have a centre of symmetry. Crystals
with this symmetry cannot have a polar axis without application of an
electric field.
It is also worth remarking that the rotoinversion symmetry 2 is equivalent to a mirror plane.

Combination of point symmetry operations:


the crystallographic point groups
It is commonly found that points within the crystal will have a combination of symmetry operations. For example, a point could be the intersection of three 2-fold rotation axes, or a 4-fold axis and one, two or
three mirror planes. Sometimes the combination of two symmetry operations will generate a third symmetry operation.
For example, two intersecting 2-fold axes automatically generate a
third. We take the origin at the point of intersection of the two 2-fold
axes, and let the 2-fold axes lie along the x and y axes. A general point
with coordinates x, y, z will have a symmetrically-equivalent position
generated by the first 2-fold axis at x, y, z, and one from the second 2fold axis at x, y, z (note that the operation of a 2-fold rotation is to
change the signs of two of the coordinates). Let us now take the first of
these new points and operate with the second 2-fold axis to create a
fourth symmetrically-equivalent position. This will change the sign of
the first and third coordinates: x, y, z x, y, z, which corresponds to the
operation of a 2-fold rotation axis along z.
A second example is two perpendicular (and hence intersecting) mirror
planes will be a 2-fold rotation axis. We consider the case where these
two mirror planes lie in the x-y and x-z planes, generating respectively
symmetrically related points at x, y, z and x, y, z (note that the operation
of a mirror plane is to change the sign of one of the coordinates). If the
second mirror plane operates on the new position generated by the first
it will generate a fourth symmetrically-equivalent position through
changing the sign of the second coordinate: x, y, z x, y, z. This corresponds to the operation of a 2-fold rotation axis along x.

41

A third example is the combination of a 2-fold axis and a perpendicular mirror plane. We take the case where the 2-fold axis (and hence the
normal to the mirror plane) is along the x axis. The operations of the 2fold axis and mirror plane respectively give new symmetricallyequivalent positions x, y, z and x, y, z. If we now operate on the position
generated by the 2-fold axis by the mirror symmetry, we generate a
fourth symmetrically-equivalent position: x, y, z x, y, z, which corresponds to the action of a centre of symmetry, of symbol 1.
To extend this analysis, it should be clear that three mutually perpendicular mirror planes will generate three 2-fold axes along the intersections of the mirror planes, and a centre of symmetry.
There are in fact 32 unique combinations of point symmetry elements
that are consistent with the existence of the crystal lattice; these are
called the crystallographic point groups.
At this point we have no reason to care about this level of detail, but
when we come to think about applications of crystal structures it will
be the case that symmetry can be extremely important. For example,
micro-scale transducers and actuators (for example in microphones),
infrared sensors, spark generators, and quartz oscillators, depend on
the absence of a centre of symmetry of the material being used. Some
optical applications depend on materials not having higher rotational
symmetry.

Combination of point symmetry operations


with translational symmetry
It is surprisingly common to find crystallographic symmetry that combines a rotation or mirror operation with a translation by some fraction
(typically, but not always, one half) of a lattice vector. These are called

screw symmetry and glide symmetry respectively. We do not need to


understand these symmetries to move forward, but we note that the
hcp structure is called hexagonal (reflecting the 6-fold symmetry that is
found in the close-packed layer) in spite of the fact that the dimple into
which the atoms of neighbouring atoms slot only have 3-fold rotational
symmetry. In fact the symmetry of the hcp structure has a 6-fold screw
axis.

Point symmetry and site symmetry


It is important to appreciate that the word point in the phrase point
symmetry means that there are a set of specific points within the unit
cell where the symmetry operates.
Let us take as an example a crystal that has three perpendicular mirror
planes, and let us suppose that these three mirror planes intersect at
the origin of the unit cell (a natural place to set the origin). We have
just noted that this crystal will also contain three 2-fold rotation axes.
Let us consider a point within the unit cell with coordinates x, y, z. The
set of symmetrically-related positions are
x, y, z

x, y, x

x, y, z

x, y, z

x, y, z

x, y, z

x, y, z

x, y, z

The origin corresponds to the point with the centre of symmetry. The
points that lie on the mirror planes have coordinates
0,y, z

x,0,z

x, y,0

If we one of these, say the first, operation of the 2-fold rotation axes
will generate equivalent positions at
0,y, z 0,y, z ; 0,y, z ; 0,y, z
42

Similarly the points that lie along the 2-fold rotation axes are
x,0,0 ;

0,y,0

0,0,z

Now if operate by the mirror planes, we get the set of equivalent positions
x,0,0 ;

0,y,0

0,0,z

Atoms and lattices: the motif


When we consider more complex crystal structures, we have to care
about where the atoms are located within the unit cell. It is common to
define the origin of the unit cell at a point that has significant point
group symmetry (often the choice is a centre of symmetry), and to give
the minimum number of atomic coordinates such that the complete
motif can be generated by application of the set of symmetry operations.
The full crystal structure can be conveniently described as the combination of the lattice and the motif, in that the crystal structure is obtained
by placing a copy of the motif at each lattice point. In mathematical
terms, this operation is called a convolution.
The atoms in the motif might be located at a general position in the
crystal, but may also lie on a point of symmetry. In the previous paragraphs we considered the symmetric sites generated by a combination
of mirror symmetry and 2-fold rotational symmetry. A motif containing
an atom at a general position will actually have eight symmetrically related versions.
In the crystal, if we have a point x, the translational symmetry will generate a point at 1 x, as illustrated in Figure 2.3. This is equivalent to

FIGURE 2.3 Three equivalent points (denoted by the crosses) in in


relationship to a unit cell (denoted by the thicker lines of the
rectangle). The far left point is generated from the point to its right
by a vertical mirror plane passing through the origin. The point on
the far right is the lattice-translation equivalent of the point on the
far left. It is clear from this construction that it is equivalently
generated by a mirror plane passing through the centre of the unit
cell.
the existence of a second mirror plane located at x = 1/2. Indeed, the
same argument holds true for centres of symmetry and rotation axes.
Thus if the crystal has a centre of symmetry at the origin, it will have
centres of symmetry at each of these points:
0,0,0

1
,0,0
2

1 1 1
, ,
2 2 2

1 1
0, ,
2 2

1
0, ,0
2
;

1 1
,0,
2 2

0,0,
;

1
2
1 1
, ,0
2 2
43

There will be a similar multiplication in the number of rotation axes.

The crystallographic lattices


This bit we have to know about. We have met two types of lattices
namely the cubic and tetragonal, but we have not said anything about
why they are defined as such, and nor have we enumerated the number of possible lattices.
The number of lattice types is defined by the point symmetry of the lattice points, and the point symmetry determines relationships between
the unit cell parameters. In the least symmetric case, namely where the
point symmetry is merely 1 (nothing at all except that it exists) or 1,
there are no constraints on the values of the lattice parameters.
The seven lattice types are
triclinic, for which the defining symmetry is either 1 or 1, with lattice
parameters a b c and .
monoclinic, for which the defining symmetry is the presence of a single 2-fold axis and/or mirror plane, with lattice parameters (in the
conventional setting) a b c and = = 90, 90.
orthorhombic, for which the defining symmetry is either a mirror
plane or 2-fold rotation axis associated with each of the three lattice
vectors, with lattice parameters a b c and = = = 90.
trigonal or rhombohedral, for which the defining symmetry is the presence of a single 3-fold axis, with two options for the lattice parameters: first is the case a = b c and = = 90, = 120, and second
is the case a = b = c and = = 90. Note that the first of these
is equivalent to the hexagonal lattice described below, and indeed

TABLE 2.4 Bravais lattices for each lattice type


Lattice

Triclinic

Monoclinic

Orthorhombic

Trigonal / Rhombohedral

Tetragonal

Hexagonal

Cubic

A, B or C

the lattice itself has a 6-fold rotational symmetry. But it is possible for
the motif to have only 3-fold rotational symmetry and take this lattice. A primitive unit cell with the second set of lattice parameters
can be recast in terms of the first if the 3-fold axis is aligned with the
c lattice of this first lattice, but in so doing the unit cell is three times
as large as that of the primitive unit cell.
tetragonal, for which the defining symmetry is the presence of a single 4-fold axis, with lattice parameters a = b c and = = = 90.
hexagonal, for which the defining symmetry is the presence of a 6fold axis, with lattice parameters a = b c ; = = 90, = 120.
cubic, for which the defining symmetry is the presence of four 3-fold
axes, with lattice parameters a = b = c and = = = 90. It is
tempting to believe that all cubic structures have 4-fold rotational
axes along the three cubic lattice directions, but there are cubic crystals for this which is not the case.

44

These 7 axes can be defined with and without centring, to give a total
of 14 lattices, called the Bravais lattices. These are enumerated in Table
2.4 and shown as 3d images in Appendix B.
It is worth remarking that when we combine the 32 crystallographic
point groups with the Bravais lattices, and allow screw axes to replace
rotation axes and glide planes to replace mirror planes, there is a total
number of 230 combinations, which are called the crystallographic
space groups.

Atomic coordination and atomic packing


In all the structures we have discussed we have noted a count of the
number of nearest-neighbour atoms. In the case of crystals composed
of pure elements, the best packing of spherical neighbours is achieved
with each atom having 12 neighbours (ccp and hcp structures). We
have, however, met structures in which the number of nearest neighbours is 8 (bcc), 6 (simple cubic) and 4 (diamond structures). In these
cases, the advantages of close packing are offset by the energy gain of
directional forces.
Coordination is an important concept in the crystal structures of compounds (next chapter). There we will find that the most efficient packing is determined by the relative sizes of the bonded atoms.
When we come to discuss the structures of molecular systems, including molecules of the elements, we will find that the structures are determined by the best way in which molecules can pack together. It is difficult to draw general principles, except to note that the best packing of
molecules with shape is often best achieved through the use of a lower
symmetry of the lattice.

Vectors between lattice points


We frequently need to formally describe the vectors between lattice
points in a structure. With our three lattice vectors a, b and c, we can
define the vector between any lattice points as a linear combination of
these vectors, as
t[U,V,W ] = Ua + V b + W b
where U, V and W are integers (when we consider primitive lattices;
where we have defined a conventional lattice we can include halfinteger values). The shorthand symbol [U, V, W ] is frequently used to
represent of t[U,V,W ]. In this context, the square brackets are the standard convention and are not arbitrary.
We can extent the convention further. In the case of symmetric lattices,
some vectors between lattice points will be symmetrically equivalent.
Thus in the case of cubic crystals, the three vectors [1,0,0], [0,1,0] and
[0,0,1] and their negative equivalents form a symmetric set. The set of
all symmetrically-equivalent vectors are designated by curly brackets;
in this case this set of six equivalent vectors are given the symbol
{1,0,0}. This shorthand might all sound somewhat is particularly useful
for calculations of lattice geometries.

Plans of crystal structures


Whilst three-dimensional images are nice, it is nevertheless the case
that they convey more of an impression that quantitative information.
The quantitative information we need is the unit cell and the coordinates of the atoms in the motif. This information is best conveyed
within a plan, which is a two-dimensional scale representation of the
unit cell viewed down one axis, showing the location of atoms and dis45

cubic, diamond and hexagonal close-packed.

Going forward
1/2

1/2

The key information that is needed going forward from this section is
that regarding the Bravais lattices. We will now meet crystal structures
of the elements that do not have cubic or hexagonal symmetry.

GALLERY 2.3 Some sample plans for the crystal structures we


have already met. Each plan shows the outline of the unit cell as
viewed down the [001] (z) axis, with each atom represented by
a circle. In all these plans there is only one type of atom. The
heights of the atoms expressed as fractional coordinates, are
represented by the numbers.
Cubic close-packed structure, showing the lattice points at the
centres of each face of the unit cell

playing their heights in fractional coordinates. Gallery 2.3 shows plans


for the structures we have met so far: cubic close-packed, body-centred
46

Crystal structures of other elements


What we have discussed so far covers many of the elements, mainly
many metals and the rare-gas elements, but we are a bit short of having accounted for all of them. Broadly we can say that the structures
of the remaining elements will be determined by very particular details of the chemical bonds they form; on the one hand, some of the
elements form simple molecules (such as chlorine and oxygen) with
weaker bonds between molecules and those within, and their crystal structures will be based on the packing of these discrete molecules. Other elements form structures where the packing of atoms
is, like in the diamond structure, determined by angular forces from
the bonding.
47

Pressure (GPa)

20
Melt
Diamond
10
Graphite
0

2000
4000
Temperature (C)

FIGURE 2.4 The phase diagram of carbon.


INTERACTIVE 2.13 Crystal structure of graphite, showing layers of
hexagonally ordered carbon atoms (now called graphene sheets).
In this section will be look at a number of specific examples, although
we will not attempt to be comprehensive, even though the number of
elements is limited, because in some cases the structures are very complex, and also because many elements can be found in more than one
structure depending on temperature and pressure. These different structures are called polymorphs. One examples is iron, which can exist in
either bcc or hcp form at ambient pressure depending on temperature,
and there are two well-known forms of tin, one of which has the diamond structure and the other has a body-centred tetragonal structure.

Carbon
For many years it was believed that there are two basic forms of carbon, namely graphite and diamond. The diamond structure is identical
to that of silicon discussed in an earlier section of this chapter, with tetrahedral coordination, whereas graphite (which is the more common
form as the stable form at ambient pressure) consists of sheets of carbon in a hexagonal arrangement, with weak bonds between the sheets.
This is shown in Interactive 2.13. The crystal structure has trigonal symmetry through the way that the hexagonal ring are stacked on top of
each other, with atoms at positions
0,0,0

0,0,

1
2

1 2
, ,0
3 3

2 1 1
, ,
3 3 2

48

MOVIE 2.2 Animation of the face-centred cubic phase of C60,


showing dynamic tumbling of the molecules.
Diamond is the high-pressure phase of carbon, and the range of temperatures and pressures under which these two phases are stable is
shown in Figure 2.3.
More recently it has been recognised that carbon can exist in the form
of large and extended molecules; examples are the spherical or ellipsoidal nanoscale molecules such as C60 and C70, and carbon nanotubes.
These molecules have strong covalent bonds within the molecule, with
weak interactions between the molecules.
The crystal structure of C60 is illustrative of a situation encountered in
many elements that form molecules. At higher temperatures C60 forms
a structure that is analogous to the ccp structure of atoms, where each

INTERACTIVE 2.14 C60 molecules crystallise in a simple cubic closepacked structure, which only develop orientational order at lower
temperatures. This image shows the ordered structure, that is
clearly based on the ccp arrangement of molecules.
molecule sits on a lattice site. The point symmetry of these sites contains 3 perpendicular 4-fold rotation axes, yet the molecule itself does
not have this symmetry. The only way for the symmetry to be reconciled is if the molecules are rotationally disordered, so that the 4-fold
rotational symmetry reflects the range of molecular orientations. In real49

INTERACTIVE 2.15 Single-wall carbon nanotube.


ity the disorder is dynamic, with the molecules in a constant state of
tumbling. This is illustrated in Movie 2.2.
At low temperature the molecules do order, which can only be accompanied by a lowering of the symmetry. In the case of C60 the ordered
structure, which is shown in Interactive 2.14, is still cubic, albeit with a
symmetry that does not include 4-fold axes. This change is called a
phase transition; the existence of phase transitions is often a reason
why materials have properties than lend themselves to very useful applications.
Carbon also exists in the form of nanotubes, both as individual tubes
and as tubes within tubes (multi-walled carbon nanotubes). A single
nanotube is shown in Interactive 2.15. Carbon nanotubes come in a

INTERACTIVE 2.16 Crystal structure of molecular chlorine, showing


the herringbone stacking of Cl2 molecules in an orthorhombic
crystal structure.
wide of diameters and length, and can be open (as in Interactive 2.15)
or capped at the ends. Where carbon nanotubes become particularly
interesting is in how they can be filled with other nanoparticles.

Diatomic molecules
Here we consider the molecules H2, N2, O2, F2, Cl2, I2 and Br2. The
halogens form crystal structures in which the molecules are aligned in
what is often called a herringbone structure, as illustrated in Interactive
2.16, typically with orthorhombic lattices. Like the case of carbon
50

INTERACTIVE 2.17 Orthorhombic crystal structure of molecular


sulphur, S8.

MOVIE 2.3 The disordered hexagonal close-packed structure of


crystalline N2. The molecules can be seen to be tumbling.
nanoscale structures, the covalent bonds with the molecules are strong
and the interactions between molecules are much weaker.
Whilst we might think of the molecules H2, N2 and O2 as dumb-bells,
with two atoms at the end of the rod representing the covalent bond, in
reality the density of the outer electrons is almost spherical, meaning
that these molecules almost interact as spheres. Thus like the case of
C60, the molecules have dynamic disorder in the higher-temperature

crystalline phase. For example, N2 has the hcp structure with tumbling
of the molecules. The structure is shown in Movie 2.3.

More complex molecular crystals


Here we simply cite the existence of some more complex element crystal structures. Sulphur forms large molecules, such as S8, and can exist
in at least two crystalline forms as this molecule. The orthorhombic
form is shown in Interactive 2.17. Sulphur is able to form other molecules, such as S6 and S18. It is able to form molecular chains, similar
to what we will see now with selenium.
51

INTERACTIVE 2.18 Crystal structure of selenium, showing spiral


molecular chains with weak interactions between chains.
Selenium and tellurium form spiral chains of covalently bonded atoms
with weak interactions between the chains. The example of selenium is
shown in Interaction 2.18; tellurium forms an identical structure.

Complex metallic phases of elements


We return now to the discussion of the crystal structures of elemental
metals. Most metals have the relatively simple phases (ccp, hcp, bcc)
that we have explored so far, but some metals have unexpectedly complex crystal structures.

INTERACTIVE 2.19 Crystal structure of a-manganese,


Interactive 2.19 shows the phase of manganese. This has a bodycentred cubic lattice (beware, this is not the bcc structure we have met
previously; standard nomenclature is not helpful here) containing 58
atoms in the conventional cubic unit cell.
Manganese forms other structures that are equally as complicated. And
it is not the only metal to form complex structures; others include the
-phase of gallium (Interactive 2.20) and boron.

52

INTERACTIVE 2.20 Crystal structure of -gallium (C-centred


orthorhombic structure).

53

CHAPTER 3

Crystal structures of simple compounds

Ionic crystals, coordination, and


structural polyhedra
General principles
The materials will look at in this chapter can, to a large extent, be
thought of as ionic materials. Even when there is significant covalent bonding, the atoms will be charged and therefore can still be
thought of as cations and anions. For example, SiO2, which has
strong covalent SiO bonds, nevertheless still appears to have
charged atoms (+2.4|e| for Si, 1.2|e| for O).
The primary bonding interactions in ionic materials are from the
Coulomb interaction (to be considered in a later chapter), and as a
result we expect to find cations having anions as nearest neighbours
55

INTERACTIVE 3.1 Tetrahedral [4] coordination

INTERACTIVE 3.2 Octahedral [6] coordination

and vice versa. And often we find symmetrical arrangements of the


ions, such as tetrahedral and octahedral arrangements, with structural
polyhedra connected via sharing of corners, edges or faces. To a large
extent the insights gained from seeing the crystal structures of compounds and being composed of structural polyhedra allow us to better
understand the behaviour of these materials.

we expect that the essential feature that determines crystal structure is


the relative size of the cations and ions. On the basis that cations have
anions as nearest neighbours, we can consider how many anions can
be packed around a single cation based on the relative radii of the two
ions. In short, the sum of the radii of the cation and anion define the
bond length, and the radius of the anion determines how many anions
can be packed around a single cation.

Size constraints

We consider the common cases of cations having 4, 6 and 8 anion


neighbours (these are common coordination numbers; other coordination numbers are more irregular). Given radii r + and r for the cation
and anion respectively, we can compute the ratios of radii for which

Before we look at some example crystal structures of the compounds,


it is worth understanding how the size of the ions will affect the possible crystal structures. If we need not worry about the shape of atoms,

56

INTERACTIVE 3.3 Cube [8] coordination

INTERACTIVE 3.4 Dodecagonal [12] coordination

the anions touch each other in each coordination, providing the case
of 12-coordination for good measure:

The values of these ratios define the limits of coordination based on


ionic size. Thus on the basis that the cation forms a bond with the anions, each coordination is only stable if the ratio is not greater than the
values cited here. We might also expect, in the absence of covalent or
directional bonding, that the most stable form of a crystal structure will
be that for which the greatest level of packing can be achieved, and
thus we expect that a material will form a structure with the highest
possible coordination number for the cation.

r
4-coordination: + =
r
r
6-coordination: + =
r
r
8-coordination: + =
r

1
3/2 1
1
21
1
31

r
12-coordination: + = 1
r

= 4.449

= 2.414
= 1.366

Examples
Values of ionic radii for many common cations and anions are tabulated in Appendix D. The ratios for some simple oxides are given in Ta57

TABLE 3.1 Ionic radius ratios for some common cations in


coordination with oxygen (ionic radius 1.25 ).

Can we really think of atoms as hard spheres?

Compound

Radius ()

Coordination

Radius ratio

MgO

0.86

1.453

CaO

1.14

1.096

TiO2

0.745

1.678

SiO2

0.40

3.125

Al2O3

0.675

1.852

ble 3.1, where we take the radius of the O2 anion as 1.25 . The fact
that Si4+ usually always forms tetrahedral coordination with O2 in compounds at ambient pressure is clearly explained by the radius ratio (at
high pressure it forms an octahedral coordination; the published tables
of ionic radii assign a larger radius for Si4+ of 0.54 this case). The examples of Mg2+, Al3+ and Ti4+ are clearly also comfortable as having octahedral coordination. The one oddity is Ca2+, which is a large ion and
would happily have 8 neighbouring oxygen atoms. In fact the coordination of Ca2+ is found to be quite variable between different materials
and crystal structures, and indeed irregular. In short, there is no requirement that the radius ratio rules be obeyed, but they do allow us some
degree of understanding of how materials behave.

Something about nomenclature


In this chapter we will need to give names to crystal structure types, because in almost every case a particular structure type will be found for
a range of compounds. For example, all the alkali halides are found as
one of two structure types.

The assumption that we can do so underpins the analysis of the


structures, but is it justified. The answer is yes and no.
Atoms are neither hard nor always spherical. In particular, interatomic separations can fluctuate by a few percent, suggesting a degree of softness or compressibility. And we know that
cations can have directional bonding, such that they prefer certain coordination states.
However, it is found that cationanion bond lengths will have
consistent values between different crystal structure and compounds, and as a result it has proven possible to draw up reasonable lists of ionic radii, albeit allowing radius values to depend on coordination number (which might not be thought satisfactory given that the radius ratio constraints on coordination
number can be satisfied by changing the ionic radii).
As such, the concept of the ionic radius does have a value in
rationalising the stabilities of crystal structures, particularly in
comparing different materials.

Sometimes these structures are named after a particular case; for example, the type of crystal structure found for NaCl is often called the NaCl
structure. This however can be somewhat complicated, not least because NaCl actually changes its structure to the other type of alkali halide structure under pressure. Instead, one common system is called the
Strukturbericht system, which is a way to designate structure types
based on letters and numbers. The letter codes are A for elements (see
58

TABLE 3.2 Strukturbericht labels for structures of the elements


encountered in previous sections
Label

A1

A2

A3

A4

A9

Structure

ccp

bcc

hcp

diamond

graphite

Table 3.2), B for binary compounds, C for AX2 compounds, D for AmXn
compounds, and so on, but quickly the scheme becomes unwieldy
since the number part labels the structure in a way that is not at all logical, and structures that are slight deformations of another structure are
not indicated as such. However, you will encounter this scheme in
your private research, and thus is worth indicating here, albeit without
being worth making a lot of fuss over.

59

Binary AX compounds
The way forward
It turns out to be useful to make reference to some of the simple elemental structures we looked at in Chapter 2, and to consider additional sites called interstitial sites or interstices that are simply
the voids int he elemental structures. We did just this when we considered the diamond structure, where we identified the octahedral
and tetrahedral interstitial sites in the ccp structure, and noted that
the diamond structure involved adding atoms to half of the tetrahedral interstices.

60

INTERACTIVE 3.5 Crystal structure of NaCl (B1), showing cations and


anions with different size and colour. This image shows a single
conventional face-centred cubic unit cell.
Clearly this approach automatically gives us the tetrahedral and octahedral coordination polyhedra that we identified in the previous section.
We will not restrict this to the ccp structure, as will quickly become evident.

MOVIE 3.1 Dynamics of atoms within a layer of NaCl at a low


temperature (100 K).

The NaCl structure type (B1)

tional unit cell), and the motif can be described as having fractional coordinates:

This structure can be described as the ccp structure with the octahedral
interstices occupied by the other atom; it is shown in Interactive 3.5.
The lattice type is face-centred cubic (4 lattice points in the conven-

A = 0,0,0 ; X =

1
,0,0
2

61

FeO). A large number of metallic compounds also crystallise with this


structure.
As we noted with the metals, the atoms are not held rigidly in place in
the crystal structure but are able to vibrate, with an amplitude that increases with temperature. Movies 3.1 and 3.2 show the thermal motion
in NaCl at two different temperatures, one low and one near the melting point.

CsCl structure type (B2)


This structure can be described as the simple cubic structure with the
central interstice occupied by the other atoms; it is shown in Interactive
3.6. The lattice type is primitive cubic, and the motif can be described
as having fractional coordinates:
A = 0,0,0 ; X =

MOVIE 3.2 Dynamics of atoms in NaCl at a temperature just below


the melting point.
with the other atoms in the cubic unit cell being generated by the operation of the face-centre symmetry. Essentially, this structure can be
described as ccp structures of the two elements, displaced by a shift
half way along the unit cell edge.

1 1 1
, ,
2 2 2

This structure looks like a body-centred cubic structure, because the


second site is at the centre of the unit cell as it is in the bcc structure,
but because the central site is occupied by a different atom it cannot
be said to be a lattice site. Thus structure is described as simple cubic
structures of the two elements displaced by each other by a shift half
way along the cube diagonal.
Examples of ionic materials that crystallise in this structure are the alkali halides CsCl, CsI and CsBr. A large number of metallic compounds
also crystallise with this structure.

Examples of materials the crystallise in this structure are all alkali halides except CsCl, CsI and CsBr, and binary oxides (such as MgO and
62

INTERACTIVE 3.6 Crystal structure of CsCl (B2), with a primitive cubic


structure. The two atoms in different colours, having 8 neighbours
each.

ZnS sphalerite structure type (B3)


In this structure type we follow the method used to generate the diamond structure, with the difference that we fill the half of the tetrahedral interstices with the second atom type. This is shown in Interactive
3.7. The lattice type is fcc, and each atom has tetrahedral coordination
with the other type of atom. The motif can be described as having fractional coordinates
A = 0,0,0 ; X =

1 1 1
, ,
4 4 4

INTERACTIVE 3.7 Crystal structure of ZnS sphalerite (B3), with a facecentred cubic structure. Each atom is in tetrahedral coordination.
This structure is found for a number of important semiconductor compounds, such as CdTe, InP and GaAs.

ZnO wurtzite structure type (B4)


Similar tetrahedral interstices can be identified in the hcp structure,
and by analogy these are now occupied in the wurtzite structure type.
This structure is shown in Interactive 3.8. The motif can be described as
having the fractional coordinates (within the hexagonal setting)
A=

1 2
1 2 3
, ,0 ; X = , ,
3 3
3 3 8
63

INTERACTIVE 3.8 Crystal structure of ZnO wurtzite (B4), with a


hexagonal crystal structure. Each atom is in tetrahedral
coordination.
This structure is found for both ZnO and ZnS, and also for the important semiconductor GaN (the important component of blue lightemitting diodes).

Other AX structures
It is not necessary to develop a longer list of structures, although in fact
there are not many standard types. The NiAs (B8) structure is one that is
based on the hcp structure with the filling of octahedral interstices. A
number of other structures are generated by intermetallic compounds.

64

Simple compounds of formula AX2


Preamble
Although the title formulae are very simple, we now have an accelerating number of possible structures. For example, TiO2 exists in
three known structures at ambient pressure. Moreover, the structure
of SiO2 is based on corner-sharing SiO4 tetrahedra, and it turns out
that there is probably an infinite number of structures that are possible. Thus here we will focus only on a few simple and important
structure types, to give as representative examples.
The principles of forming structures from structural polyhedra based
on relative ionic radii applies equally for more complex materials.
65

INTERACTIVE 3.9 Crystal structure of CaF2 (fluorite, C1), with the


fluorine atoms (light green) in tetrahedral coordinate and the
calcium atoms (dark green) having 8 neighbours.

CaF2 fluorite structure type (C1)


The fluorite structure recalls the way we generated the diamond structure, but now the F atoms occupy all the tetrahedral interstices in the
ccp structure rather than just half. The structure is shown in Interactive
3.9. In this structure, because of the filling of all the tetrahedral interstices, the Ca cations have 8 nearest neighbours.

INTERACTIVE 3.10 Crystal structure of the -cristobalite form of SiO2


(C9). Silicon atoms are presented by blue spheres, and oxygen
atoms by red spheres. We have shaded the SiO4 tetrahedra to
guide the eye. As you rotate the structure, observed the linear Si
OSi bonds. Also look for the 6-membered rings of SiO4
tetrahedra.

SiO2 cristobalite structure (C9)


The cristobalite structure is found as the high-temperature form of silica, SiO2; the stable phase at room-temperature is the better known
quartz form. We show the structure in Interactive 3.10. It can be described well as the diamond structure for the silicon atoms, with oxygen atoms placed half-way between the nearest-neighbour silicon contacts. Since the silicon atoms are in tetrahedral coordination with their
66

The curious point about this structure is that it has linear SiOSi linkages, which are not energetically favourable, and instead the structure
deforms by tilting of the rigid SiO4 tetrahedra to generate a tetragonal
structure with bent SiOSi linkages. At low temperature this occurs by
a phase transition to a tetragonal structure, but at high temperature this
is a dynamic process, involving large-amplitude rotations of the SiO4
tetrahedra as shown in Movie 3.3.
The cristobalite structure is the first structure that is crystallised by annealing silica glass, and indeed the density of silica glass is much
closer to that of cristobalite than of quartz.

Other silica structures


The most common form of silica is quartz. Like cristobalite, it consists
of an infinite network of corner-sharing SiO4 tetrahedra. At high temperature quartz has a hexagonal crystal structure, with spirals of tetrahedra along the z axis. In fact the structure deforms slightly through small
rotations of the tetrahedra on cooling below a temperature of 858 K,
and the lattice symmetry is lowered to trigonal (this is given the symbol
C8). The hexagonal structure of quartz is shown in Interactive 3.11.
MOVIE 3.3 Motions of the SiO4 tetrahedra in -cristobalite, showing
the large rotations as they move to bend the SiOSi bond to a
more favourable configuration. We have coloured a central 6membered ring as a guide to the eye. The view is along a direction
normal to the layers of six-membered rings.
neighbour silicon atoms, it follows that this structure now has SiO4 tetrahedral units. The structure retains the fcc lattice type.

There is probably an infinite number of infinite networks that can be


formed from corner-sharing tetrahedra. Silica structures can be created
with a range of densities, as networks can be formed that contain large
pores and channels. These structures are known as zeolites. In nature
they are frequently found with Al3+ cations substituting for Si4+, and
with charge-compensating cations located within the pores. These natural zeolites tend to contain a lot of water, and the very name zeolite
represents the fact that this water is boiled out of the rocks when
heated.
67

INTERACTIVE 3.11 Crystal structure of quartz, SiO2, where the SiO4


tetrahedra are shaded. Note the hexagonal spirals of cornersharing tetrahedra.
Interactive 3.12 shows the zeolite called chabazite. This structure has
three-dimensional channels between the pores, which are built from
rings containing 8, 6 and 4 tetrahedra. The density of the silica form of
chabazite is about 60% of that of quartz, even though they have the
same chemical composition, and the unit cell of chabazite has a volume that is nearly 7 times larger than the unit cell volume of quartz.
And chabazite is one of the simpler zeolites!

INTERACTIVE 3.12 Crystal structure of the zeolite chabazite, showing


the network of SiO4 tetrahedra, with a charge-compensating cation
shown in the cavities. This structure is rhombohedral but with a unit
cell angle close to 90 so that it almost looks cubic.

Cu2O structure (C3)


This structure is illustrated in Interactive 3.13. It can be described in either of two ways. The simplest is to think of the monatomic bcc structure, with sites occupied by the oxygen atoms, and to then place the
copper atoms at the mid-point along half the vectors from the central
oxygen site to the eight corners in a way that generates a perfect tetrahedral arrangement of copper atoms around each oxygen atom. Taken
alone, the Cu atoms form a face-centred cubic arrangement, with the
O atoms occupying half the tetrahedral interstices that are occupied in
68

INTERACTIVE 3.13 Crystal structure of Cu2O (C3). This has a


primitive structure with a network of corner sharing OCu4
tetrahedra (oxygen atoms are represented by red spheres, copper
atoms by blue spheres). The structure can be described as two
interpenetrating but unlinked cristobalite structures.
the sphalerite structure (and one quarter of those occupied in the fluorite structure). But taking the structure as a whole, it has a primitive cubic crystal structure.
The second way to view this structure is as two interpenetrating cristobalite structures, made from OCu4 tetrahedra rather than SiO4, displaced relative to each other by one unit cell in any of the x, y and z
directions. Unlike in cristobalite, though, the shortest distances be-

INTERACTIVE 3.14 Crystal structure of rutile, TiO2 (C8). The structure


consists of TiO6 octahedra (green, with the oxygen atoms as pink
spheres at the corners), which are linked together at shared
corners and on shared edges.
tween the corners of the tetrahedra within and between tetrahedra are
the same.

Octahedral AX2 structures


As we have said, there are simply too many other structures that could
be explored, except that it will become overwhelming. Many of these
structures consist of octahedral coordination polyhedra, typically con-

69

INTERACTIVE 3.15 Crystal structure of pyrite, FeS2, consisting of FeS6


octahedra (highlighted in this image) connected at corners.
nected through shared edges or faces, and here we will show two examples.
TiO2 exists in at least three different crystal forms, but all share some
common characteristics. The rutile (C4) form is shown in in Interactive
3.14. It consists of TiO6 octahedra that are connected to form an infinite network, through shared corners and edges. This forms a tetragonal
structure.
The mineral pyrite, FeS2, better known as fools gold on account of its
golden colour and metallic feel, has a primitive cubic structure (symbol
C2; interestingly, this is an example of a cubic structure that does not

INTERACTIVE 3.16 Crystal structure of CdI2 (C6), showing layers


of edge-sharing CdI6 octahedra arranged in layers, with weak
interactions between neighbouring layers.
have 4-fold rotation axes). The structure is shown in Interactive 3.16.; it
consists of FeS6 octahedra that are connected in groups of three at corners.
An interesting contrast is the case of CdI2, shown in Interactive 3.16
(trigonal structure, symbol C6). This consists of CdI6 octahedra that
share a lot of edges, with the result that the structure consists of layers
held together by weak forces. There are a number of ways that neighbouring layers can be oriented with respect to each other, and as a result there are a number of stacking sequences found between and
within different examples of this type of structure.
70

More complex crystal structures


The perovskite structure, ABX3 (E2)
The one example that we must consider in some detail is the
perovskite structure, named after the mineral of chemical formula
CaTiO3. The perovskite structure is one of the most important crystal
structures for modern life.
The parent structure is of primitive cubic lattice, and is shown in Interactive 3.12. The fractional coordinates of its motif can be described in either of two ways, depending on where we chose to
place the origin:

71

INTERACTIVE 3.17 Crystal structure of the ideal cubic structure of


perovskite (E2). We have drawn this with origin choice 2, with the
octahedrally coordinated atom (B, green sphere) at the centre of
the unit cell, and the atom with 12 neighbours (A, cream sphere) at
the origin of the crystal. The X atoms, which are usually oxygen,
are now drawn as ellipsoids that represent the amplitude of
thermal motion; this shows that the oxygen atoms tend to move
more in the direction perpendicular to the BXB vector.
Origin 1: A at 0,0,0 ; B at
Origin 2: A at

1 1 1
1 1
1 1
1 1
, , ; X at , ,0 , ,0, , 0, ,
2 2 2
2 2
2 2
2 2

1 1 1
1
1
1
, , ; B at 0,0,0 ; X at ,0,0 , 0, ,0 , 0,0,
2 2 2
2
2
2

Regardless of how we chose to write the structure, the key characteristic is that the B site has six X neighbours forming a perfect octahedron

MOVIE 3.4 Thermal motion in the perovskite structure, where the


BX6 octahedra are shown as shaded polyhedra.
around B, with the BX6 octahedra connected via corner sharing atoms
to form an infinite three-dimensional network. The A cation occupies a
site with twelve X neighbours.
An interesting point about this structure is that it is effectively overconstrained in terms of the ionic radii. If the atoms had radii such that
72

both the A and B atoms were touching the X atoms, we would have the
following two formula based on ionic radii r :
a = 2(rB + rX)
a=

2(rA + rX)

It is never possible to satisfy both equations, and as a result the structure is liable to distort in some way. Sometimes this is achieved through
rotations of the BX6 octahedra, as indeed occurs in the type structure
CaTiO3. In other cases the distortions to accommodate size mismatch
involve the cations being slightly displaced from their symmetric sites
(which in the ideal cubic structure are sites with a centre of symmetry).
The fact that the perovskite structure is able to do so means that it can
accommodate a huge range of chemical compositions, which in turn
means that the physical properties can be finely tuned.
The X (oxygen) atoms are represented by ellipsoids that reflect the anisotropy of their thermal motion. It can be seen that the main motions
are in the direction perpendicular to the BXB direction. This is highly
suggestive of significant rotational motion of the BX6 octahedra. This is
confirmed through animations of the thermal motion of the structure,
seen in octahedral and atomic representations in Movies 3.2 and 3.3
respectively.
Perovskites can be formed both as oxides and halides (some reports of
perovskites made replacing oxygen by sulphur). In the case of oxides,
the total cation charge will need to add up to +6 | e | ; examples are
where one cation has charge +5 | e | and the other +1 | e | , such as
NaNbO3; where one cation has charge +4 | e | and the other +2 | e | ,
such as BaTiO3; and where both cations have charge +3 | e | , such as
LaAlO3. It is possible to have the A site vacant and the B site occupied
by a hexavalent cation, as in WO3. In the case of the halides the sum of

MOVIE 3.5 Thermal motion in the perovskite structure, showing the


motions of individual atoms. Blue spheres represent the B atoms at
the centre of the BX6 octahedra, pink atoms represent the X atoms,
and the green atoms represent the B cations.
the cation charge is +3 | e | , so we have examples such as ReO3 or ScF3
without the A cation. The crystal structure of the latter compound is
shown in Interactive 3.18, where the transverse thermal motion of the
linkage F atom is very clear.
73

INTERACTIVE 3.18 Crystal structure of ScF3 (cubic symmetry, symbol


D09), where the spheres represent the Sc3+ cations and the
ellipsoids represent the F anions. The shape of the ellipsoid
represents the extent of thermal motion of of the F anions, showing
considerable motion in directions perpendicular to the ScFSc
linkage.
The perovskite structure often allows the formation of solid solutions.
This for example it is possible to prepare samples like PbTixZr1xO3,
with all values of the composition variable x. In this case the two end
phases have different distortions, and in middle compositions the
structure can easily be flipped from one distortion to another. This
enhances some of the physical properties, such as the dielectric

INTERACTIVE 3.19 Crystal structure of the orthorhombic perovskite


CaTiO3, showing the TiO6 octahedra as solid objects and the Ca2+
cations as green spheres. Note that the degree of rotation of the
TiO6 octahedra is different about each of the three axes.
constant and its piezoelectric properties, which is extremely useful for
many applications.
The perovskite structure can typically deform in one of two ways.
Firstly, if the cation occupying the A site is smaller enough, the
structure can deform by rotations of the octahedra. These can occur
about all three crystal axes, and different examples show different
combinations of rotations. The example of CaTiO3 is shown in
Interactive 3.19. This has the same pattern of rotations as the silicate

74

Polarisation (Cm2)

0.8

0.6

0.4

0.2

0.0
300

400

500
600
Temperature (K)

700

800

FIGURE 3.1 Temperature-dependence of the dielectric polarisation in


the tetragonal perovskite PbTiO3.
INTERACTIVE 3.20 Crystal structure of PbTiO3 in its distorted
tetragonal phase. The symmetry-breaking displacements of the ions
along the c axis is clear.
perovskite MgSiO3, which is believed to be a major phase in the Earths
inner mantle region.
A second deformation involves displacements of the cations and
anions in different directions, breaking the centre of symmetry of the
crystal structure and leading to the formation of a static dielectric
polarisation. The example of PbTiO3 is shown in Interactive 3.20. The
distorted phase has tetragonal symmetry, with displacements of the
cations and anions along the crystallographic [001] axis. The distortion
of the TiO6 octahedron is clear. The direction of the atomic

displacements and hence of the dielectric polarisation can be switched


relatively easily by application of an external electric field.
The structural distortion only exists below a certain temperature, Tc,
which defines a phase transition between the high-temperature
paraelectric phase, and the low-temperature ferroelectric phase. Tc is
called the critical temperature, or sometimes the Curie temperature
through the analogy with the ferromagnetic phase transition. Indeed,
the analogy between the ferroelectric state and the ferromagnetic state
are is really quite close, which this phenomenon is called
ferroelectricity. A large number of ferroelectric materials are know,
many of which, but not exclusively so, have the perovskite crystal
structure.
75

The atomic displacements in PbTiO3 produce an expansion of the c


axis relative to the a and b axes. Application of stress along the [001]
axis will change the size of the atomic displacements and hence of the
dielectric polarisation. This is called the piezoelectric effect, and the
size of the effect in perovskites is both large and tuneable with
chemical composition (eg in the solid solution with PbZrO3).
A third effect of this type of phase transition in the perovskite structure
is that close to the transition temperature Tc the dielectric susceptibility
diverges, typically varying as
Tc T

This leads to enormous values of dielectric constants, in the thousands


or more when conventional materials have dielectric constants in the
tens at most. The temperature dependence of the dielectric constant of
the perovskite LiTaO3 is shown in Figure 3.2, where the divergence of
the dielectric constant at Tc can be seen.

103

20

104 /

The atomic displacements in the tetragonal phase are dependent on


temperature, becoming larger on cooling, and the extent of the
displacements directly determines the size of the dielectric
polarisation. It thus follows that the dielectric polarisation varies with
temperature, actually to quite a large degree compared with many
other materials, which gives a direct coupling between the electrostatic
field and the temperature. The temperature dependence of the
polarisation in PbTiO3 is shown in Figure 3.1. Such an effect gives rise
to the capability to measure small changes in temperature, enabling
perovskites to be used as the active component in heat and infrared
sensors.

10

0
700

800

900

1000

Temperature (K)
FIGURE 3.2 Temperature dependence of the dielectric constant of
the perovskite LiTaO3 together with its inverse to show the linear
relationship with temperature.
The perovskite structure is one that you will want to be familiar with.
Materials with this structure have properties that lead to a huge impact
on modern life, along with silicon. The rapid progress in technology,
particularly digital technologies, would not be possible without the
perovskite structure.
76

Molecular crystals
Introduction
In the chapter on the crystal structure of the elements, we noted that
some elements (for example, carbon, nitrogen and the halogens) exist as molecules that retain their molecular character completely
when they form crystal structures. In this section we will look at
crystals that are formed from discrete molecules.
We will find an interesting contrast between molecular symmetry
and crystal symmetry. In some cases, the most efficient packing of
molecules disregards the symmetry of the molecule. In other cases,

77

INTERACTIVE 3.21 Crystal structure (monoclinic) of naphthalene,


C10H8, showing a characteristic herringbone arrangement.

INTERACTIVE 3.22 Crystal structure (monoclinic) of


diphenylacetylene, C14H10.

the molecules are almost spherical and have dynamic rotational disorder, as we met previously for N2 and C60.

of the molecule that is utilised in the crystal structure is the centre of


symmetry.

Herringbone structures

The molecular stacking in naphthalene is a characteristic herringbone


structure that is repeated in many other molecular crystals. A more
complex example is shown in Interactive 3.22.

Many molecular crystals form crystal structures in a low-symmetry lattice often monoclinic in an arrangement that maximises packing efficiency without exploiting the symmetry of the molecule. One typical
example is the crystal structure of the molecule naphthalene, C10H8.,
shown in Interactive 3.21. The molecular symmetry has a centre of symmetry, three mirror planes and three 2-fold axes. Yet the only symmetry

More symmetric molecular crystals


A few molecular crystals do make use of molecular symmetry. One example is sym-triazine, C3N3H3. Its molecular symmetry of a 3-fold rota-

78

INTERACTIVE 3.23 Crystal structure (rhombohedral but in the setting


of the non-primitive hexagonal lattice) of sym-triazine, C3N3H3,
showing molecules arranged in planes with the molecular 3-fold
rotation axes lying along the crystallographic 3-fold axes.
Neighbouring molecules along the c axis are rotated by 180 with
respect to each other, such that the view down the c axis gives the
impression that each molecule might have 6 hydrogen atoms.
tion axis and a perpendicular mirror plane are reflected in the rhombohedral crystal structure, as shown in Interactive 3.23.
In fact the symmetric structure of sym-triazine is only stable at higher
temperatures (from 198 K to its sublimation point), and at lower temperatures the structure distorts to a monoclinic structure through a
shear of the unit cell and rotations of the molecules about a plane pass-

MOVIE 3.6 Animation of the molecular vibrations in the


rhombohedral crystal structure of sym-triazine, viewed down the
trigonal c axis.
ing through the molecular mirror planes. This new structure does not
exploit the symmetry of the molecule at all.
Movies 3.6 and 3.7 show the vibrations of the molecules at high temperatures, viewed along two different directions. These animations
show the extent of molecular rotational motions possible within an ordered molecular crystal structure.
79

MOVIE 3.7 Animation of the vibrations of sym-triazine in the hightemperature rhombohedral phase, viewed edge-on to the
molecular mirror plane.

Orientationally-disordered structures
The examples of molecular crystals we have shown so far involve molecules with distinct shapes. On the other hand, some molecules have
such highly-symmetric structures that they are almost spherical, and in
these cases the orientational forces within the crystal structure are not

MOVIE 3.8 Dynamic disorder in the crystal structure of SF6. One


molecule has been coloured as a guide to the eye.
strong enough to keep the molecules in a fixed orientation. Thus these
molecules can form crystals in which the molecules are allowed to tumble. We met two examples previously, namely C60 and N2. Here we
consider two other examples. The first is SF6, which turns out to be
rather surprising in the sense that the perfect octahedral symmetry of
the molecule is exactly reflected in the symmetry of the site in the crystal structure (body-centred cubic structure with one molecule per lat80

unit cell vectors. However, this alignment (presumably favoured in


terms of the interaction between the molecules at the origin and centres of the unit cell) places the neighbouring F atoms along the unit cell
edge at too short a distance. Hence the molecules are in a constant
state of tumbling as they seek to form favourable correlated orientations. This tumbling is seen in Movie 3.8. On cooling there is a phase
transition to an ordered C-centred monoclinic structure.
The molecule of adamantane, C10H16, forms a globular cage shape. On
first forming it has a face-centred cubic structure with one molecule
per lattice point. Because the crystallographic site has higher symmetry
than that of the molecule (for example, the molecule does not have 4fold rotational axes that the crystallographic site has), the molecules
are orientationally disordered, as shown in Movie 3.9. This is very similar to the case of C60. On cooling there is a phase transition to an ordered body-centred tetragonal structure.

Hybrid metal-organic structures

MOVIE 3.9 Orientationally disordered phase of adamantane,


C10H16, showing dynamical reorientational motions of the
molecules. One has been coloured as a guide to the eye. The
crystal structure has face-centred cubic symmetry, where the nearspherical molecules pack in a manner analogous to the packing of
spherical atoms in the cubic close-packed structure. This view is
down the cubic a axis.
tice point). On average the orientation of the molecules are aligned
with the crystal lattice, with the SF bonds pointing along the cubic

In the field of crystal design, there is a lot of interest in hybrid structures made from metal cations and organic ligands. It is possible to design structures that are analogues of zeolite frameworks, for example
where a metal cation replaces the silicon site and the ligand replaces
the linkage oxygen site, but because of the size of the organic molecules it is possible to create structures with very large pore size and
concomitant low density.
One example is zinc imidazolate, Zn(C3N2H3)2, where the zinc atoms
bond to the nitrogen atoms in the ligand. The analogy with silica is
made complete by the fact that the zinc atoms can bond to four
ligands, forming near-perfect ZnN4 tetrahedra which appear to be reasonably rigid entities. An example is shown in Movie 3.10, which
81

MOVIE 3.10 Crystal structure of the hybrid metal-organic material


zinc imidazolate, Zn(C3N2H3)2, showing ZnN4 tetrahedra and the
large pores enabled in such a structure.
shows firstly the large pore sizes achievable, and secondly the extent of
thermal motion allowed in such structures.

82

CHAPTER 4

Bonding

Introduction
In this chapter we pose the question of what holds the atoms together in a crystal structure. One one hand the answer might be considered to be quite simple: it is the electrostatic energy coupled
with the existence of electrons. But on the other hand, the way in
which this combination works out in practice gives us a wide variation of bonding. Here we mostly consider bonding in non-metals,
and we

84

Coulomb interaction
The primary binding force in ionic crystals comes from the Coulomb interaction between charges (which is important even in crystals where the nearest bonds could be said to be covalent, such as
in SiO2).
The Coulomb energy between two ions of charge Q1 and Q2 has the
form
E=

Q1Q2
40r

where r is the distance between charges (assumed to be points), and


0 = 8.8542 1012 F/m is the permittivity of free space. The key fac85

15
10

Energy

5
0

5
Distance

10

FIGURE 4.1 Contributions to the Coulomb energy of the NaCl structure, where the distance scale is set as the nearest-neighbour has
distance value of 1, and where the energy is in units where the charge values and 1/40 are set to be equal to 1. The connected blue
points represent the contribution to the crystal energy neighbour-by-neighbour. The red curve represents the running sum of the
neighbour-by-neighbour points. The black line represents the summation over 8-ion cubes. The result at infinite distance is the Madelung
constant, which for the NaCl structure is equal to 1.74756.
tor here is that the energy varies as the inverse of the distance. If we assume a uniform distribution of charged sites in the crystal that is characterised by the quantity that defines the number of atoms per unit volume, a thin spherical shell of space at a distance r from an origin
charge with thickness r will have a volume 4r 2r and hence will
contain 4r 2r charged sites. Thus the energy of interaction between
the origin charge and the charges within the shell will be equal to

E=

Q1
QQ
Q2 4r 2r = 1 2 rr
40r
0

The problem is that this energy is now proportional to r, so when we


sum to very large values of r (indeed, towards infinite values) the Coulomb energy might diverge. This is countered by the fact that there is
an equal distribution of positive and negative charges, which will cancel out the infinity values, but cancelling infinities is a dangerous
mathematical task.
86

We illustrate this by calculating the energy of a NaCl structure


neighbour-by-neighbour, starting from one ion as the origin. This is illustrated in Figure 4.1. For the NaCl structure the charge of the nearest
neighbour is of opposite sign to that of the origin ion, second neighbour has the same sign, third of opposite sign, and so on. Figure 4.2
shows the neighbour-by-neighbour energy together with the running
total at each neighbour. It is seen that even going to large distances the
running total is oscillating a lot; for reference the value if you could
sum out to infinity is 1.73, and clearly we are not converging towards
this value even at the large distances in the figure. This is the infinite
sum problem.
So we could be a bit more clever and do the sum using groups of ions
instead of individual ions. The first thought might be to use a pairs consisting of a cation and an anion, which would be neutral. This however
has the problem that the pair has a dipole moment, and the interaction
between two dipoles varies as 1/r 3. Whilst this might seem better than
1/r as for the pure charges, it turns out to be not much better. When we
multiply by 4r 2r we have a shell-by-shell energy that varies as 1/r,
and the integral to evaluate the sum to infinity diverges logarithmically.
So the next thing might be a square of two cations and two anions, but
even better would be a cube of 4 cations and 4 anions, namely one cubic unit cell. The energy in this case will vary as 1/r 7, which should
converge well. And indeed it does, as we see in Figure 4.1 which also
shows the running total obtained by adding different neighbours of
cubes. Clearly now the running total is converging well.
This sum can be done even more rigourously. For cubic binary ionic
crystals (that is, of representative chemical formula AX, AX2, AX3)
where the positions of all cations are determined by symmetry (thereby

TABLE 4.1 Values of the Madelung constant for various highsymmetry cubic crystal structures, together with the relationship
between nearest-neighbour distance and the unit cell parameter.
Lattice type

d/a

NaCl (B1)

1.747567

1/2

CsCl (B2)

1.762677

3/2

Sphalerite (B3)

1.638057

3/4

Fluorite (C1)

2.519396

3/4

Cuprite (C3)

4.442482

3/4

Cristobalite (C9)

4.453383

3/8

ReO3 (D09)

2.984664

1/2

excluding examples such as FeS2), the Coulomb energy can be described by a simple formula:
Q +Q
E = NA
40 d
where Q + and Q are the charges of the cation and anion, d is the distance between nearest-neighbour ions, NA is Avogadros number, and
is a numerical constant called the Madelung constant that is a fixed
number for each structure type. Examples are given in Table 5.1.
For all other crystal structures (ionic crystals with more than 2 ion
types, cubic crystals where the ion positions are not determined by
symmetry, and crystals belonging to another lattice type; basically almost all crystal structures!) there are sophisticated methods to compute
the Coulomb energy accurately and efficiently. These are all based on
Fourier transform methods.
87

Note that the Coulomb energy of a crystal is always negative and will
become more negative as the volume is decreased.
We do have the question of what values of ionic charge to use. In some
cases it might be quite reasonable to use the formal valence charges
(for example, Q + = + | qe | and Q = | qe | , where qe is the charge of
the electron, works well for NaCl), but there will be cases where there
is some degree of covalent bonding and when charge values lower
than the formal valence charge are more realistic. For example, many
models for SiO2 have Q + = + 2.4 | qe | and Q = 1.2 | qe | instead of the
formal values +4 | qe | and 2 | qe | respectively, as determined by quantum mechanical calculations.

88

Short-range repulsive interactions


The Coulomb energy of a crystal favours short interatomic interatomic distances, but when atoms are close the electrons of the two
atoms begin to overlap and thereby raise the energy. This increase in
energy arises from the Coulomb interaction between electrons, together with the effects of quantum mechanics which create new but
higher energy states. These effects only work over short distances,
but the energy rises very quickly over small changes in distance. For
this reason atoms and ions appear to have a particular size which
justifies the idea of using ionic radii in calculations, but the repulsive energy is a smooth function of interatomic distance and hence
there will be some variability of observed interatomic distances.
89

We do not have a derived formula for the repulsive energy. However,


quantum mechanics calculations of the interactions between closedshell ions appear to be consistent with an energy expression of the
form

5
4

where r is the interatomic separation, E0 is an uninteresting constant


which is usually set equal to zero in applications of the model, and B
and are constants for a particular ion pair. This function is known as
the BornMayer potential energy function.
Since the repulsive interactions operates only over a small range of interatomic separations, it might be expected that the particular form of
the function is not important. Accordingly there is a history of researchers using simpler functions for the repulsive energy, particularly inverse
power-law functions of the form
E(r) = ar n
where a is a constant for a particular ion pair, and n is a large integer
chosen often arbitrarily. A common case, for reasons that will be clear
shortly, n = 12 is often used.
These two functions, using n = 12 in the inverse power law function,
are compared in Figure 4.2, scaled to pass through a common point
and such that they have the same value of the first derivative at that
point. The two functions look remarkably similar, but that is a bit of a
trick; the two parameters of the BornMayer function have been tuned
to fit with the one-parameter inverse power law function, but it would
be much harder to work the other way round and get a good agreement by fitting the parameter of the inverse power law function to an
arbitrary parameterisation of the BornMayer function.

Energy

E(r) = E0 + B exp(r /)

3
2
1
0
0.8

1.2

1.4

FIGURE 4.2 Comparison of the BornMayer and inverse power


law (n = 12) functions for the repulsive energy, scaled that both
equal 1 and both have the same value of the first derivative at
the point r = 1 (arbitrary units).
For many ionic crystals simply combining the BornMayer function
with the Coulomb energy gives an adequately good representation of
the crystal energy, provided the parameters in the BornMayer function
and the values of the ionic charge are suitable chosen or tuned (for example, to predict the correct crystal structure or physical properties
such as elasticity). We will discuss this point shortly.

90

Dispersion interactions
There are crystals for which the Coulomb interaction will not provide adequate binding. These include the rare-gas solids (Ne, Ar, Kr,
Xe), molecular crystals (for example, organic molecules or the molecules of the elements such as N2, F2 S8 and C60), and crystals of elements where parts of the crystal are not held together by covalent
bonds (such as the interactions between layers in the graphite structure, or between chains in Se). In many of these cases there will be
Coulomb interactions, but these will be weak. Thus the binding energy of these systems needs to come from another origin.

91

dipole moment on the second atom. The energy of this interaction will
be proportional to 12 /r 3, where 2 is the induced moment. Taking
account of our previous argument for the size of 2, we have an energy
that is proportional to 2 12 /r 6. When averaged over time, this is pro-

r
FIGURE 4.3 Two atoms whose electron density is polarised
(separated by distance r), where the darker shade represents an
excess of electrons over the the average, and lighter shade
represents a corresponding deficiency of electrons.
The most common interaction in this case is one called the dispersion
interaction (alternatively the dispersive interaction). This arises from fluctuations in the electron density in the interacting atoms.
Whilst the spatial average of the electrons in an atom will have a spherical shape, at any instance in time the charge distribution will not be
spherical. Indeed, at any instance in time the fluctuation in the charge
distribution will lead to a fluctuation electrostatic dipole moment. Although the time-average of this dipole moment will be zero, so that we
cannot obtain an interaction between two atoms merely on the basis of
the interaction between these fluctuating moments, there is a part of
the fluctuations within the two atoms that will be correlated.
Figure 4.3 illustrates this correlation. One atom is instantaneously polarised, generating an electric field. This field is seen by the second atom,
and the field will then polarise this atom further over the top of the
naturally fluctuating polarisation. The size of this polarisation will be
proportional to 2 1 /r 3, where r is the interatomic separation, 2 is the
electronic polarisability of the second atom, and 1 is the instantaneous
dipole moment on the first atom. There will be a direct interaction between the fluctuating dipole moment on the first atom and the induced

portional to 2 12 /r 6.

There are two things about this argument. First, the minus sign follows
from the fact that two aligned moments have a negative (attractive) energy. Second, the square of the average is necessarily of positive value,
and thus the overall energy of this interaction is always negative and
hence attractive.
We also remark that this energy is proportional to the square amplitude
of fluctuation of the moment. This in turn is proportional to the electronic polarisability (you do not expect the amplitude itself to be proportional to the polarisability because the amplitude will have sign and
the polarisability is a fixed constant value). Hence the energy is proportional to 12 /r 6.
It follows that if we can write the dispersion energy in the form Cr 6,
the coefficient C can be said to follow a combining law given its proportionality to the produce of the polarisabilities of the two atoms. If
we have atoms labelled i and j, it is likely to be the case that
Cij =

CiiCjj . On this basis, if the coefficients for a set of atoms are to

be determined empirically rather than by calculation we have a constraint to reduce the number of parameters; if we have a set of N different atoms, there will be N(N 1)/2 terms to be determined, but exploiting this constraint reduces the number of parameters down to N.
Researchers often combine the dispersion interaction with the repulsive
interaction to create functions that are tuned together. One example is
the Buckingham interaction:
92

E(r) = B exp(r /) Cr 6
This function is popularly used, in conjunction with electrostatic interactions, in the study of inorganic materials and crystals formed from
small organic molecules. In fact the coefficients B, and C can be
tuned to work well for a wide range of organic crystals, giving a predictive power to such models.

E(r) = 4

[( r ) ( r ) ]
6

E/

Another popular form is the Lennard-Jones function

12

This function has been written in a particular way to highlight the fact
that the combination of the +ar 12 and C/r 6 terms leads to a particular interpretation. The constant represents the energy of a bond, and
the constant characterises the equilibrium separation of the two atoms, which we write as r0 = 21/6. It is easy to show that = (a /C )1/6
and = C 2 /4a. This function tends to be used when there is a strong
desire to reduce the number of variables. Given that is fixed by the
size of the atom, the only free parameter is , and with the combining
constraint the problem is made even easier.
We compare the Buckingham and Lennard-Jones potential in Figure
4.4, by assuming the minima and and the energy at the minima are the
same, and that both have the same dispersion energy. Thus the value of
C in the Buckingham potential is equal to C = 4 6. The energy at the
minimum of the Buckingham potential can be written as
= B exp(r0 /) 2
which means that B exp(r0 /) = . The minimum of the Buckingham
potential is obtained as the solution of

0.8

1.2

1.4
r/

1.6

1.8

FIGURE 4.4 Comparison of the Lennard-Jones (black curve) and


Buckingham (red curve) potential tuned to gave the same
minimum point and the same dispersion energy term.
E
r

=
r=r0

B exp(r0 /) + 6Cr07 = + 12 = 0

r0

from which we obtain = r0 /12. Thus we obtain the plots for these two
functions shown in Figure 4.4, with (by construction) the same depth of
the potential energy functions and the same behaviour at larger atomic
separations, but different behaviour at short distances. In fact the Buckingham function is badly behaved at short distances because the dispersion energy will tend towards a value of as r 0 but the repulsive
93

energy will only tend towards the value of B. Quantitatively the main
difference between the two functions is in the value of the second differential of the energy at the minimum. This is 20% larger for the
Lennard-Jones potential than for the Buckingham potential when the
Buckingham potential is tuned against the Lennard-Jones potential. The
second differential is the quantity that determines the forces constants
associated with atomic vibrations and properties such as elastic constants. And as we remarked when we discussed the repulsive interactions, here we are using the greater flexibility of the Buckingham
model (3 parameters) to fit the two-parameter Lennard-Jones model; it
would not be so easy to get as good a fit by tuning the parameters of
the Lennard-Jones function to give best agreement with an arbitrary
Buckingham function.
It should be noted that when we combine these functions with Coulomb interactions, the concept of a minimum in the function becomes
meaningless, since the minimum we require in the crystal is the minimum with respect to the crystal energy, not to specific bond parameters. Each atom will be at a minimum of the potential energy it sees
(that is the definition of the state of equilibrium, and corresponds to the
state of experiencing no force), but this potential energy will consist of
contributions from more atoms than the nearest neighbour atoms. Thus
long-range interactions that push the crystal inwards to reduce volume
will also force the nearest-neighbour atoms to be closer than they
might ordinarily like to be.

94

Hydrogen bonding
There are many crystals containing hydrogen where the interactions
considered this far are inadequate to explain the stability of the crystal structure.
Many examples are found in molecular crystals where there are
cases when a bond containing hydrogen, such as NH, will align
with a feature in another molecule. For example, the crystal structure of the molecule imidazole, C3N2H4, has an arrangement of atoms such that the NH bond in one molecule points towards a nitrogen atom in the neighbouring molecule that does not have its own
bonded H atom. This effect is primarily electrostatic in origin, because the moments along the NH bond in the first molecule are
95

INTERACTIVE 4.1 Crystal structure (monoclinic lattice) of imidazole,


C3N2H4, showing the hydrogen bond in which the NH bond on
one molecular is aligned with the electrons on the other nitrogen
atom in a neighbouring molecule.
aligned with the cloud of electronic charge on the nitrogen atom of the
second molecule that is mimicking a bond (so-called lone pair electrons). Similar bonds are found in polymers and form part of the interaction between the base pairs in the DNA molecule.
A similar effect occurs for water molecules, for example in the crystal
structures of ice. The distribution of electrons around the oxygen atom
in the water molecule have a tetrahedral shape, with two lobes lying
along the OH bonds, and the other two pointing out of the molecule.
Thus crystalline water (ice, but beware that there are over a dozen dif-

INTERACTIVE 4.2 Crystal structure of phase-II of H2O ice


(rhombohedral lattice), viewed down the 3-fold axis. The bonds
within the water molecules are coloured grey, and the hydrogen
bonds between molecules are coloured pink. The network of
hydrogen bonds forms a three-dimensional tetrahedral framework
similar to the frameworks of silica, SiO4.
ferent phases of ice with rather different structures) has the molecules
aligned in tetrahedral coordination with neighbouring molecules.
A rather different type of hydrogen bond is found in cases where a proton binds to two anions. In many cases it is energetically favourable for
the proton to lie closer to one anion rather than being situated at the
mid point of the bond, but typically the effect is weak and at high temperature the position of the proton is disordered. One key example is
96

INTERACTIVE 4.3 Crystal structure of the disordered phase of KH2PO4


(tetragonal lattice), showing two positions (small pink spheres) for
the hydrogen atoms between the oxygen atoms on neighbouring
PO4 tetrahedra.
the ferroelectric KH2PO4, where the protons are located along OO
vectors joining two PO4 tetrahedra.

97

Lattice energy
Introduction
We now define a quantity called the lattice energy, which is equivalent to the energy of the crystal structure with all atoms at rest, essentially the potential energy of the classical crystal at a temperature
of zero. Note the irony of the fact that having gone to some trouble
in this book to ensure that we use names correctly, the conventional
name for the crystal potential energy has confused the words crystal
and lattice (a similar confusion occurs for the topic of lattice dynamics, which really concerns waves travelling through the crystal with
the lattice itself having no dynamics). Thus the alternative name cohesion energy is sometimes used.
98

The lattice energy is defined easily in the case where it can decompose
into interactions between pairs of atoms that are dependent on the distance between atoms. In this case, we can denote two atoms by the labels i and j, their separation as rij and their interaction energy as ij(rij ),
the lattice energy is given as
W=

1
ij(rij )
2
ij

where the factor of 1/2 is used to cancel the fact that the sum involves
counting all pairs twice. The sum is over all atoms in the crystal and
over all neighbours. In practice the sum over all neighbours is very sensible, but the sum over the crystal is normalised to a fixed quantity
such as one mole of formula units.

Role of the lattice energy


If we can neglect the effects of temperature and quantum fluctuations
(zero-point motions), the equilibrium structure is that for which the lattice energy is a minimum. This means firstly that of all possible structures (for example, the NaCl, CsCl and other structure for a simple binary oxide) the one with the lowest lattice energy will be the stable
structure. And secondly, the minimum of the lattice energy also defines
the equilibrium values of the lattice parameters and atomic coordinates
where these are not fixed by symmetry.
The lattice energy is routinely calculated by two methods. First, the sort
of models we have discussed can be tuned for a specific chemical composition and used in a calculation of lattice energy. Second, it is now
quite feasible to calculate the bonding energy directly via modern implementations of the quantum mechanics for many electrons.

TABLE 4.2 Crystal data for the rare gas crystals


Neon

Argon

Krypton

Xenon

a ()

4.466

5.313

5.656

6.129

W (kJ/mol)

1.88

7.74

11.2

16.0

In addition to giving information about the crystal structure, a calculation of the equilibrium crystal structure can be used to calculate physical properties. Equations exist to obtain properties such as dielectric,
piezoelectric and elastic constants from appropriate differentials of the
lattice energy with respect to ionic displacements and lattice strain.
The lattice energy, and its relationship to the equilibrium structure and
physical properties, is often used to tune models of interatomic forces
(namely to tune values of ionic charges and the parameters in the Buckingham model say), on the basis that the best model should be capable
of reproducing the structure and experimental data for properties.
Whilst this approach carries some risk (your model is designed to get
the right answer), it nevertheless can be useful to derive predictive models, for example for organic crystals.

Example: lattice energy of rare gas solids


The rare gas solids, namely of the elements Ne, Ar, Kr and Xe, have a
ccp structure. Experimental data for these materials are given in Table
4.2. We compute the lattice energy using the Lennard-Jones potential
in order to illustrate some general principles (we have to use the
Lennard-Jones model because we only have two items of data for each
material).

99

The interatomic distance in a ccp crystal is equal to a / 2, and in terms


of the Lennard-Jones parameters is equal to 21/6. Thus = a /22/3, values for each material are easily deduced from Table 4.1. Evaluation of
the values of follow directly from the lattice energy. If we assume
nearest-neighbour interaction only, from our earlier discussion of the
Lennard-Jones potential we see that the lattice energy is calculated by
adding the number of bonds per mole of atoms and multiplying by .
In a ccp structure, each atom has 12 neighbours, but the energy of
each bond is shared by the two atoms within the bond. Thus the lattice
energy per mole of atoms is equal to 6NA. Expressed in unit of kJ/
mol, the value of is simply equal to the value of the lattice energy
given in Table 4.2 divided by 6.

Example: lattice energy of ionic alkali halides


with the NaCl structure
We work with a simple mode, namely charged ions and a BornMayer
repulsive interaction. In the NaCl structure, each ion has 6 neighbours
of the opposite charge. If we choose to calculate the lattice energy per
mole of formula units, we can compute the BornMayer energy from
one ion type only because we will automatically capture the energy of
the second ion and avoid double counting. Thus the lattice energy per
mole of formula units is equal to
W = WC + WR = NA

q
+ 6B exp(a /2)
40 a /2

where we use the fact that the nearest-neighbour distance is a /2, and
where q is the charge on an ion (the cation and anion have equal but
opposite charge values, with the sign taken account of through the leading minus sign).

TABLE 4.3 Experimental data for the set of alkali halide crystals with
the NaCl structure: a is the lattice parameter, W is the lattice
energy, and K is the bulk modulus.

Li

Na

Rb

Cl

Br

a ()

4.028

5.140

5.502

6.000

W (kJ/mol)

1014.3

832.6

794.5

743.9

K (GPa)

67.1

29.8

23.8

17.1

a ()

4.634

5.640

5.978

6.474

W (kJ/mol)

897.5

764.4

726.7

683.2

K (GPa)

46.5

24.0

19.9

15.1

a ()

5.348

6.294

6.596

7.066

W (kJ/mol)

794.5

694.0

663.5

627.5

K (GPa)

30.5

17.4

14.8

11.7

a ()

5.630

6.582

6.890

7.342

W (kJ/mol)

759.3

666.8

638.8

606.6

K (GPa)

26.2

15.6

13.0

10.6

The equilibrium value of the lattice parameter is determined by the condition W/a = 0. This is written as
W
= 0 = WC /a WR /2
a
It thus follows that
=

1
W
a 1
2 (
WC )

100

FIGURE 4.5 Lattice energy (W) and


bulk modulus (K) data for the alkali
halides plotted as a function of the
lattice parameter (a). The curves are
guides to the eye. The fact that all
data lie on a general curve indicates
a high degree of consistency
between the different compounds.

70
60

K (GPa)

W (kJ/mol)

50
40
30
20
10
4

4.5

5.5
6
a ()

6.5

7.5

0
4

4.5

If we know the value of the ionic charge and if we have experimental


data for the lattice energy and lattice parameter, it is straightforward to
obtain a value for the BornMayer parameter and subsequently for
the parameter B. Experimental data for the set of alkali halides with
NaCl structures are given in Table 4.3.
The bulk modulus is a measure of the stiffness of the material when under uniform compression. It is defined as
V
K=V
( P )

where V is the crystal volume and P is the pressure (note that various
texts use either B or K as the symbol for the bulk modulus, and we use
K to avoid confusion with the parameter in the BornMayer function,
but its inverse, namely the bulk compressibility, is sometimes also
given the symbol K). This gives a measure of the extent of volume reduction under pressure, with smaller volume changes corresponding to

5.5
6
a ()

6.5

7.5

large values of B. To put this into what will be a more useful form, we
use the relationship at zero temperature P = W/V to give
2W
K=V
V 2
To proceed, we will assume we can use formal charges for the ions. At
this point we should ask a question of the consistency of the data,
since the use of formal charges throughout relies on this. The data for
the lattice energy and bulk modulus are plotted against lattice parameter irrespective of ion type in Figure 4.5. The fact that the data fall onto
common curves suggests that we can treat all alkali halides within the
same model framework, which in turn relies on being able to use the
same charges for all compounds.
It is straightforward to use a spreadsheet or similar work package to determine values of and B for the model, but as numbers they are simply parameters of a model are of limited interest. What is more interest101

The calculated value of K for each system is compared with the experimental data in Figure 4.6. You can judge for yourself whether the agreement is good; personally for such a simple model with just two parameters per system, I feel that the agreement is really quite good, particularly given that different data sources give slightly different values for
the experimental data of the lattice energy and bulk modulus, and I suggest that the model captures the essence of the bonding in these materials.

70
60

Kcalc (GPa)

50
40
30
20
10
0
0

10

20

30
40
Kexpt (GPa)

50

60

70

FIGURE 4.6 Comparison of calculated and experimental values of


the bulk modulus for the suite of alkali halides with the NaCl
structure. The straight line has a gradient of 1.

The same can be done for many other systems, and indeed this forms
the basis of modern computational materials physics and chemistry. Different systems may require a different level of sophistication, for example accounting for charge differently, allowing electronic polarisability,
and allowing for the effects of covalency on specific bonds and to account for bending of bonds. For some materials (and the perovskites
are one particular case) such simple modelling cannot capture all the
interesting science, particularly when the phenomena are concerned
with subtle electronic effects.

ing is to test whether these models can predict the experimental values
of the bulk modulus. We rewrite the equation by noting that the volume of one mole of formula units is V = NA a 3 /4. It this follows that
4 2W
K=
9NA a a 2
It is straightforward to show that
2WC WR
2W
=
+ 2
a 2
a2
4

102

CHAPTER 5

Diffraction

Radiation beams
Preamble
We know from school science that light is diffracted when it passes
through a mask cut with lines and holes, and that each type of mask
has a characteristic pattern. This phenomenon is not due to the light
per se, but arises from the fact that the wavelength of light is not too
dissimilar to the physical dimensions of the lines and holes in the
mask. On this basis we can imagine that any radiation with a wavelength that is of the order of the spacing between atoms might be
diffracted in a way that is characteristic of the arrangement of atoms. Thus, provided we can interpret the diffraction pattern, diffrac104

tion of radiation by a crystal might give us a tool by which we can


learn something about crystal structures.

GALLERY 5.1 Examples of laboratory x-ray diffractometers

So we first need to be able to generate beams of radiation with the appropriate wavelength. In terms of electromagnetic radiation, given that
the typical interatomic separation is of order 12 , the radiation we
need is within the x-ray spectrum. X-rays were discovered in 1895 by
Wilhelm Rntgen, observing them as generated when a beam of electrons of sufficient energy struck a target. This method of production of
x-rays was quickly refined and improved. In 1912 the first diffraction of
x-rays from a crystal was reported by von Laue and co-workers, and in
the following year the father and son team of William and Lawrence
Bragg interpreted the diffraction patterns and turned the observation
into a scientific method.
Many universities and research laboratories now have x-ray diffraction
facilities, optimised for measurements from single crystals and from
powder samples, all based on the idea of

Powder diffraction instrument

105

FIGURE 5.1 View of the UKs Rutherford Appleton Laboratory, showing


the ISIS neutron and muon facility in the centre and the Diamond
synchrotron radiation facility as the large silver ring to the right.

106

Scattering of radiation from particles


Some general ideas
We start by considering the scattering of a wave from a single point.
For the moment we consider that the wave is scattered equally in
each direction, although we can relax this assumption later (for example, the scattering of electromagnetic radiation by an electron
will have an angular dependence on the intensity of the scattered
beam). We will consider only elastic scattering, in which the radiation beam does not have its energy changed through the scattering
process.

107

Define origin
ki

ks

= r cos
2

= r cos

Position r
FIGURE 5.2 Path of radiation beam scattered by two particles.
The radiation beam is defined as a wave with a given direction and
wavelength, both of which are encapsulated in the quantity called the
wave vector, which is given the symbol k. This is defined such that the
vector is aligned with the direction in which the wave front travels, and
the modulus of the wave vector is defined by the wavelength:
2
k =

For several reasons, as we will see, wave vector is a much more useful
quantity than wavelength.
What we are more interested in is how the scattering from a collection
of particles causes interference effects, because interpreting these interference effects is the key to understanding structure. We start by considering the scattering of two particles. What we chose as an origin is arbi-

trary, so we place one particle at the origin and one at a position r relative to the origin. The situation is encompassed in Figure 5.1, which
shows the path when the wave is scattered by two particles through the
same angle. In this figure, one beam travels further than the second
beam. If the incoming beam has a common wave front, the scattering
from two particles at random orientation to the wave front will introduce a change in path length for one particle relative to the other.
The additional path length in Figure 5.1 is equal to 1 + 2. If we define
the wave vector of the incoming beam as ki, the angle identified in
Figure 5.1 and hence the value of 1 can be obtained from the vector
dot product ki r:
ki r =

2
2
r cos =

Similarly we can define angle and the value of 2 from the dot product with the scattering wave vector ks:
ks r =

2
2
r cos(180 ) = 2

The change in phase of the wave is equal to


2
((1 + 2) = (ki ks) r

The important variable here is neither the wave vector of the incoming
wave nor the wave vector of the scattering wave, but the difference between these two wave vectors, and thus we give this its own symbol
Q = ki ks

108

and we note that the phase difference from the scattering from the two
particles is equal to Q r. Thus the scattering of the wave from these
two particles will be moderated by the factor
F(Q) = 1 + exp(iQ r)
If we have a collection of particles, each with position rj relative to
some origin, the modulation in amplitude of a wave that is scattered is
given as
F(Q) =

exp(iQ rj )

Here we have explicitly assumed that the scattering from each particle
is the same. We can account for this by a simply per-particle multiplicative factor:
F(Q) =

fj exp(iQ rj )

The quantity Q is called the scattering vector. To link it back to the


wave vectors of the incoming and scattered wave we calculated its
modulus-squared value
Q

= ki

+ ks

4 2
2ki ks = 2 (2 2 cos )

where is the angle subtended by the vectors ki and ks and is equal to


the angle through which the wave has been scattered. Using the trigonometric identify 1 cos = 2 sin2(/2) we can write
Q =Q=

4
sin

where we define = /2. We will use the angle a lot.

Scattering from a continuous distribution of


particles
If rather than being able to define a set of point particles we have the
case where there is a density of particles that in some way varies in
space, we could do the trick of pixelating space and treating each pixel
as a point. This works if the pixels are considered to be infinitesimally
small. Thus we define a density function (r), and we select a set of
small volume elements dV. The scattering from each volume element at
position r will be proportional to (r)dV; this corresponds to the perparticle multiplicative factor fj introduced above.
In this case, the sum over all particles becomes instead an integral over
all positions of the volume element, and we can simply obtain the scattering modulation factor as
F(Q) = (r) exp(iQ r) dV

Where this function is important is in considering the scattering of xrays from the electrons in an atom. We are not able to identify specific
positions for the electrons, but instead the wave functions define a function for the electron density. This function gives us one important limiting case, namely forward scattering where = 0 and Q = 0:
F(0) = (r) dV

In the case of scattering of x-rays by the electrons in an atom, the amplitude of forward scattering is simply equal to the integral of the electron
density and hence to the number of electrons multiplied by the charge
of the electron.

109

Scattering as a Fourier transform


Particles and Fourier transforms
Writing the scattering equation in the form
F(Q) = (r) exp(iQ r) dV

is nothing more than writing the scattering amplitude as the Fourier


transform of the density of the object that the radiation is being scattered from, where Q is the Fourier variable associated with r . This
point is important, because we now have a well-known mathematical tool with which to understand scattering theory.
110

In the case of individual point particles, we can represent the points as


densities through the use of the Dirac delta function, defined as
(r rj ) = when r = rj

We can represent the density of the collection of particles as


(r) =

(r rj )

(r rj ) = 0 when r rj

The Fourier transform of this density function is thus given as

Thus the function is defined as having a non-zero value only when the
position is the same as the position of the particle. Since the particle
exists with probability 1, we are consistent with the value of the integral of the Dirac delta function:

F(Q) =

formalism.

(r) dr = 1

We next need to know the Fourier transform of the Dirac delta function. This is best evaluated in the infinitesimal limit, where separate the
Fourier integral into the region of space immediately in the vicinity of
the point, where we can say that the factor exp(iQ r) does not vary
with r, and the rest of space. In the rest of space the integral is equal to
zero, so we only need to consider the integral within the vicinity of the
point. This is best illustrated in one dimension, where we consider the
variable x instead of r:

(x xj ) exp(iQx) dx =

xj +

(x xj ) exp(iQx) dx

= exp(iQxj )

xj +

(x xj ) dx = exp(iQxj )

Hence in three dimensions we can write


F(Q) = (r rj ) exp(iQ r) dV = exp(iQ rj )

(r rj ) exp(iQ r) dV =

exp(iQ rj )

This brings us back to our starting point. And if we wanted to weight


our points by a factor fj, this possibility is easily incorporated into the

Convolution
To exploit the use of Fourier transforms to study scattering from crystalline materials, it turns out to be useful to define the process of convolution. This is used extensively in defining a crystal structure, in that the
crystal structure can be defined in terms of placing an identical set of
atoms into the space associated with each lattice point (the unit cell).
This process can be identified with the convolution operation, which
formally is the process of replacing all points of one function with a
copy of a second function to obtain a third function. This is written
mathematically as
f (x) = g(x) h(x)
and is defined in one dimension such that
f (x) = g(w) h(x w) dw = g(x w) h(w) dw

111

Convolution theorem
The convolution theorem states that the Fourier transform of a convolution is the product of the Fourier transforms of the separate functions.
The Fourier transforms are defined as
F(q) = g(x) exp(iqx) dx

and similarly for G(q) and H(q). We can write the Fourier transform of
the convolution as
F(q) =

g(w)h(x w)exp(iqx)dwdx

which we can re-write as


F(q) = g(w) h(x w)exp(iqx)dx dw
(
)

We can now define a change in variable: z = x w $z=x-w$, so that


dz = dx . Thus
FIGURE 5.3 The convolution process in action, where we
replaced each point of the first function by a weighted copy of
the second function.
The process is that we place a copy of one function at each point of the
second function, weighted by the value of the second function at that
point, and then we sum over all the copies of the first function through
the integral. It is illustrated in Figure 5.2, where we show how we replace one function by a weighted copy of the second function point by
point, then just add all copies as an integral.

F(q) = g(w) h(z)exp (iq(z + w)) dz dw


(
)

We can rearrange this as


F(q) = g(w)exp(iq w) h(z)exp(iqz))dz dw
(
)

which can be further rearranged as


F(q) = g(w)exp(iq w)dw h(z)exp(iqz))dz

112

For example, suppose you perform a measurement of a spectrum that


is broadened by a function that defines the resolution of the instrument. This is a convolution operation. If you know this function, you
can recover the spectrum you want by taking the Fourier transform of
the measurement, diving by the Fourier transform of the resolution function, and doing the reverse Fourier transform. Well, it may not always
be quite as simply as this sounds, because you have effects of noise
and the like, but the principle is sound.

sin(qa)/qa

This is simply the expression of F(q) = G(q) H(q). This is a fantastically useful theorem in science for a wide range of applications.

Fourier transforms and size


It is useful to have a feel for how Fourier transforms work in practice.
We consider a simple one-dimensional rectangular function (sometimes called a slit function or a box function) defined as
f (x) = 1/2a when a x + a
f (x) = 0 for all other values of x
This function has an integral of 1. Its Fourier transform is
1 +a
F(q) =
cos(qx) dx
2a a
where we only need the real part of the complex exponential because
the function is symmetric around x = 0. The solution is
F(q) = sin(qa)/qa
This function is illustrated in Figure 5.3, and is colloquially known as
the sinc function. The function has zero values when the argument is

qa
FIGURE 5.4 The sinc function.
equal to multiplied by an integer, that is when q = n /a where n in
an integer, except for the case n = 0, where instead the function has a
value of 1 . The width of the sinc function might be defined as twice
the point at which it has value of 1/2. This has value q = 1.207 /a.
The point is that the width of the Fourier transform is proportional to
1/a, where we recall that a is the width of the original function. Thus if
we have one function with a peak, its Fourier transform will also have a
peak but with a width that is inversely proportional to the width of the
first function.

Scattering of x-rays from the electrons in an


atom
We now go back to the idea of the scattering of radiation from a collection of particles, and explicitly consider the scattering of x-rays from
113

10

Scattering factor

Si4+ cation
Al3+ cation

Mg2+ cation

4
2
0
0.0

F anion

0.5

1.0

1.5

sin /
FIGURE 5.5 Amplitude of the scattering of x-rays from four
isoelectronic ions
the electrons in an atom. From what we have done so far, we can immediately say that the scattering of x-rays from an atom will give the
Fourier transform of its electron density (after accounting for the angular dependence of the scattering of electromagnetic radiation from a
charge particle, and accounting for any polarisation of the x-ray beam).
In Figure 5.4 we show the amplitude of scattering of x-rays from four
isoelectronic ions as a function of sin / = Q /4. The extent to which
the scattering functions are spread out in Q is determined by the charge
on the atomic nucleus, with the F anion having the lowest nuclear
charge and the Si4+ cation having the largest. The larger charge means
that the electron orbitals are closer to the nucleus, as reflected in the
ionic radii given in the Appendix. Thus Si4+ is the smallest ion of the
four, and thus we expect its Fourier transform to extend further.
114

Scattering from a crystal


Preface
We now come to consider what happens when we scatter radiation
from a crystal. We have in place all the tools we need to do this
properly and with a minimum amount of effort.
We have already remarked that the crystal can be considered as a
convolution of the lattice and the motif. In turn, for x-ray scattering
the atom must be considered with its electron rather than just as a
point in the motif; thus the atom in the motif is a convolution of the
atom position and the distribution of electrons. Finally we know
that atoms are vibration about their average positions, and so the
115

last convolution is that of the atoms position in the motif and the
spread of positions arising from thermal motion.
Thus formally we can write this as Crystal Structure = Lattice Motif
Atomic Electron Density Thermal Motion. The scattering of x-rays
will therefore be given as
F(Q) = FT(Lattice) FT(Motif) FT(Electrons) FT(Atomic Motions)
Our problem is made easier by the fact that we simply calculate the
Fourier transforms of the separate components and multiply them together.

Fourier transform of the lattice


The lattice can be defined mathematically as the infinite collection of
points defined by three lattice vectors:
L(r) =

U,V,W

(r (Ua + V b + Wc))

Its Fourier transform is equal to


R(Q) =

U,V,W

(r (Ua + V b + Wc)) exp(iQ r) dr

We use the property of the Fourier transform of the Dirac delta function
discussed previously to reduce this function to
R(Q) =

U,V,W

cos (Q (Ua + V b + Wc))

where the imaginary component cancels because the lattice has a centre of symmetry. The sum is over an infinite number of terms, and will
only have a non-zero value for a set of values of Q that cause the argu-

ment of the cosine to be equal to 2 multiplied by an integer. To see


how this can work, we define a new vector d*hkl as a combination of
three vectors in the three-dimensional Fourier space:
d*hkl = ha* + kb* + lc*
And consider the situation where Q = d*hkl. At the present time we

know nothing of these three vectors a*, b* and c*, but we label them
as such by analogy with the crystal lattice vectors a, b and c. We thus
require that
(ha* + kb* + lc*) (Ua + V b + Wc) = n 2
for all values of U, V and W, which in turn will set constraints on values
of h, k and l and on the vectors a*, b* and c*.
Now we need to decide how to work with these new quantities. Let us
set the case that V = W = 0. Thus the phase relation gives us
(ha* + kb* + lc*) Ua = n 2
We can define the direction of the vectors such that b* a = c* a = 0.
This means that we would be left with
hUa* a = n 2
We are free to make whatever definition we like, so we could set
a* a = 2. This means that h U must be an integer (any integer), and
this can only be achieved if h in turn is an integer. We can perform a
similar analysis for the other quantities, such that
a* a = b* b = c* c = 2
and that all other dot products, such as a* b and c* a are all equal to
zero. In turn this means that the vector a* is perpendicular to both b
116

and c, b* is perpendicular to a and c, and c* is perpendicular to a and


b. Any vector perpendicular to two other vectors is parallel to the vector cross product of these two vectors. Hence we can write that
a* b c

b* c a

c* a b

Since we have our normalisation condition, namely a* a = 2, it follows that


a* = 2

bc
a (b c)

b* = 2

ca
a (b c)

c* = 2

ab
a (b c)

The denominators are all the same because the combinations of products are all the same and equal to the volume of the unit cell.
Now we consider the general case:
(ha* + kb* + lc*) (Ua + V b + Wc) = 2(hU + kV + lW )
which is an integer multiplied by 2.

Interpretation: the reciprocal lattice


The result of the Fourier transform of the lattice is nothing other than a
new lattice with basis vectors a*, b* and c*, where the lattice points
are defined as the set of values of d*hkl, where the values of h, k and l

FIGURE 5.6 Example of an electron diffraction image, showing


scattering only at well define points as described the the reciprocal
lattice.

We stress that the reciprocal lattice is not something condensed matter


scientists have invented for convenience. The space of the reciprocal
lattice, namely reciprocal space, is the space inhabited by the vectors

of waves, be they the waves that are diffracted in an experiment designed to measure the crystal structure, or the waves of atoms vibrations (including sound), or the waves of electrons in metals. The reciprocal lattice vectors are special vectors in this space. Here we are beginning to see that they are the vectors associated with diffraction, but for
vibrations and electrons they are vectors where we find a new translational symmetry for the wave vectors.

are restricted to the set of integers. This new lattice is called the reciprocal lattice, the name chosen perhaps because it is the Fourier inverse of
the original lattice, and perhaps because the scale of the vectors in this
space have dimensions of 1/length.

117

FIGURE 5.8 Representation of a crystal plane projected onto the


vectors of the unit cell.
sponds to a point in the reciprocal lattice, that is explicitly when
Q = d*hkl .

FIGURE 5.7 Photograph of x-ray diffraction on an instrument


designed to record an undistorted image of reciprocal space.

The reciprocal lattice, diffraction and planes


of atoms
The point of this analysis was to compute the Fourier transform of the
lattice to multiply by all the other Fourier transforms in order to understand the scattering of radiation from a crystal. The conclusion is simple; the scattering is zero for all values of Q except when Q corre-

This point is illustrated in Figures 5.6 and 5.7, which show photographs
recording the scattering of electrons and x-rays respectively from crystalline materials. The key observation is that scattering is limited to specific points in reciprocal space, which correspond to the points of the
reciprocal lattice.
We now relate the description of the reciprocal lattice to the arrangements of planes of atoms in the crystal. Consider the situation shown in
Figure 5.8. We say that this plane intersects the unit cell at distances
a /h, b/k and c/l along the three vectors of the unit cell, where h, k and
l are integers. There will be a parallel plane that passes through the origin, as part of an infinite set of planes. We can define vectors within
118

the plane that will of course have a zero dot product with the plane
normal. We consider one that connects two intercepts
r=

Incoming beam
of radiation

Scattered beam
of radiation

1
1
a b
h
k

If the plane normal is equal to


d*hkl = ha* + kb* + lc*

dhk

we have
r d*hkl =

1
1
a b (ha* + kb* + lc*)
(h
k )

From the previous definitions, we have the result that


r d*hkl = 1 1 = 0. This is true for all vectors within the plane. Thus we
have confirmed that the vector d*hkl is perpendicular to the plane.

If we define n*hkl = d*hkl / d*hkl as the normalised vector normal to the


plane, and an vector p as a vector from the origin to the any point in
the plane, n*hkl p = dhkl, where dhkl is the distance from the origin to the

plane, and hence the separation of planes. Taking p as any of the intercepts of the plane with the unit cell vectors, say p = a /h, we have
1
d*hkl

(ha* + kb* + lc*) a /h =

2
d*hkl

It thus follows that the spacing of planes denoted by the three indices
hkl, which we denote as dhkl, is given as
dhkl = 2 / d*hkl

180 2

FIGURE 5.9 Illustration of the Bragg equation. A beam is scattered


from planes of atoms.
If we go back to the association of d*hkl with the scattering vector Q, we

recall that we also obtained a value for the modulus of this vector in
terms of the radiation wavelength and scattering angle. Thus we can
write
dhkl = 2 / d*hkl = 2

4 sin

We can rearrange this to obtain


= 2dhkl sin
This is the famous Bragg law for diffraction, discovered by Lawrence
Bragg immediately after the first x-ray diffraction patterns were measured by Max von Laue and co-workers in 1912.

119

Miller indices
The set of integers h, k and l is used to define a particular planes in a
crystal, as per Figures 5.6 and Figure 5.8. These planes are not not just
a ingle plane of atom, but the complete set of parallel planes displaced
relative to each in a way that reflects the translational symmetry (Figure
5.10). The three numbers that define a set of planes have their own
name, Miller indices, and are also designated by the three integers we
have used. They are enclosed in round brackets to denote the set of parallel planes, (hkl ), with the set of equivalent planes being denoted by
curly brackets as {hkl}. For example, the (100) plane of a crystal with
cubic symmetry is part of the set {100} which includes also (010),
(001), and also (100), (010) and (001).

FIGURE 5.10 illustration of a set of planes here they are one of


the set {120} showing the planes separated by a symmetry
translation.
The Bragg equation is illustrated in Figure 5.9, where we see scattering
from planes of atoms. The additional path length in the figure is equal
to the wavelength for constructive interference. Taking one of the two
segments, we clearly have

/2 = dhkl cos(90 )
which immediately gives us the Bragg law. This derivation is rather
quicker than approaching the problem via the reciprocal lattice, but
we nevertheless need the reciprocal lattice for a complete analysis.

Actually the complete set of negatives do not make sense from one perspective, in that the set of planes designated by (h, k, l ) are equivalent
to those designated by (h, k, l ). What is different, however, is how we
define the plane normal. The vector normal to the planes is the vector
d*hkl, and for the set of planes (h, k, l ) the plane normal will point in the

opposite direction, namely d* = d*hkl. This is more than being pedanhkl

tic: the scattering of a beam of radiation from a crystal will follow as


the Fourier transform of the crystal structure, and the cases where
Q = d*hkl and Q = d* correspond to the beam being scattered in oppohkl

site directions.

Fourier transform of the motif


We now return to the main point, which is how to define the Fourier
transform of the crystal as a product of the Fourier transforms of the individual components. The motif consists of the convolution of the atom
positions and the function describing the electron density of each
120

atom, the latter which we have already discussed as being represented


by the function f (sin /). This is called the atomic scattering factor or
the x-ray form factor. The Fourier transform of the motif, where each
atom has position rj = xj a + yjb + zj c, is given as
F(Qhkl ) =

fj exp (2iQhkl rj) =

fj exp (2i(h xj + k yj + lzj ))

Thermal motion
Thermal motion arising from harmonic waves will give a Gaussian distribution of atomic displacements. The Fourier transform of a Gaussian
is also a Gaussian. In the case of harmonic waves, the thermal motion
in the isotropic approximation is described by a Gaussian function of
2
the form exp (uj2Qhkl
/2), where uj2 is the mean square displacement

of atom of label j. Note that we have previously remarked that


Qhkl = 4 /sin(hkl ) = 2 /dhkl.

of x-ray diffraction, the inverse Fourier transform will give the electron
density:
(x, y, z) =

hkl

F(hkl )exp (2i(h x + k y + lz))

In the case of neutron diffraction, the corresponding quantity will be


the nuclear density; although nuclei are effectively points on the length
scale of the crystal structure, these points will be smeared through the
effects of thermal motion.
Unfortunately we cannot measure the structure factor, but instead the
quantity that can be measured is the intensity, which is proportional to
2

F(hkl ) . By taking the square of the modulus we lose all information


on the phase of F(hkl ). For centro-symmetric structures the phase is
the sign of F(hkl ), but for structures without a centre of symmetry
F(hkl ) is a complex quantity. How we proceed is discussed in the next
section.

Putting it all together: the structure factor


When we perform the final multiplications of the components, we arrive at the result
F(hkl ) =

fj exp (2i(h xj + k yj + lzj )) exp (uj2 Q2hkl /2)

This quantity is called the structure factor, and provides the link between the structure and diffraction. Note that we now use the symbol
F(hkl ) in place of F(Qhkl ) following normal convention.
If we could measure a complete set of structure factors we could perform the inverse Fourier transform and obtain the structure. In the case
121

Finding the crystal structure


Determining the lattice parameters
The first task is to determine the lattice parameters, and this means
that we need to measure some lattice reflections and index them
with the correct Miller indices.

Symmetry of the crystal


A crystal that we can hold on the tip of our finger will contain of the
order of 1020 atoms. Because of translational symmetry we only
need to identify the positions of the atoms in the unit cell, which
may be a trivial exercise for the very simple structures, but these ex122

amples were solved in the very early days of crystallography. Most crystals do not have the positions of the atoms determined exactly by symmetry, but symmetry will help to reduce the number of coordinates.

which has a value of zero if h + k + l is an odd number. Similar relationships between the values of the Miller indices and non-zero structure
factors exist for other centred lattices.

We have previously identified a number of symmetry operations, some


of which involve symmetry about a point (centre of symmetry, mirror
planes, and rotation and rotoinversion axes). Many crystals also contain glide planes (mirror reflections coupled with a translation along a
direction of symmetry) and screw axes (rotations coupled with a similar type of translation). These symmetry operations can be combined in
only a fixed number of ways 230 to be precise, with the combination
called a space group and the first task in a crystal structure analysis is
to identify the space group. This can be determined by inspection of
the diffraction.

On this basis, the determination of the Bravais lattice is usually fairly


straightforward.

The first task is to identify the Bravais lattice. The unit cell parameters
will be able to identify the lattice types (triclinic, monoclinic etc), although there are cases when two unit cell lengths are coincidentally
indistinguishable, and when unit cell angles are coincidentally too
close to 90 to be sure that they are not of that value. The existence of
a centred lattice will be clear from the diffraction patterns because
some reflections will be systematically absent. We take as an example
a body-centred lattice. The structure factor based on the conventional
unit cell will contain terms of the form
exp 2i(h x + k y + lz) + exp 2i (h(x + 1/2) + k(y + 1/2) + l(z + 1/2))
This can be written as
exp 2i(h x + k y + lz) (1 + exp i(h + k + l ))

The second task is to identify the symmetry of the space group, having
significantly reduced the number of possibilities by identifying the Bravais lattice. The diffraction pattern will have a symmetry that reflects
the underlying crystal symmetry. For example, planes of relection and
rotation axes will be seen in the symmetry of the diffraction pattern.
Consider a crystal with a mirror plane perpendicular to the z axis, such
that an atom at position x, y, z will have a copy at x, y, z . The structure
factor will contain terms of the form
exp 2i(h x + hk + lz) + exp 2i(h x + hk lz)
Clearly if we reverse the sign of l we have the same result. Thus in the
case of a crystal with this mirror plane it follows that F(hkl ) = F(hkl ).
A similar result will follow for rotation axes.
Symmetry operations that contain a translation give unambiguous evidence of their existence in the form of systematic absences, namely
sets of structure factors that have zero value because of the symmetry.
Consider the case of a screw axis along the y axis, so that a position
x, y, z has an equivalent position x, y + 1/2,z. The structure factor will
contain terms of the form
exp 2i(h x + k y + lz) + exp 2i (h x + k(y + 1/2) lz)
When h = l = 0, this can be written as
exp 2i(k y) (1 + exp ik)
123

This will equal zero for all odd values of k. Similar conditions exist for
glide planes. From these systematic absences it is possible to identify
the presence of screw axes and glide planes. On this basis, coupled
with the symmetry of the diffraction pattern, it is possible to narrow
down the set of symmetry operations in a particular crystal.

This function gives us information about the vectors between pairs of


atoms, and it is sufficiently important that this function is known as the
Patterson function named after the person who first recognised it. In a
crystal, the shortest vectors are likely to be visible in a three-

We note however that there is one problem that will haunt us, and that
is that by definition the sets of structure factors has its own centre of
symmetry, namely that

interpret these peaks, and there will be many of them, but often we are
helped by symmetry.

F(hkl ) = F(hkl )
This is known as Friedels law. It arises from the fact that the complex
conjugate of the structure factor can be written as F*(hkl ) = F(hkl ),
and by construction

F(hkl )

= F(hkl ) F*(hkl ). In consequence,

we might expect to find some cases where we are left with some ambiguity.

The phase of the structure factor


The big challenge now is to go from measurements of F(hkl ) to the
absolute value of F(hkl ), noting that there is no experimental way to
determine the phase of F(hkl ) directly.
Historically there have been a number of approaches to tackle this
problem. The first is to exploit the meaning of what we actually measure. We can write the modulus square of the structure factor as
F(hkl )

jk

exp (iQhkl (rj rk ))

dimensional map of the function F(hkl ) . The challenge then is to

For example, suppose we have a crystal with a mirror symmetry perpendicular to the y axis. Thus an atom at position x, y, z will have an equivalent atom at position x, y, z. Thus the vector between these atoms will
lie at 0,2y,0, and the Patterson function should show a peak at this special position. Similarly, if the structure has a mirror plane such that an
atom at position x, y, z will have an equivalent at x, y, z, and the Patterson function will show a peak on the plane 2x,0,2z. Thus it will become possible to obtain estimates of the atomic coordinates from the
Patterson function. Although there will be, in principle, many peaks, it
turns out that in practice if the crystal contains one atom that is a lot
heavier than the other atoms (that is, it has many more electrons), the
Patterson function will be dominated by peaks associated only with
this atom. Given that this atom will contribute heavily to the value of
the structure factor, by identifying the positions of the heavy atoms it
will be possible to identify the phases of many of the stronger structure
factors.
Nowadays the Patterson function is not necessary because we have statistical methods called direct methods to identify the phases of
structure factors directly from the data. Discussion of these methods
takes us beyond the scope of these notes, but the gist is readily explained. The statistics of structure factors works best if we replace each
atom by an equivalent point atom, that is, we deconvolve out the elec124

trons and thermal motion by dividing by the average x-ray scattering


factor and an estimate of the effects of thermal motion.
We can illustrate the point by considering one early example. We use
the Cauchy-Schwartz inequality for two functions
2

f j2

gj2

fj gj

We write the normalised structure factor of a centro-symmetric crystal


structure as
UH =

nj cos(2iH X)

where we denote the Miller indices hkl as the set H, the coordinates
x yz by the set X, and the dot product H X = h x + k y + lz. nj represents the per-atom normalisation factor, and the overall normalisation
n = 1.
requires that
j
j

We apply the inequality by writing fj = nj1/2 and gj = nj1/2 cos(2iH X).


Thus

fj gj = UH

We also have

f j2 =

nj = 1

We can write

gj2 =

nj cos2(2iH X) =

1
1
nj (1 + cos(2i2H X)) = (1 + U2H)
2
2
j

Thus the inequality becomes


1
2
(1 + U2H) UH
2
Since the right hand side is necessarily positive, it follows that for the
case UH2 1/2 the sign of U2H must be positive.
Of itself this inequality is only of limited practical use, but it does suggest that it will be possible to use more detailed statistical analysis of
the distribution of values of structure factors to determine the phases of
some structure factors, and indeed this has proven to be the case. In
practice it is now straightforward to assign the phases of enough structure factors to be able to determine the crystal structure with good accuracy.

Refinement of the crystal structure


Crystallographers demand precision on the crystal structure, and this is
obtained using least squares methods. Typically the atomic coordinates
and parameters that describe the thermal motion are adjusted so as to
minimise the function
2

hkl

Fobs(hkl ) Fcalc(hkl )

where Fobs(hkl ) and Fcalc(hkl ) are the observed (that is, measured and
normalised) and calculated structure factors respectively. Note that at
this stage we do not explicitly need to know the phases of the structure
factors; these will be set by the crystal structure at each stage of the
least squares procedure.
125

Once this process is completed we will have a very good determination of the crystal structure, and with modern data good precision is
possible (third decimal place of the fractional coordinates at worst).

126

Glasses

What is a glass?
The title of this section assumes there is a generally accepted answer, but this is far from the truth. We have some general observations that can guide us to a pragmatic, if not ideal, answer.
A glass is an apparently solid object formed by rapid cooling of a
liquid. Quite how rapid will depend on the solid; a silicate liquid
can be cooled quite slowly and still form a glass (this is exploited by
glass makers), but a metal has to be cooled incredibly quickly to
form a glass phase.
A glass does not have a well-defined crystal structure, as can be
seen in a diffraction experiment. Whereas a crystal will give well128

defined reflections of a beam of x-rays or neutrons, a glass will give a


very broad diffraction pattern. Thus we can conclude that there is no
long-range order the atoms in the structure, but nevertheless there are
features in the diffraction that suggest some degree of atomic order.
The glass phase is established at a particular temperature on cooling,
but unlike other phase transitions the actual temperature is dependent
on the cooling rate. Nevertheless, there is a clear transition, and properties such as heat capacity and thermal expansion change at the transition point (on the other hand, there is not a change in density at the
glass transition).
So far we have simply characterised the main features of a glass, but
we have not yet said what a glass actually is. This question comes to
whether a glass is a real solid or a liquid that flows so slowly that its
flow cannot be observed on human time scales (it is often cited that
glass in medieval buildings is thicker at the bottom than at the top,
thereby suggesting flow over a long time, but this is a myth; medieval
glass is not of uniform thickness because of the manufacturing process,
but there is no correlation between thickness process and how the
glass was arranged within the window).
It should also be remarked that crystalline solids can also flow under
gravity. You need a very soft crystal to be able to see this, but there is
no reason why over time a crystal to lower its energy within a gravitational field by having atoms move over the surface to a lower energy
state.
It is quite easy to imagine that some materials can form states with no
long-range order but with a high degree of short-range order. Figure
6.1 shows a two-dimensional arrangement of atoms in which all bonds
are of similar length and where each atom has the same number of

FIGURE 6.1 Representation of the arrangement of atoms in a twodimensional continuous random network, showing no long-range
order but nevertheless a high degree of short-range order with a
clear average bond length and bond angle, and with each atom
having three nearest neighbours.
neighbours. It will follow that each atom will also have a reasonably
well-defined set of second and third neighbours too.
There are a number of well-understood glass-forming materials. The
most common is any material containing silicon, ranging from pure silicon dioxide, SiO2, which we have met before when considering crystals of simple formulae, to various compounds that could variously be
described in terms of the general formula (SiO2)x(Al2O3)y(XO)z(Y2O)1-x-y,
where X is a divalent cation such as Mg2+ or Ca2+, and Y is an alkaline
cation such as Na+ or K+. In this case the Si4+ and Al3+ cations form tet129

bonds is slow, and in part because the energy of the disordered state
is not much higher than the energy of the ordered structure.
Another important family of glasses are materials containing S and Se
anions, often called the chalcogenides. These materials also form covalent networks, albeit with lower dimension than the threedimensional networks of the silicate glasses.
A number of organic materials can also form glasses. In part this is
because the shapes of the molecules allow them to become trapped
in their orientations within the liquid state as they are cooled.

FIGURE 6.2 Glass fragility for a range of glass-forming materials


rahedral coordination with oxygen, and these tetrahedral units form an
infinite network as they do in the crystalline form. However, unlike the
crystalline form, these networks are randomly connected, and these networks can form without adding stress to the tetrahedra. These materials
form glasses because in the fluid state there is a high degree of shortrange order through the formation of the covalent tetrahedra, which in
turns causes a high viscosity (think of the slow movement of volcanic
lava). On cooling it is easy to retain this network without forming crystalline order, in part because the time scales for breaking and reforming
130

Atomic structure of glasses


With crystals, the atomic structure of the whole crystal could be decomposed into separate components namely the lattice and motif
which made description of the structure relatively easy. The lattice
describes the long-range part, and the list of atomic coordinates
within the motif is usually quite short. Moreover, the list of atomic
coordinates is significantly reduced when taking account of symmetry. We do not have this advantage in a glass.
However, it cannot be said to be at all useful to describe the structure of a glass in terms of a set of absolute positions. Such a list
would be impossibly long and impossible to determine, it would dif-

131

ferent for each sample, and it would contain no real information of


value without subsequent analysis.
Instead, the key useful feature is the structure over a short distance, not
the absolute structure, but a measure of the average structure. For example, in a silicate, the silicon cations will almost always be surrounded
by a group of 4 oxygen anions at a fixed distance (around 1.6 ), and
because the oxygens form a near-perfect tetrahedron the nearestneighbour oxygenoxygen distances will be around 8/3 times the Si
O distance. The SiOSi bond angle has a well-defined distribution
around 150, which will mean that the distribution of first-neighbour
SiSi distances is also well defined.
The key then is to have a way to describe the average local structure,
which comes down to a way of describing the distribution of interatomic distances from any atom type.
We start by defining a spherical shell around each atom, at distance r
and thickness dr. We consider one atom type, and consider the density
of atoms of a second atom type, at this distance, 2. The number of atoms of this second type within the shell is going to be related to 2. Naively we could write that the number of atoms in the shell is equal to
4r 2 dr, but this will be wrong because it assumes there is no correlation between the positions of atoms. For example, we know that the
second atoms cannot come closer to the first atom by a distance nominally given by the sum of the ionic radius. This might define a first shell
of neighbours, and then we know that the next shell of neighbouring
atoms cannot come closer to the first shell than another sum of ionic
radii. Thus we have to modify how we count the number of atoms.
We define a quantity called either the pair distribution function (pdf) or
the radial distribution function (rdf). This is give the symbol g(r), and is

FIGURE 6.3 Shells of neighbours surrounding a single atom in the


two-dimensional continuous random network.
the function that quantifies the degree of structural order around an
atom. Taking our example of one atom and its neighbours of a second
type, the number of atoms of type 2 found within a spherical shell of
thickness dr at distance r from atom of type 1 is given as
n(r)dr = 4r 2 2 g12(r)dr
For a random distribution of atoms, g12(r) = 1, and because at large distances we do not expect correlations between atom positions we expect g12(r ) = 1. On the other hand, in the limit of short distance,
namely r < r0, we expect g12(r < r0) = 0. Because we expect a welldefined first shell of neighbours, we expect g12(r) to have a peak
132

FIGURE 6.4 First shells of neighbours for several atoms in the twodimensional continuous random network.
around the first-neighbour distance. Indeed, if we integrate over a band
of distances than encompasses the first neighbour distance, we expect
R+

n(r)dr = 42r

R+

g12(r)dr = N1

where N1 is the average number of nearest neighbours, which have a


average distance of R and a spread of distances between R and
R + .
We can picture this using our model two-dimensional glass shown earlier. In Figure 6.3 we show shells of neighbouring atoms around one

FIGURE 6.5 Second shells of neighbours for several atoms in the


two-dimensional continuous random network.
central atom. In Figures 6.4 and 6.5 we show that these shells describe
the coordination of all atoms.
Thus the function g(r) characterises the short-range atomic structure in
a material, whether a glass, fluid or indeed even a crystalline material.
It has distinct peaks associated with shells of neighbours, with widths
that quantify the spread of neighbour distances. Furthermore, the extent of detail within g(r) characterises the extent of short-range order,
the range of which is determined by the point at which g(r) more-orless becomes equal to one.

133

g(r)
4

SiSi

4
2

T (r)

3
2

SiO
20

0
0

r ()

10

12

14

FIGURE 6.6 PDF function T(r) for silica glass, measured by neutron
scattering.
Figure 6.6 shows an experimental measurement of the PDF of amorphous silica. Actually the plot shown is not g(r) but a related function
T(r) = 4rg(r); the key point about this function is that it is proportional to g(r) multiplied by r. Hence at higher values of r, where
g(r) 1, T(r) becomes a straight line. The data were obtained from neutron scattering (we will discuss the methodology later).
The PDF shows a number of features at low r, which as seen in the
pure g(r) functions obtained by using the experimental data to obtain a
model for the structure. The first two distinct peaks are due to SiO and
OO nearest-neighbour distances. The average distances are read from

10
0
OO

6
4
2
0

6
8
r ()

10

12

14

FIGURE 6.7 PDF functions g(r) for separate atoms pairs in


amorphous SiO2
134

INTERACTIVE 6.1 Configuration of SiO4 tetrahedra in amorphous


silica.

INTERACTIVE 6.2 Configuration of an amorphous metal organic


framework, here with Zn in tetrahedral co-ordination with
imidazolate (C3N2H3) ligands.

the peak positions. The mean SiO (around 1.6 ) and O-O (around
2.6 ) distances are consistent with the existence of near-perfect SiO4
tetrahedra as noted earlier. The third peak corresponds to the first SiSi
stance. It is much weaker than the other two peaks because it is
broader, because the scattering factors are weaker for Si, and because
there are only half as many silicon atoms as oxygen atoms. It represent
a mean distance of around 3.2 , which is consistent with a mean Si
OSi angle of 150. Beyond this distance there is an overlap of various
peaks in the individual g(r) functions and indeed overlap of the different functions, which means that interpretation of the information
within the PDF is difficult without some modelling activity.
135

Measuring the pair distribution


function
The pair distribution function can be measured by similar diffraction
experiments that give information about crystal structure, except
that the analysis is rather different.
We recall the equation for the amplitude of scattering from a collection of particles:
F(Q) =

fj exp(iQ rj )

The quantity we measure is the squared modulus of this function:

136

S(Q) =

1
F(Q
N

1
fj fk exp (iQ (rj rk ))
N
j,k

where N is the number of atoms in the material. This function explicitly


contains information about the separation of atoms in the material (before, we noted that we had lost the phase factor of F(Q) when we
measured the intensity; here this doesnt matter because we are not after the positions of atoms directly).
By itself this sum over all neighbour distances is not particularly useful.
What we want to do is make use of the PDF, which contains information about the spread of interatomic distances. But first we should remove the explicit directional information contained within this formula. A glass will look the same from all directions, thus we really
want to work with the isotropic scattering vector Q = Q . To do this
we need to perform an orientational average of the function. We denote rjk = | rj rk | , and perform the integration over all angles between
Q and rj rk:

1 2
exp(iQ (rj rk ) = 4 d exp(iQrjk cos ) sin d
0
0

We substitute x = cos and hence that sin d = dx, and perform the
trivial integration over , to obtain
sin(Qrjk )
1 +1
exp(iQ (rj rk ) = 2 exp(iQrjk x) dx = Qr
jk
1
Hence we can write the scattering intensity as
sin(Qrjk
1
S(Q) =
fj fk
N
Qrjk
j,k

Note that this function includes terms where j = k, which we now separate from the scattering from pairs of different atoms:
S(Q) =

sin(Qrjk
1
1
f j2 +
fj fk
N
N
Qrjk
j
j,k

Now we can make use of the PDF. We define cm as the relative concentration of atoms of species type m; thus in SiO2 cSi = 1/3 and cO = 2/3.
Thus the first term simply becomes
1
f j2 =
cm f m2

N j
m
The second term is a double sum over all atoms. The first of these can
be replaced by a sum over atom types, since we are considering the average behaviour associated with each atom type. The second sum is the
sum over all neighbours, which can be replaced by scanning distances
and using the probability of finding the next atom. This probability is
defined by g(r), but we do need to be careful about what happens at
large distance. Therefore we use instead the probability 4nr 2 multiplied by the correlation g(r) 1, and the bit we subtract will be added
back in at the end. On this basis, we can write
sin(Qrjk
1
sin(Qr)
fj fk
= 4
cmcn fm fnr 2 (gmn(r) 1)
dr

N
Qr
Qr
jk
j,k
m,n
The bit we left out when we went from g(r) to g(r) 1 integrates to a
term that is only non-zero at Q = 0. Thus we can write
S(Q) =

cm f m2 + 4

m,n

cmcn fm fnr 2 (gmn(r) 1)

sin(Qr)
dr + S0
Qr

137

where S0 is the term that exists at Q = 0, which actually is never measured in an experiment.

0.4

sin(Qr)
4
cmcn fm fnr (gmn(r) 1)
dr = S(Q)
cm f m2 S0 = i(Q)

Qr
m,n
m
2

It follows by multiplying by Q and cancelling factors of r that


4

m,n

cmcn fm fnr (gmn(r) 1) sin(Qr) dr = Qi(Q)

c c f f g (r) 1) = Qi(Q)sin(Qr) dr
m n m n ( mn

m,n

The left hand function is given its own symbol:


D(r) = 4r

0.2
0.0

-0.2
0

By reverse transform we have


4r

i(Q)

It helps to rearrange this equation in order to focus on the middle term:

m,n

cmcn fm fn (gmn(r) 1)

The functions D(r) and Qi(Q) are the primary functions for measurements of the PDF. The i(Q) function for amorphous silica is shown in
Figure 6.8. Both D(r) and Qi(Q) oscillate in value about zero at high
values of r and Q respectively, which is achieved by the subtractions in
both cases. And because of the subtractions, the functions are negative
at low r and Q, and because both these functions contain a factor of r
and Q respectively they vary as r and Q at the start.
There is one very important point to note, which is always a problem
with performing Fourier transforms in practical terms regardless of what
sort of data we mean. In our case, we can measure Qi(Q) only over a

10

20
Q (-1)

30

40

FIGURE 6.8 Scattering function for amorphous silica, measured by


neutron scattering with the different curves (displaced vertically
for clarity) corresponding to different banks of detectors that
cover different ranges of values of Q.
finite range of values of Q, up to a maximum value Qmax and down to
a minimum value Qmin. In practice, this means that the function we
have is Qi(Q) for an infinite range of values of Q multiplied by the box
function, namely a value of 1 for Qmin Q Qmax. When we take the
Fourier transform, the result will be the true D(r) function convolved
with the Fourier transform of the box function, namely the sinc function. The lower the value of Qmax the broader the sinc function, and this
means that the peaks in D(r) will be broadened too much by the convolution, but worse than this, the ripples in the sinc function will also be
convolved into the extracted D(r) function and we will not be able to
identify real peaks amongst the rippled peaks.
The common solution is to multiply the Qi(Q) by a function of the form
M(Q) = sin(Q /Qmax)/(Q /Qmax), which is a sinc function. It means that
138

the function to be transformed has zero value at Qmax, and it significantly reduces the ripples in the transform. However, because of the
convolution theorem, when we transform Qi(Q) M(Q) we obtain
D(r) FT[M(Q)], where FT[M(Q)] is the Fourier transform of M(Q). We
have previously identified that the sinc function is the Fourier transform
of the box function, and hence the Fourier transform of M(Q) is in turn
a box function. Thus the experimentally-derived D(r) function is broadened by convolution with a box function of full width 2 /Qmax.

139

CHAPTER 7

Magnetic materials

Magnetic ions
A number of atoms or ions have magnetic moments, although by no
means all. There are three origins of magnetism in atoms. First is the
magnetic moment of the nuclear, which arises when the numbers of
protons and neutrons in the nucleus are such that there is not a cancellation of the moments of these particles. We will find, however,
that in terms of the magnetic structures of materials the nuclear moments are not important (although they are extremely important in
experimental methods such as magnetic resonance techniques, and
are are important at very low temperatures).
The second origin of atomic magnetism is from the spin orbital angular moment of the electrons. Clearly in open-shell ions with odd
141

numbers of electrons there will be incomplete cancellation of the spin


moments (but we note that the basis of forming ionic bonds is to transfer electrons in order to form closed shells). However, in transition and
rare earth metal ions, the lowest-energy electronic configurations actually favour adding electrons to the part-filled shells initially all with the
same spin orientation and then adding electrons with opposite spin orientations only after the shell is half filled. Thus there is a significant
magnetic moment arising from the non-cancelling electron spin moments.
The third origin of atomic magnetism is from the orbital angular momentum of electrons in partly-empty shells. For partly filled shells there
will be non-cancellation of the orbital angular momentum of the atom.
In a classical picture, an electronic with orbital angular momentum
will act like an electrical current loop, which will generate a magnetic
field.
This discussion relies on the fact that in the transition and rare earth elements, the electronic energy levels are filled in a way that avoids cancellation of moments.
The d levels have 5 levels from lz from +2 to 2. Electrons first enter

each state (beginning with the lz = + 2 state) one electron by electron,


to give for the first five states:

S = 1/2

L =2

S = 3/2

S=1

L =3

S=2

L =3
L =2

S = 5/2

L =0

l . Subsequent electrons then fill the remain states with


z
pairings of spins:
where L =

S=2

L =2

S = 3/2

L =3

S=1

L =3

S = 1/2

L =2

S=0

L =0

There is a separate rule that determines how the vectors S and L combine to form J.
Similar rules are found for the filling of the 7 partially-empty f levels (lz
from +3 to 3) as found in rare-earth metals. These rules are called
Hunds rules.
There is one difference between transition metal and rare earth metal
ions. In the transition metals, the unfilled shells are in the outer part of
the atom, which means that when the ion is within a crystal these electrons see the fields from the local environment, and the net effect is
that the orbital angular momentum does not operate to contribute to
the overall magnetic moment, which is given by the spin angular moment only. This is not the case for the rare earth ions; in their case the
unfilled shells are below some closed shells and are shielded from the
effects of the crystalline fields.
The size of the magnetic moments can be measured by measuring the
magnetisation of a sample when in an applied external magnetic field.
Broadly it is found that the qualitative description given here is correct,
although there is a small number of exceptions. The response of a crys142

tal containing non-interacting magnetic ions to an applied magnetic


field is discussed in the next section.

143

Response of a material to an
applied magnetic field
Diamagnetism
We consider first a relatively weak effect, which is found in all materials regardless of whether there is a net magnetic moment in the atoms. Without going into detailed analysis, the key point is that in a
classical sense we can think about the orbital angular momentum
corresponding to an electric current loop in the atom. Application
of a magnetic field, by Lenzs law, will generate an opposing current
that will generate a magnetisation in the opposite direction to the
applied field. Thus we can write
M = Bext
144

where Bext is the externally-applied magnetic field, and where for this
effect , which is called the susceptibility (formally = M/B), has a
negative value.

Paramagnetism
Of more interest is the effect of a magnetic field on a crystal containing
non-interacting magnetic ions. Here we discuss the behaviour of the
magnetic susceptibility as a function of temperature, but using plausibility arguments rather than detailed thermodynamic analysis.
At high temperatures, the magnetic ions have a lot of thermal energy,
and a high value of the externally-applied magnetic field is needed to
align the moments. On cooling, the thermal opposition to ordering is
lessened, until at zero temperature it is very easy to order the magnetic
moments. Thus we expect the paramagnetic susceptibility is expected
to have the form
= C/T
This temperature dependence, known as Curies law, is found for many
paramagnetic systems. It can be derived quite easily from thermodynamics, although like many laws it is obtained as a limiting behaviour
(here for the case of a weak field). The constant C can be calculated
from the magnetic properties of the ion, and in particular it is directly
determined by the size of the magnetic moment. From quantitative
measurements of the susceptibility is it therefore possible to obtain values for the magnetic moments, which can be compared with the analysis given in the previous section. It is generally found that there is excellent agreement, albeit with a few exceptions.

145

Magnetic interactions
Phase transitions
In some cases the magnetic moments interact, which leads to a
spontaneous ordering of the atomic moments. Once the moments
order, they create a magnetisation. It is possible (but for the moment
dont take this literally) that this magnetisation, M, then acts with the
external magnetic field to create a new field
Beff = Bext + M
When we apply this to the Curie law, we obtain

146

M=

C(Bext + M )
T

We can rearrange this as


M(T C) = CBext M =

C
Bext
T C

It thus follows that the susceptibility now has a new form, known as
the Curie-Weiss law:
=

C
T Tc

where Tc = C is the temperature below which there is a spontaneous


magnetisation. It is called either the critical temperature or the Curie
temperature; the latter having an obvious historical link but without reflecting the universal behaviour that is now recognised to be characteristic of many different types of phase transition. The important point is
that the susceptibility diverges at T = Tc, at which point the material is
very easily ordered by the external magnetic field. We saw a similar feature when we met ferroelectricity earlier.
Actually the Curie-Weiss form of the susceptibility is only partly correct
for magnetic materials, but that is another story.
The key question to ask is what is the field M; is it simply the magnetisation? It actually turns out that the magnetisation is too weak to produce ordering directly via the magnetic field. At best the ordering temperature produced directly via the magnetic field would be of the order
of 23 K, as compared to cases that can be two orders of magnitude
higher. So we need to search for other ways for magnetic moments to
interact.

FIGURE 7.1 Example of the super-exchange interaction between


two d-electron systems interaction with the p electrons on an
intermediate oxygen cation.

Exchange interactions
We now understand that the interactions between two magnetic ions
occurs as a result of their overlapping wave functions. There is a coupling between the spin angular momentum of electrons in two atoms
of the form
E = J Si Sj
which is purely a result of quantum mechanics. This interaction is
much stronger than the result of direct magnetisation.
Direct exchange occurs when two magnetic ions are touching such
that their electrons overlap. However, this is very rare.
Super-exchange is a mechanism in which two magnetic ions can interact via an intermediate ion. It is illustrated in Figure 7.1 for the case of
two d-electron ions (such as Mn2+) interacting via an intermediate O2
anion. In this case the moment on one cation interactions with the electrons in one lobe of electrons in the oxygen in a way that favours antiparallel alignment. This immediately polarises the electron on the opposite lobe of the oxygen atom, which causes the second cation to have
147

TABLE 7.1 Transition temperatures (K) for some selected


ferromagnetic materials.
Fe

Co

Ni

Gd

EuO

EuS

1043

1394

631

289

70

16.5

its magnetic moment pointing in the opposite direction to that of the


first cation. This mechanism, or variants of it, is much more common
than direct exchange.
In metals, the exchange interactions are via the conduction electrons.
This can lead to quite long range interactions with oscillating signs; in
turn this can lead to quite complicated ordered structures.

Ordered phases
The scenario described above is for the case when the magnetic order
leads to a macroscopic magnetic polarisation. For materials containing
only one type of magnetic ion, this will typically involve all the magnetic moments being aligned together. This phenomenon is called ferromagnetism. A ferromagnetic phase transition will occur at a fixed temperature, and below the transition temperature the magnetisation will
increase gradually up to a saturation value namely when all the moments are perfectly aligned at low temperature. Some examples are
given in Table 7.1.
In a ferromagnet, the exchange interactions favour parallel arrangement of the moments, but in fact this is not the common case. It is
more usual for exchange interactions to favour alignment in opposite
directions. In the magnetic crystal, this type of moment orientation also
occurs at a particular transition temperature, but there is no bulk magnetisation. Instead, the existence of the magnetic order can be seen in
diffraction experiments.

INTERACTIVE 7.1 Magnetic structure of MnO (NaCl structure),


showing (111) sheets of ferromagnetic ordered moments with
alternate layers being aligned in an antiferromagnetic
configuration.
Consider the case of MnO show in Interactive 7.1. The magnetic structure consists of (111) planes of moments ordered in a ferromagnetic arrangement, but with neighbouring planes having an antiferromagnetic
alignment of the moments. The dominant super-exchange interactions
are along the [100] directions, which give the antiferromagnetic order.
The 90 super-exchange interactions between nearest neighbours
within the (111) planes are ferromagnetic.
One thing that is apparent from the antiferromagnetic structure shown
in Interactive 7.1 is that the ordering is such that two atoms related by
148

TABLE 7.2 Transition temperatures (K) for some selected


antiferromagnetic phase transitions.

TABLE 7.3 Transition temperatures (K) for some selected


ferrimagnetic phase transitions.

MnF2

MnO

CoO

FeO

Cr2O3

Fe2O3

Fe3O4

CoFe2O4

NiFe2O4

67

116

292

116

307

950

858

793

858

a simple lattice translation are no long equivalent. This means that the
unit cell is now a multiple of the unit cell described by the atoms. In
the case of MnO it is clear that the unit cell doubles along the a, b and
c axes (although in fact it is possible to define a smaller primitive unit
cell). This means that the (100) Bragg peak, which is absent in a facecentred cubic diffraction pattern, now corresponds to the (200) reflection of the new unit cell, and hence is visible in a neutron diffraction
measurements (the neutron has its own magnetic moment, and thus is
able to detect magnetic ordering in a material).
A number of examples are given in Table 7.2.
The magnetic susceptibility for an antiferromagnet needs to be described in terms of the two magnetic sublattices, each of which will
generate an equivalent of the M term that we used to describe the susceptibility of a ferromagnet. But now the interaction is negative, and
were we to follow a similar analysis that derived the susceptibility of
the ferromagnet but carried by by sublattice, we would find that the susceptibility varies as
=

C
T + T

Y3Fe5O12 Gd3Fe5O12 Ho3Fe5O12


560

564

567

A third state of ordering occurs for cases when there is more than one
type of magnetic ion; the mineral magnetite, formula Fe3O4, containing
both Fe2+ and Fe3+ cations is an example. These materials are called ferrimagnetic. In many cases, and Fe3O4 is one, the magnetic moments of
the two ions do not cancel and there is a net magnetisation in the ordered state.
Several examples of ferrimagnetic materials are given in Table 7.3.
The fourth state we will discuss is where the magnetic moments order
in a spiral structure, a property called spiral magnetism or helical magnetism. We consider a simple energy of the form
E = NS 2(J1 cos J2 cos 2) = NS 2 (J1 cos J2(2 cos2 1))
where is the angle between neighbouring moments, J1 is the exchange interaction for nearest neighbours, and J2 is the exchange interaction for second-nearest neighbours (hence the 2 argument in the cosine). This equation holds for interactions in a one-dimensional system,
or between ferromagnetic planes. The equilibrium structure is that for
which E / = 0:
(J1 4J2 cos )sin = 0

where T is called the Weiss temperature. The magnetic susceptibility is


rising on cooling, but will diverge at a negative value of temperature.

The solutions to this are = 0 or = , which correspond to the cases


of ferromagnetism or antiferromagnetism, and

We make one remark about the direction of the magnetic moments in


MnO. The orientations of the moments do lie within the planes, but

cos =

J1
4J2
149

ues of J2), and for all negative values of J2 either of these states is favoured over the spiral state. However, when J2 is positive, the secondneighbour interactions oppose the same alignment of second-nearest
neighbours, and if they are strong enough they make the system prefer
the spiral ordering. We take the case of positive J1 (the argument is exactly the same for negative J1 but in that case we would compare with
the antiferromagetic energy). The spiral state is preferred if the energy if
the spiral phase is lower than that of the ferromagnetic state:
J12
Eferro > Espiral NS (J1 J2) > NS
+ J2
( 8J2
)
2

This reduces to
J12
J12
(J1 J2) >
J2 8J1 + 16J2 >
8J2
J2
INTERACTIVE 7.2 Example of a spiral magnetic structure in a simple
cubic structure.
The energies for the different cases are
= 0 Eferro = NS 2(J1 J2)
= Eantiferro = + NS 2(J1 + J2)
J1
J12
2
cos =
Espiral = NS
+ J2
4J2
( 8J2
)
The ferromagnetic case is favoured over the antiferromagnetic case
when J1 > 0 for all negative values of J2 (regardless of whether we have
ferromagnetic or antiferromagnetic ordering, the second neighbours always have the same orientations, which are favoured by negative val-

Writing J1 = a J2 we have the condition


8a + 16 > a 2 (a 4)2 > 0
This is of course always true, but we have also to be consistent with the
condition that | cos | > 1, so the spiral state is stable under the condition 4J2 > J1, where we are in the regime that both J1 and J2 have positive values.
Thus to obtain spiral magnetism we need to have a sufficiently large
value of J2. This is most commonly possible in metallic crystals of the
rare earth elements, where the exchange interactions are reasonably
long range and of either sign, owing to the interaction between the
magnetic moments and the conduction electrons.

150

Bulk magnetisation
Our final point concerns the macroscopic measurements of magnetisation. In principle it would seem that all you might need to do is
put a sample into an instrument that merely measures magnetism
and you would get an answer. However, it is not as simple as this,
mostly because different parts of the sample may form a net magnetic moment independent from other parts of the sample, with the
result that the overall sample magnetisation remains close to zero.
Figure 7.2 shows and example of how a material can form small regions called domains in which the magnetisation points along
different directions. The shapes of these domains will be determined
by a number of factors associated with the interactions between
151

M,P
P= E
M= B

M0,P0

Bc,Ec

FIGURE 7.2 Magnetic domains in a metal observed by a magnetic


optical effect called the Kerr effect.
magnetic moments around the boundaries called the domain walls
and the way that the domain walls interact with strain. The same effect
is also observed in ferroelectric systems that we met earlier, but there is
one difference. In ferroelectrics, the domain walls are likely to be quite
small, on the scale of the unit cell, but in ferromagnetism the energy of
the domain wall can be reduced by forming a spiral with a shallow angle, which will meant that magnetic domain walls can be quite wide
(many unit cells).
You can imagine that solution to the problem of forming domains is to
align the moments into one single domain by applying a large mag-

Bc,Ec

B,E

M0,P0

FIGURE 7.3 Hysteresis loop for both ferromagnetism and


ferroelectricity.
netic field. Indeed, this is exactly what is done, but what is interesting
is the shape of the magnetic response as the applied magnetic field is
first applied, then reversed, and cycled back and forth, The behaviour
is illustrated in Figure 7.3, and the characteristic shape (which is also
seen for the analogous experiment on a ferroelectric material) is called
a hysteresis loop.

152

Hysteresis loops are relatively easy to understand. At high field the magnetisation is the equilibrium value for the temperature plus a contribution caused by the applied field, namely M = M0 + B. When the direction of B is reversed, the magnetisation doesnt flip but instead is reduced slightly from the value of M0 until a certain field strength, called
the coercive field strength and given the symbol Bc, is reached. Only
then is the magnetisation reversed.
Clearly it costs energy to reverse the magnetisation, and to understand
this we need to have a picture of what is happening at the atomic level.
It costs a lot of energy to flip the direction of a single atomic moment
because of its interaction with its neighbours, so the easiest place for
the moments to flip is at the domain wall. What you might expect,
therefore, is for the applied field to move the domain walls. Unfortunately they are pinned by defects within the crystal; Figure 7.4 shows
an example of the sort of defect that will trap or pin a magnetic domain wall.

FIGURE 7.4 Example of a line defect seen in an electric microscope.


The white blobs are not atoms but do represent an arrangement of
atoms.

153

Appendices

A. Point symmetry and lattice types

All three-dimensional crystal structures have lattices that are one of 7


crystal systems defined by symmetry, but there are 14 lattice types,
known as the set of Bravais Lattices, that Lacan be mapped onto these
crystal systems.

Point symmetry operations


A lattice point will see the symmetry of the arrangement of lattice
points around it. There are three types of symmetry operations we must
consider.
Mirror symmetry is described in terms of plane in space such that whatever is above and below the plane are mirror images of each other. In
terms of coordinates, if the mirror plane is perpendicular to the z axis
in a coordinate system, the mirror symmetry will transform all points as
x, y, z x, y, z
where we use the nomenclature z = z.
Rotational symmetry defines an axis in the structure such that rotation
by a rational fraction of a complete cycle, that is rotation by an angle

360/n, where n is an integer, brings the structure back onto itself. For
three-dimensional crystals, the only allowed values of n are 1, 2, 3, 4
and 6. The case n = 1 does nothing of course, but is included within
the formalism specifically to denote the absence of other symmetry.
Transformations for different values of n assuming that the rotation axis
is along the z axis are
n = 2 : x, y, z ; x, y, z
n = 3 : x, y, z ; y, x y, z ; x + y, x, z
n = 4 : x, y, z ; y, x, z ; x, y, z ; y, x, z
n = 6 : x, y, z ; y, x y, z ; x + y, x, z ;
x, y, z ; y, x + y, z ; x y, x, z
Rotoinversion symmetry is rather harder to imagine. It is a two stage
process, involving a rotation by 360/n followed by the transformation
x, y, z x, y, z. The rotoinversion symmetry is given the symbol n. Thus
we have the following transformations, where the rotation axis is assumed to be along z:
1 : x, y, z x, y, z
155

2 : x, y, z ; x, y, z
3 : x, y, z ; y, x y, z ; x + y, x, z ;
x, y, z; y, x + y, z ; x y, x, z
4 : x, y, z ; y, x, z ; x, y, z ; y, x, z
6 : x, y, z ; y, x y, z ; x + y, x, z ;
x, y, z ; y, x y, z ; x + y, x, z
In this case, 1 corresponds to what is called a centre of symmetry,
namely that any point has a corresponding image where all coordinates have reversed sign. This is a particularly important symmetry for
the properties of crystals. The case 2 is equivalent to the operation of a
mirror plane.

It turns out that there are 32 point groups consistent with the presence
of a crystal lattice. This is slightly fewer than the maximum number of
permutations because some of the combinations are equivalent; this is
illustrated by the previously-noted fact that the 2 symmetry is equivalent to a mirror symmetry.

Symmetry of lattices
A lattice is a collection of points in a periodic arrangement. By construction, all lattices contain a centre of symmetry.
The seven lattice types are given in the table above. There is scope for
confusion here, because the table actually has eight entries. This is because the rhombohedral type is a special form of the trigonal lattice.

Point groups
These symmetry operations can be combined to form sets, which are
know as the point groups. For example, consider the a simple example
of combining a 2-fold axis with a perpendicular mirror plane. This is
given the symbol 2/m, and there are 4 positions generated by the combination; one as the starting point, one produced by the 2-fold axis,
one generated by the mirror plane, and one generated by the operation
of both the 2-fold rotation and the mirror plane. This set of equivalent
coordinates is
2/m : x, y, z; x, y, z; x, y, z; x, y, z
Note that the operation of both the 2-fold rotation and mirror plane has
generated the centre of symmetry.

156

B. Bravais lattices

The Bravais lattices are 14 lattice formed from the seven lattice types
and now allowing for lattice points to also exist at face centres and

INTERACTIVE 8.1 Primitive triclinic lattice. a b c,

body centres. In this appendix we give representations of each of the


Bravais lattices. Note that in the captions F-centred corresponds to the

INTERACTIVE 8.2 Primitive monoclinic lattice. a b c,


= = 90
157

INTERACTIVE 8.3 C-centred monoclinic lattice. a b c,


= = 90

INTERACTIVE 8.4 Primitive orthorhombic lattice. a b c,


= = = 90

case where there is a lattice point at the centre of each of the faces of
the unit cell, C-centred corresponds to the case where this is a lattice
point in the centre of the face opposite the c lattice vector, and Icentred corresponds to the case where there is a lattice point at the centre of the unit cell. Note that it is also possible to refined A-centred and
B-centred lattices, where there are lattice points at the centres of the
faces opposite the a and b lattice vectors respectively; it only makes
sense to define these for the cases of orthorhombic and monoclinic lattices, where the choice of C-centre is a matter of convention.

ways. The one lattice that has a 3-fold rotation axis is the rhombohedral lattice as shown here. However, it is possible to populate the hexagonal lattice with a motif with only 3-fold rotation axis so as to generate a trigonal crystal, although the underlying lattice is hexagonal. It is
possible to draw a hexagonal lattice within the rhombohedral lattice
with a unit cell that is three times larger than that of the rhombohedral
unit cell.

One interesting example is trigonal. The trigonal symmetry is defined


as having one 3-fold rotation axis. This can be accomplished in two
158

INTERACTIVE 8.5 C-centred orthorhombic lattice. a b c,


= = = 90

INTERACTIVE 8.6 I-centred orthorhombic lattice. a b c,


= = = 90

159

INTERACTIVE 8.7 F-centred orthorhombic lattice. a b c,


= = = 90

INTERACTIVE 8.8 Rhomohedral trigonal lattice. a = b = c,


= = 90

160

INTERACTIVE 8.9 Primitive hexagonal lattice. a = b c,


= = 90; = 120

INTERACTIVE 8.10 Primitive tetragonal lattice. a = b c,


= = = 90

161

INTERACTIVE 8.11 I-centred tetragonal lattice. a = b c,


= = = 90

INTERACTIVE 8.12 Primitive cubic. a = b = c, = = = 90

162

INTERACTIVE 8.13 I-centred cubic. a = b = c, = = = 90

INTERACTIVE 8.14 F-centred cubic. a = b = c, = = = 90

163

C. Calculation of coordination ratios of ionic radii

Here we calculate the geometry for four coordination polyhedra consisting of a central cation surrounded by 4, 6, 8 and 12 anions. The cations and anions are treated as spheres with radii r + and r respectively,
and here we calculate the radius ratios r /r + for the cases where the
cationanion bond distance is equal to r + + r and where the distance
between closest anions is equal to 2r , namely the case where anions
are in contact both with the bonded cation and their neighbouring anions. In practice you can imagine that in

4-coordinated cation
In this case, the anions form a perfect tetrahedral arrangement around
the cations, with the angle between two cationanion bonds being
equal to cos1(1/3) = 109.47. By geometry we find that
2r =

8 +

(r + r )
3

which gives a ratio of


r
=
r+

1
3/2 1

= 4.449

6 coordinated cation
In this case the anions form a perfect octahedral shape, with the angle
between two cationanion bonds being 90. The geometry shows that
2r =

2 (r + + r )

which gives the ratio


r
=
r+

1
21

= 2.414

8 coordinated cation
In this case, the anions form a perfect square around the cation, with
the angle between two cationanion bonds being equal to
cos(1/3) = 70.53. From geometry we have
2r =

2
3

(r + r )

which gives the ratio


164

r
=
r+

1
31

= 1.366

12 coordinated cation
The anions form a perfect icosahedron, with the angle between two
cationanion bonds being 60. Geometry gives
2r + = r + + r
and hence
r
=1
r+
which is only achieved in practice for the case where the anion is the
same as the cation.

165

D. Ionic radii

A number of attempts have been made to develop a consistent set of


ionic radii. The problem is always that ions come in pairs because
you only observe distance between ion centres and never isolated ions
and separating out one radius is therefore difficult. Nevertheless, it
has proven possible to obtain a consistent set of ionic radii, which are
reported here. These values have been taken from one extremely useful
web resources called WebElements, which has taken its values from
work by Shannon and Prewitt (Acta Crystalographica B25, 925, 1969;
B26, 1046, 1970) and developed further by Shannon (Acta Crystalographica A32, 751, 1976).
We present the data separately for anions and cations with different coordination numbers. Our lists are not comprehensive but cover most of
the elements you might encounter in materials; the WebElements web
site will fill in any that we have not included.

Because of the ambiguity with respect to isolating separate ions, it will


Tetrahedrally-coordinated cations
Cation

Radius ()

Cation

Radius ()

Al3+

0.53

Mn4+

0.53

B3+

0.25

Mn5+

0.47

Be2+

0.41

Mn6+

0.395

Cd2+

0.92

Mn7+

0.39

Cu+

0.74

Mo5+

0.60

Cu2+

0.71

Mo6+

0.55

Fe2+

0.77

Na+

1.13

Fe3+

0.63

Ni2+

0.69

Ga3+

0.61

P5+

0.31

Ge4+

0.53

Si4+

0.40

Anion

Radius ()

Anion

Radius ()

In3+

0.76

Sn4+

0.69

Br

1.82

O2

1.25

K+

1.51

Ti4+

0.56

Cl

1.67

S2

1.70

Li+

0.73

W6+

0.56

1.19

Se2

1.84

Mg2+

0.71

Zn2+

0.74

2.06

Mn2+

0.80

Zr4+

0.73
166

Octahedrally-coordinated cations

Cube-coordinated cations

Cation

Radius ()

Cation

Radius ()

Cation

Radius ()

Cation

Radius ()

Ag2+

1.08

Mo5+

0.75

Ba2+

1.56

Na+

1.32

Al3+

0.675

Mo6+

0.73

Bi3+

1.31

Pb2+

1.43

B3+

0.41

Na+

1.16

Ca2+

1.26

Rb+

1.75

Ba2+

1.49

Ni2+

0.83

Cd2+

1.24

Sc3+

1.01

Be2+

0.59

Ni3+

0.70

Cs+

1.88

Sn4+

0.95

Bi3+

1.17

Ni4+

0.62

Fe2+

1.06

Sr2+

1.40

Ca2+

1.14

P5+

0.52

Fe3+

0.92

Ta5+

0.88

Cd2+

1.09

Pb2+

1.33

In3+

1.06

Ti4+

0.88

Cs+

1.81

Rb+

1.66

K+

1.65

Y3+

1.16

Cu+

0.91

Sc3+

0.885

Li+

1.06

Zn2+

1.04

Cu2+

0.87

Si4+

0.54

Mg2+

1.03

Zr4+

0.98

Cu3+

0.68

Sn4+

0.83

Mn2+

1.10

Fe2+

0.75

Sr2+

1.32

Fe3+

0.69

Ta3+

0.86

Ga3+

0.76

Ta4+

0.82

Ge4+

0.67

Ta5+

0.78

In3+

0.94

Ti2+

1.00

K+

1.52

Ti3+

0.81

Li+

0.90

Ti4+

0.745

Mg2+

0.86

W4+

0.80

Mn2+

0.81

W5+

0.76

Mn3+

0.72

W6+

0.74

Mn4+

0.67

Y3+

1.04

Mn7+

0.60

Zn2+

0.88

Mo3+

0.83

Zr4+

0.86

Mo4+

0.79

always be the case that the greatest value of these tables is firstly to calculate expected bond lengths, and secondly to account for differences
in size of two ions. They might be less accurate for tasks such as calculating the radius ratios r /r + that we have talked about in terms of understanding the relative stabilities of different structure types.

167

Radiation beams for materials research

The study of materials at an atomic level requires the use of beams of


radiation with wavelengths at the atomic scale. Wave-particle duality
gives us the prospect of using both electromagnetic radiation and matter waves.

Laboratory x-ray sources and detectors


The part of the electromagnetic spectrum with wavelengths in the 1
range are called x-rays. X-rays with well-defined wavelengths are produced within atoms when a electron orbital that lies close to the nucleus is empty, and an electron in an outer orbital drops down into the
vacant orbital and emits a photon as it does so. Typically the vacant orbital is creating by knocking out an electron, usually by bombarding a
target with a beam of electrons. This is achieved by a relatively easy experimental set up, as illustrated in Figure 6.1. The beam of electrons is
produced within an evacuated vessel from the cathode, and accelerated by electric fields and directed onto a metal target that acts as an
anode.
The x-rays emitted are characteristic of the particular metal used in the
target. Typical spectra for two common target materials used in labora-

Ua

Uh

Wout
C

A
Win

X
FIGURE 8.1 Illustration of a vacuum tube used to create beams of
x-radiation. C represents the cathode that emits the electrons,
and A represents a metal target that acts as the anode. The
beams of x-rays emitted from the target are denoted by X.
tory x-ray sources are shown in Figure 6.2. There are two features of
these spectra. The first is a set of sharp peaks in the spectra, which in
168

4p
4s

3d

N
3p
3s

M
2p
2s

L
K

M
M

1s

2
1

L2
L

FIGURE 8.2 Characteristic x-ray spectra for two typical target


materials.
Figure 6.2 are labelled K and K. These are generated by the beam of
electrons knocking the electrons in the target atoms out of the atom,
leaving a vacancy in the orbital shell. Quickly an outer electron will
fall into this lower-energy vacancy, and it will emit a characteristic xray photon as it does so. There are are course different shells from
which the first electron can be knocked out, and different shells from
which the second electron can come from, leading to a number of possible spectral lines. The relationship between the spectral lines and the
atomic structure is shown in Figure 6.3.
The lines that are usually of most value are the K lines. The other spectral lines can be largely removed by the use of absorption filters. X-rays
are absorbed by their energy transferred to the electrons. Such a process requires that the energy of the x-ray needs to be above a certain

FIGURE 8.3 Origin of characteristic x-ray spectral lines.


threshold energy, and as a result the absorption is very low for x-ray energies below the threshold and high for x-ray energies above. Thus an
absorbing material whose energy threshold is just above the energy of
the K radiation of the x-ray source is ideal as a filter material.
Some x-ray wavelengths of standard laboratory sources are given in Table 6.1.
In addition to the characteristic wavelengths, Figure 6.2 also shows a
broad continuous spectrum, which is the second feature of note. This
arises from the slowing down and deflection of the electrons that are
169

TABLE 8.1 Characteristic wavelengths of x-rays emitted from


standard laboratory sources.
Anode material

K wavelength ()

K wavelength ()

Fe

1.935

1.757

Cu

1.541

1.392

Mo

0.709

0.621

Ag

0.559

0.497

fired onto the target. Any accelerating charge, which includes deceleration and deflection, will emit electromagnetic radiation, and in an xray source this effect is seen as the continuous distribution. This is
called Bremsstrahlung radiation (from the German for slowing
down). The cut-off at low wavelength corresponds to the maximum
energy of the electrons as defined by the accelerating voltage of the xray tube. Again, much of the Bremsstrahlung spectrum can be absorbed by filters.
The principles of creating x-rays within the laboratory have not
changed significantly over many years, although there are specialised
sources that have more intense and more focussed beams. The biggest
changes have come in the methods used to detect the scattered beams
of x-rays. Much early work used photographic file, with instruments designed to optimise the way that the photographic sheets were exposed
to the diffracted beams of x-rays. It became possible to measure the
darkening of the photographic film with a high degree of accuracy.
However, film has largely been replaced by electronic detectors, now
based on semiconductors (where the x-ray photon is converted to an
electron-hole pair) and/or scintillation devices (where the x-ray photon
is converted to an optical photon). The Braggs actually used ionisation
detectors, where the x-ray photon ionises a gas and produces a pulse

of electrical current, and such detectors were used for a long time after,
but the modern detectors have the advantage of having much better
spatial resolution and thus can be assembled in an array that covers a
wide area of the scattering space but enables the specific x-ray photon
to be pinpointed to a specific pixel within the array.

Synchrotron radiation
There is an inherent limit to the intensity of x-ray beams produced in
the laboratory, in spite of the development of specialised tubes designed to increase the intensity. This limitation is most keenly felt in
two regards. First, when measurements need to be fast because many
measurements are required, for example on a large suite of samples or
when wanting to perform measurements at many temperatures, and secondly when sample environment (particularly high-pressure apparatus)
absorbs the beam.
Synchrotron radiation is generated by rapid acceleration of beams of
electrons. In a modern facility, beams of electrons are accelerated
around a rings using bending magnets. Through being directed in a circular orbit the electrons continuously emit radiation. However, more
intense and focussed beams of radiation can be produced through devices inserted in the path of the electron that cause sudden changes in
direction of the beam. These sudden changes are much more efficient
at producing intense beams of x-rays.
Specific wavelengths are selected by what is called a monochromator
crystal, in which the by a particular alignment of crystal planes only xrays of a specific wavelength are diffracted.

170

Beams of neutrons from a nuclear fission


reactor
The production of neutrons in a fission reaction lies at the heart of the
chain reaction that underpins the technology. The probability of a nuclear fission reaction increases for slower neutron speeds, and hence a
typical reactor will operate at around room temperature. Each fission
reaction produces a certain number of neutrons, which are cooled
down through interaction with the atoms in materials inserted specifically for this task and which are called the moderators.
Thus in principle neutron beam production at a nuclear fission reactor
should be straightforward. Indeed, in operation a reactor needs to have
neutrons absorbed in a controlled way in order to avoid a runaway reaction. So there are plenty of neutrons to harvest. However, you nevertheless need to get the neutrons out of the reactor and transported
safely to the instruments; they will not get there on their own in sufficient quantity. For this reason, the design of a research beam reactor
will differ significantly from that of a power reactor.

Beams of neutron from a spallation source


A spallation (the word means to break bits off an object when it is
struck) source of neutrons refers to neutrons knocked from a neutronrich nucleus when hit by an object, most typically a high energy proton. Thus a spallation source consists of a beam of protons incident on
a heavy metal target, with neutrons knocked from the atoms in the target. They are released with high energy and need to be cooled by passing through a moderator at low temperature.

171

E. Nobel prizes for materials physics and chemistry


GALLERY 8.1 Nobel prize winners in the study of crystalline
materials

It is interesting to note the large number of Nobel Prizes that have been
won for work connected with and around the topic of this book,
namely the physics and chemistry of materials.
The range of work is large. It encompasses discovery of underpinning
phenomena (such as x-rays and the diffraction of x-rays), developments
of experimental measurements techniques and analysis (such as x-ray
diffraction, electron microscope, neutron and light scattering methods,
surface scanning methods, nuclear magnetic resonance), experimental
methods (such as cooling to liquid helium temperatures and high pressures), theory and computation (such as theories of superconductivity,
quantum Hall effect, understanding universal phenomena associated
with phase transitions, and development of simulation methods), discoveries of new effects or materials (such as graphene, hightemperature superconductivity, tunnelling effects in superconductors),
and phenomena associated with technological developments (such as
the transistor and the large magnetoresistence).
Here we list, with photographs, all Nobel laureates in the disciplines of
Physics and Chemistry whose work can more-or-less be said to be in
the area of materials physics and chemistry.

Nobel Prize in Physics 1901


Wilhelm Conrad Rntgen
"in recognition of the extraordinary services he has rendered by the
discovery of the remarkable rays subsequently named after him"
1 of 55
172

You might also like