Sie sind auf Seite 1von 193

notes on

Ph a s e Transitions

Martin T. Dove and Anthony E. Phillips

Contents

Preface

Viewing crystal structures

vii

List of Symbols

ix

.
.
.
.
.
.
.
.
.
.
.
.
.
.

1
1
1
4
6
7
9
13
13
16
17
19
20
21
22

Magnetic phase transitions


2.1 A macroscopic view of magnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

25
25

Phenomenology
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2 The role of symmetry and the onset of order . . . . . . . . . . . .
1.3 Switching of the degree of order . . . . . . . . . . . . . . . . . . .
1.4 Example of atomic site ordering . . . . . . . . . . . . . . . . . . .
1.5 Ferroelectric phase transitions . . . . . . . . . . . . . . . . . . . . .
1.6 How to observe a phase transition . . . . . . . . . . . . . . . . . .
1.7 Order of a phase transition . . . . . . . . . . . . . . . . . . . . . .
1.8 General aspects of the thermodynamics of a phase transition . .
1.9 Seeds of a theoretical model . . . . . . . . . . . . . . . . . . . . . .
1.10 Various types of phase transitions . . . . . . . . . . . . . . . . . .
1.11 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.12 Appendix on the statistical mechanics approach to susceptibility
1.13 Short questions to test your basic knowledge . . . . . . . . . . . .
1.14 Longer questions . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

Contents

ii

.
.
.
.
.
.
.
.

26
27
29
31
33
34
35
36

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

39
39
39
41
44
44
47
48
50
52
53
53
62
64
67
71
72
74
78
79

Symmetry
4.1 Introduction to symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

81
81

2.2
2.3
2.4
2.5
2.6
2.7
2.8
2.9
3

A microscopic view of magnetism . . . . . . . . . . . . . .


Non-interacting atoms in a magnetic field: paramagnetism
Interacting atoms in a magnetic field: ferromagnetism . .
Critical exponents revisited . . . . . . . . . . . . . . . . . .
Successes and failures of the mean-field model . . . . . . .
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Short questions to test your basic knowledge . . . . . . . .
Longer questions . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

Landau theory
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2 Quantification of the free energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3 Results for second-order phase transitions . . . . . . . . . . . . . . . . . . . . . . .
3.4 Field-dependence of the order parameter at the transition temperature . . . . . . .
3.5 Taking account of spatial variations . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.6 Validity of Landau theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.7 Ferromagnetism, the mean-field approximation, and Landau theory . . . . . . . .
3.8 First-order phase transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.9 The case when the free energy is allowed to have odd-order terms . . . . . . . . .
3.10 Tricritical phase transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.11 Relevant aside: Definition of the strain and elastic tensors . . . . . . . . . . . . . .
3.12 Phase transitions and elastic strain: case of quadratic coupling . . . . . . . . . . .
3.13 Phase transitions and elastic strain: case of linear coupling . . . . . . . . . . . . . .
3.14 Practical example: Landau theory for the ferroelectric phase transition in PbTiO3
3.15 Comments on Landau theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.16 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.17 Appendix: Correlation functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.18 Short questions to test your basic knowledge . . . . . . . . . . . . . . . . . . . . . .
3.19 Longer questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

Contents
4.2
4.3
4.4
4.5
4.6
4.7
4.8
4.9
4.10
5

iii
Point group symmetry operations . . . . . . . . . . . .
Space group symmetry operations . . . . . . . . . . . .
Groups and their representations . . . . . . . . . . . . .
Symmetry of the order parameter . . . . . . . . . . . .
Symmetry of the spontaneous strain . . . . . . . . . . .
Group-subgroup relationships across phase transitions
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . .
Short questions to test your knowledge . . . . . . . . .
Longer questions . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

Displacive phase transitions and soft modes


5.1 Displacive phase transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2 Phenomenology of the soft mode model of displacive phase transitions . . . . . . .
5.3 Lattice dynamics theory of the soft mode . . . . . . . . . . . . . . . . . . . . . . . . .
5.4 Lattice dynamical theory of the low-temperature phase . . . . . . . . . . . . . . . . .
5.5 Phase transitions, soft modes, and structure flexibility: the Rigid Unit Mode model
5.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.7 Short questions to test your basic knowledge . . . . . . . . . . . . . . . . . . . . . . .
5.8 Longer questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Orderdisorder phase transitions
6.1 Orderdisorder phenomenology . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.2 Mean-field theory of orderdisorder phase transitions: the BraggWilliams model
6.3 Computational methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.4 Beyond BraggWilliams theory: the Cluster Variation Method . . . . . . . . . . . .
6.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.6 Short questions to test your basic knowledge . . . . . . . . . . . . . . . . . . . . . .
6.7 Longer questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

82
84
87
91
93
95
96
98
98

.
.
.
.
.
.
.
.

101
101
107
112
116
120
125
126
127

.
.
.
.
.
.
.

129
129
130
136
143
147
148
148

Critical phenomena
151
7.1 Recapitulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
7.2 Correlation functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

Contents

iv
7.3
7.4
7.5
7.6
7.7
7.8
7.9
8

The Widom scaling hypothesis: relationships between critical exponents


Introduction to the renormalisation group . . . . . . . . . . . . . . . . .
Deriving the Widom scaling hypothesis . . . . . . . . . . . . . . . . . . .
A sketched example: the 1D Ising model . . . . . . . . . . . . . . . . . .
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Short questions to test your basic knowledge . . . . . . . . . . . . . . . .
Longer questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Reconstructive phase transitions


8.1 Introduction and definition . . . . . . . . . . . . . . .
8.2 Examples . . . . . . . . . . . . . . . . . . . . . . . . . .
8.3 Thermodynamics of reconstructive phase transitions
8.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . .
8.5 Short questions to test your basic knowledge . . . . .

Further reading

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.

155
160
165
166
168
168
169

.
.
.
.
.

171
171
171
175
177
178
179

P reface

This document was written for the Phase Transitions module for fourth-year Physics students at Queen
Mary, University of London and within the University of London. Although phase transitions might
seem at first to be a somewhat specific area of condensed matter physics to study, in fact they are
both intimately linked to some of the most profound theories that describe our universe, and on the
other hand of immense practical and technological value in device and materials physics. Our aim in
this module is to bridge the gap between the theory of critical phenomena, arising from the worlds of
statistical physics and quantum field theory, and the real-world applications of these theories to the
many materials that exhibit phase transitions.
These notes were originally written over the course of the first iteration of the module in 2013 and
were been substantially revised in 2014 with minor changes in 2015. Nonetheless typos and perhaps
more serious errors are sure to remain. We will be grateful if you let us know about any you spot, or
indeed about any suggestions you might have for improving the notes or module.
Martin Dove
martin.dove@qmul.ac.uk
Anthony Phillips
a.e.phillips@qmul.ac.uk
September 2015

Viewing crystal structures

In this module you will meet many crystal structures. We want to encourage you to become familiar
with some of these. The best way to do so is to play with the structures themselves. Practically this
means manipulating the structures with software, and we recommend that you install the CrystalViewer
software and play with some of the structure files we will make available on the course web site.
You can download the software from http://www.crystalmaker.com/crystalviewer/. Versions are
available for both Windows and Mac OS. The program should be reasonably self-explanatory, but will
be demonstrated in the first lecture. The basic screen is shown in Figure 1.
The main window is the actual crystal structure, which can be rotated or zoomed. You can change
the viewing direction, and you can also measure distances. There is a top-bar menu together with a
set of icons for the more common commands. You are able to hide atoms using the window in the top
right hand side.
With each structure we will embed some notes, as seen in the bottom right hand side.

vii

viii

Viewing crystal structures

Figure 1: Screen shot of the CrystalViewer program. The user menu is along the top together with graphical icon
short cut to the most common tools. The bottom right hand corner contains some notes that we will provide for
each structure.

List of S ymbols

Unless otherwise specified, an italic version of the symbol for a vector represents its magnitude; thus
for instance B = |B|. A notable exception to this is in the case of angular momentum variables, where
p
the italic version represents the corresponding quantum number: for instance, |J| = h J ( J + 1).
Numbers refer to the equation where the relevant variable is first defined.

Greek symbols

eij

ij

Critical exponent representing variation of the heat capacity with temperature (3.11)
Critical exponent for the variation of the order parameter with temperature (1.2)
Critical exponent for the variation of the susceptibility with temperature (3.17)
Critical exponent for the variation of the order parameter with the corresponding external field
(2.29)
Component of the strain tensor (2nd rank); may be written in Voigt notation
Order parameter (1.3)
Critical exponent related to the pair correlation function at the critical temperature (3.34)
Magnetic dipole moment
Bohr magneton
Critical exponent for the variation of the correlation length with temperature (3.32)
Correlation length (3.30)
Component of the mechanical stress tensor (2nd rank); may be written in Voigt notation
Site occupancy variable
Field that changes the order parameter

ix

List of Symbols

Susceptibility

Roman symbols
a, b, c
a , b , c
B
cijkl
C
dijk
E
Ec
E
f
F
F (hk `)
gJ
h
H
H
h, k, `
J
J
J
l
k

Vectors that define the edges of the crystallographic unit cell and hence form the basis vectors for
the crystal lattice
Reciprocal lattice basis vectors
Magnetic field (flux density)
Component of the fourth-rank elastic constant tensor, can be given also using the Voigt notation
as cij
Heat capacity: can be defined at constant volume (CV ) or constant pressure (CP )
Coefficient of the third-order Piezoelectric constant tensor, which may be given in Voigt notation
instead
Energy, including the internal energy
Critical field
Electric field
Atomic scattering factor for x-rays, often used more generally for other radiation
Free energy
Crystallographic structure factor
Lande g-factor
Plancks constant divided by 2
Enthalpy
magnetic field (field strength)
Miller indices for points of the reciprocal lattice
Total many-electron angular momentum quantum number
Exchange constant in energy functions for coupled spins or site occupancy variables
total angular momentum
Electron orbital angular momentum quantum number
Wave vector; can also use q for phonons

xi
L
L
M
me
ml
ML
ms
MS
N
P
P
q
r
s
S
S
S
T
Tc
TN
u
V
x, y, z
x , y,
z

Many-electron orbital angular momentum quantum number


many-electron orbital angular momentum
Magnetisation
Electron mass
Secondary electron angular momentum quantum number
Secondary many-electron angular momentum quantum number
Secondary electron spin quantum number
Secondary many-electron spin quantum number
Typically used for number of atoms or number of unit cells; be careful which
Pressure
Dielectric polarisation
Phonon wave vector; can also use k
Position vector, denoting for example the position of an atom with respect to some origin
Electron spin quantum number
Many-electron spin quantum number
Entropy
Spin angular momentum
Temperature
Critical temperature (sometimes called the Curie temperature)
Neel temperature
Atomic displacement vector
Volume; be careful regarding what object the volume refers to, whether it be the crystal or a unit
cell or indeed something else
Fractional coordinates of atoms in the unit cell
Unit vectors

Chapter 1

P henomenology

1.1 Introduction
This module will focus on phase transitions. Not all phase transitions, but a fairly wide group of
interesting phase transitions, including many in which the two phases have a symmetric relationship.
Examples we will meet in these lectures include magnetic, displacive, order-disorder and reconstructive
phase transitions, some of which will be introduced in this introduction. We will not make any
assumptions about what you might already know in detail, but we are confident that you will be
familiar with some of these phase transitions from everyday life. The obvious phase transition is melting,
whether you know about ice and water, or the melting of silicate rocks in the inner earth and the
solidification of volcanic magma. You will know about the transition from liquids to gases, or perhaps
about more exotic examples like the transitions from paramagnets to ferromagnets or from miscible to
immiscible in a mixture of two liquids.
The point of the first session of this module is to look at basic phenomenology and the ways in
which we study phase transitions. We will introduce some specific phase transitions so that we can
look at the general points in their context, but we will return to look at these in more detail later in
the module. Through looking at a range of examples at this point we will come to a natural definition
of what a phase transition is, although we caution you to be very wary of forming too premature a
definition, because, as in much of life, it is always possible to find counterexamples.

1.2 The role of symmetry and the onset of order


Many important phase transitions involve a change in symmetry at the phase transition. Whereas
the actual changes in structure at an atomic level may be slight (even infinitesimally slight exactly
1

Chapter 1. Phenomenology

at the phase transition point), nevertheless symmetry is a binary property in that a specific element
of symmetry either exists or doesnt exist. Lets start with a trivial but visual example, namely the
West front of Ely Cathedral, Figure 1.1. It looks as if it originally had a tower either side of the main
central tower, because like in many medieval (and modern) buildings the architect was very keen on the
vertical plane of symmetry in which the right and left sides of the building are mirror images of each
other. Crystallographers, who have provided much of the basic structural information that provides
the foundation for our knowledge of phase transitions, call this mirror symmetry. But what you can see
is that at some point (actually in the 15th century) one of the side towers fell down, and the mirror
symmetry was destroyed; this was like a phase transition.
So lets look at some simple examples of phase transitions in real physical systems, and to start
with we will think about magnetic phase transitions. Figure 1.2 shows a representation of the hightemperature disordered phase of a one-dimensional magnetic material.

Figure 1.1: The West front of Ely Cathedral, showing an extant tower on the South side and the space
where the Northern tower, which collapsed in the 15th
century, once stood. The two side towers would have
created a vertical line of reflection bisecting the central
tower. Ironically, in the world of phase transitions, this
image would be said to possess broken symmetry.

Figure 1.2: One-dimensional representation of the paramagnetic state of a material in which each magnetic moment
can point in either up or down directions. In this image, the orientation of each moment is random, giving a total
magnetic moment, obtained as the sum over all individual moments, of zero.

In general the magnetic moments can be oriented in any direction, but in our our example we restrict
the range of orientations to up and down. In this state, called the paramagnetic phase, the actual direction
of any magnetic moment appears to be random (it isnt, but that is another story for later), and when
we sum all N individual moments j we get zero net moment:
M=

1
N

(1.1)

However, on cooling we can imagine two possible ordering processes. One, shown in Figure 1.3, has
all the magnetic moments pointing in the same direction; this is called the ferromagnetic phase. In the
ferromagnetic state the net moment of the sample M is no longer zero. On heating up to the phase
transition the ferromagnetic state will become progressively disordered whilst still retaining a non-zero

1.2. The role of symmetry and the onset of order

Figure 1.3: One-dimensional representation of the ferromagnetic ordered state of a material, in which each magnetic
moment points in the same direction at low temperature, and preferentially aligns in the same direction at all
temperatures up to the phase transition.

value of M. In many cases, M decreases gradually up to the transition temperature Tc , often with
limiting behaviour that follows a power law:1

|M( T Tc )| | Tc T |

(1.2)

The change in symmetry occurring through the phase transition can be identified by noting that in
the ferromagnetic state the up direction is no longer equivalent to the down direction, whereas these
are symmetrically equivalent directions in the paramagnetic state.

Figure 1.4: One-dimensional representation of the antiferromagnetic ordered state of a material in which each
magnetic moment can point in either up or down directions.

A second ordered state in which neighbouring moments point in opposite directions is shown in
Figure 1.4. In this state the the net magnetic moment M will be equal to zero in both ordered and
disordered states, but we can define a new quantity to describe the order if we take account of the
position of the moment when performing the sum. Formally we do this by assigning a wave vector
qc = 12 a (here |a | = 2/a, where a is the distance between first-neighbour atoms and is usually called
the unit cell length), and writing
1
=
j exp(iqc r j )
(1.3)
N
j

symbol here has a precise meaning that we will revisit


later; for now you can think of it as meaning proportional
to.

1 The

Chapter 1. Phenomenology


where r j = ja. Again, the value of will decrease gradually up to the phase transition and follow a
power-law behaviour similar to that given in Equation 1.2. In fact, might be taken as a general variable
that defines the extent to which the system has changed or become ordered. will have a maximum
value at T = 0, and will decrease on heating up to the transition temperature Tc ; we will find cases
where changes continuously to a zero value on heating to Tc , and other cases where the change is
discontinuous. is given the name order parameter.
It can be claimed that the change in symmetry does not differentiate between the up and down
directions, but nevertheless there is a new symmetry. Essentially the antiferromagnetic state now has a
unit cell that is twice as large, what we call a change in translational symmetry, as defined by the wave
vector qc .2
A three-dimensional representation of an antiferromagnet is shown in Figure 1.5. This shows a
set of magnetic atoms on a simple cubic lattice, with each spin aligned along the z axis, and with all
nearest-neighbour moments.
We have in this small example seen a number of important concepts that will be central to our study
of phase transitions:

2 In

the ferromagnetic state, the corresponding quantity will


have value qc = 0

1. The fact that symmetry changes at the phase transition;


2. That symmetry can include directional and translational elements;
3. That symmetry may be characterised by a wave vector that defines spatial aspects of the ordering;
4. Whilst symmetry changes absolutely at the phase transition, the actual atomic changes can occur
continuously with temperature;
5. That the degree of order can be quantified by the order parameter, which we will frequently denote
with the symbol .
6. That the phase transition occurs at a fixed temperature, which is called the critical temperature and
given the symbol Tc .
Figure 1.5: Representation of a three-dimensional antiferromagnet on a simple cubic lattice.

1.3

Switching of the degree of order


It is well known that application of an external magnetic field can switch the direction of the magnetisation of a ferromagnet. In terms of Figure 1.3, the external field in the downwards direction will cause

1.3. Switching of the degree of order

the atomic magnetic moments to rotate and point along the direction of the field. The ability to switch
the spatial ordering with an external field is one characteristic of a phase transition, but it is the manner
in which the spatial ordering is switched that is interesting.
When an external magnetic field is applied to a ferromagnet in the opposite direction to the ordered
atomic magnetic moments, it acts on each moment individually to give a driving force to rotate the
moment. However, each atomic moment is also interacting with its neighbouring atoms, which act
to maintain the existing degree of order. There is a significant energy cost involved in switching one
individual moment due to the interactions with the neighbouring moments, which want all to point in
the same direction. Therefore for the external magnetic field to switch the direction of alignment, all the
moments need to move almost in unison.
Experimentally, the response of the magnetisation of a ferromagnet to a sinusoidally varying applied
magnetic field has a characteristic form that is represented in Figure 1.6. This is known as a hysteresis
curve. To understand this shape, we consider the start point of the state of order at a finite temperature,
noting that thermal fluctuations will reduce the size of the overall sample magnetisation from its
fully-ordered value. Application of an external field will reduce the extent of these fluctuations and
enhance the overall magnetisation, an effect which will increase the larger the external field. This
explains the linear sloping part at the top of the loop. After increasing the value of the external field,
it is then reduced down to zero, whereupon the magnetisation returns to its equilibrium value. Then
the field direction is reversed and increased in value. The magnetisation is not switched immediately,
but the external field works with the fluctuations to reduce the size of the magnetisation gradually;
the cost of a particular moment fluctuating away from its fully aligned position is reduced because the
field now acts in favour of this. Only when the size of the external field has passed beyond a particular
value called the coercive field strength does the whole magnetisation switch direction to be aligned
with the direction of the magnetic field. And increasing further the size of the field increases the degree
of alignment of the atomic moments in exactly the same way as the start of the process. Then the
size of the external field is reduced towards zero and the overall sample magnetisation reduces to its
equilibrium zero-field value, except now in the negative sense. The external field is then raised in the
positive sense, again reducing the size of the magnetisation until, once the positive value of the coercive
field strength is reached, the magnetisation is switched back to the original direction of alignment. The
process remains reversible time after time.
The value of the coercive field is actually dependent on the quality of the specific sample, not just

M
M = B

M0

Bc

Bc

M0

Figure 1.6: Hysteresis loop.

Chapter 1. Phenomenology

the chemical composition of the material. For example, materials that are poorly crystalline or have
a high number of defects will have a different shape of the hysteresis loop and a different value of
the coercive field strength. This is not surprising; it reflects the fact that for magnetic moments to flip
direction, they need to flip with their neighbours rather than individually. The presence of defects and
poor crystallinity can create locally locked regions of magnetisation which act against switching.
When whole regions flip, at some point these regions will have an interface with regions of opposite
magnetic alignment. And in a high-symmetry crystalline material, where there are equivalent directions
at 90 or 60 apart say, there can be a much larger number of symmetrically equivalent alignment
directions. These regions are called domains, and the interfaces between domains of different alignment
are called domain walls. Domain walls in perfect crystals are fairly mobile, and can be moved (with a
finite velocity) by the applied external magnetic field. On the other hand, domain walls can be pinned
by defects and imperfect crystalline order. It is the pinning of the domain walls that can control the size
of the coercive field.
An illustration of magnetic domains and domain walls wall is shown in Figure 1.7.

1.4
Figure 1.7: Two-dimensional representation of domains in a ferromagnet, shown with different coloured
background. Three types of domain walls are illustrated.

Example of atomic site ordering


In materials with atoms of similar size and electronegativity, it is possible for atoms of different type to
be disordered over equivalent crystal sites. Examples include metallic alloys and some oxides. This
situation is illustrated in Figure 1.8. In this example, the material has an equal number of two atomic
species, and in the disordered phase each site on average is occupied 50% by one species and 50% by
the other.
Frequently site-disordered materials undergo phase transitions in which the atomic species order on
specific sites. This is also illustrated in Figure 1.8, where it can be seen that each atomic species orders
with nearest neighbours of the opposite species, forming a pattern like a chessboard.

3 We

have two subscripts to represent the two dimensions of


space.

The phase transition represented here is not so dissimilar to the antiferromagnetic phase transition.
We can associate a variable with each site, jk ,3 with values 1 depending on whether the site is
occupied by one atomic species or the other. In this case, the order parameter can be written as
=

jk exp(iqc rjk )
j,k

(1.4)

1.5. Ferroelectric phase transitions

Figure 1.8: Left: two-dimensional representation of a metal alloy with equal number of two different atom systems,
with disorder of the atoms in each site in the crystal. Each site can be said to have an average occupancy of 50% of
each of the two types of atoms. Right: ordered phase of the same system.

where r jk = ja + kb, a and b are the two-dimensional lattice vectors, and the critical wave vector for
ordering is qc = 12 a + 12 b . Clearly the variable jk is analogous to the magnetic moment in the
antiferromagnetically ordering system. We will later learn that this similarity means that we can treat
these distinct physical systems within the same theoretical treatment, obtaining very similar results.4
It is found that many metallic alloys undergo atomic site ordering phase transitions. As for the
magnetic phase transitions, these occur at a fixed transition temperature, and often the order parameter
is found to decrease continuously up to the transition temperature.

1.5 Ferroelectric phase transitions


Commercially, the most important phase transition is the dielectric analogue of the ferromagnetic phase
transition, known as the ferroelectric phase transition. The analogy is such that the phase transition is
seen as a spontaneous dielectric polarisation that can be switched by the application of an applied

4 The

only difference is that the sum of the variables has a fixed


value, equal to zero for systems containing equal numbers
of the two atomic species, whereas the sum is allowed to
fluctuate in value in an antiferromagnet. Moreover, there
are classes of magnetic system that are not subject to the
constraint that the magnetic moment is constrained to point
in only two directions. However, the extent to which these
points of detail matter or not is one of the things that makes
the study of phase transitions so interesting in the opinion of
the module lecturers.

Chapter 1. Phenomenology

electric field, and indeed shows the same type of hysteresis loop as seen in ferromagnetic materials. It is
because the analogy with ferromagnetism is so clear that the name ferroelectricity was originally coined.

5 The

centre of symmetry is a point in a crystal in which any


atom at a relative position r is also found at the position r.

There are broadly two types of ferroelectric materials. One involve atoms hopping between two sites
that are equivalent in the high-temperature phase, and then progressively order in one site on cooling
below Tc . A famous example is KH2 PO4 , where the protons hop between two sites along the vector
between two neighbouring oxygen atoms. The crystal structure is illustrated in Figure 1.9, showing the
existence of the two sites for the protons, and also highlighting the existence of PO4 tetrahedral units.
On cooling below 123 K the protons order in one site in a way that collectively breaks the crystalline
centre of symmetry.5 Not surprisingly the transition temperature changes when the hydrogen atoms are
replaced by deuterium atoms; in some cases deuteration changes the transition temperature by roughly
a factor of 2. There are many examples that are associated with the order of hydrogen atoms or the
orientations of OH molecular dipoles.
In some cases, the ordering can involve cations around the centres of local structure polyhedra, such
as Ti4+ cations that lie within TiO6 polyhedra in the perovskite phase of BaTiO3 ; see Figure 1.10. At this
point it is worth noting that the perovskite structure is one of the most important structures we will
meet, and it is essential to understand the main features of this structure.
The perovskite structure can be represented by the general formula ABO3 , where A and B denote
cations whose charge adds up to +6 to counter the 6 charge of the three oxygen anions.6 In the
high-temperature phase, the crystal structure has lattice type primitive cubic with one formula unit
per unit cell. The B cation has 6 oxygen neighbours that give the appearance of an octahedral group
of atoms, and the A cation has 12 oxygen neighbours. Whether you draw the structure with the A
or B cation at the origin makes no difference, and you may find both representations in the scientific
literature; the difference is merely a shift of 12 , 21 , 12 in the coordinates of each atom.

Figure 1.9: Crystal structure of the high-temperature


phase of KH2 PO4 , showing the PO4 tetrahedra as
shaded objects, and the positions of the H atoms represented by two sites along the O. . . O vectors.

6 Note

that there are also halogen-based perovskite structures.

In the case of BaTiO3 , the Ti4+ cations hop between sites that are slightly displaced from the centres
of the TiO6 octahedra along the set of h1, 1, 1i directions; clearly the centres of the octahedra are sites
of maximum potential energy, and the Ti4+ cations are attached to the faces with 3 O2 anions. On
cooling below 393 K, the cations occupy preferentially the top 4 sites (but are still disordered over these
sites), to generate a dielectric polarisation along the z axis, with the crystal symmetry being reduced to
tetragonal. On further cooling the structure preferentially occupies 2 of the sites, and eventually there is
a final phase transition to a state where the Ti4+ cations all occupy one of the sites. The low-temperature
fully-ordered state has its dielectric polarisation pointing along the [1, 1, 1] axis. Clearly this situation

1.6. How to observe a phase transition

looks not unlike the ordering of magnetic moments in a ferromagnetic phase transition.
The second type of ferroelectric involves what is called a displacive phase transition. This type of
transition involves small symmetry-breaking displacements of atoms on cooling into the low-temperature
phase.7 The perovskite PbTiO3 is an example of a displacive ferroelectric phase transition. At high
temperatures (Tc = 763 K) it has the ideal cubic structure, and below Tc there are displacements of the
cations and anions in opposite directions along the [0, 0, 1] direction giving an orthorhombic structure.
There appears to be no significant orderdisorder behaviour associated with this phase transition. The
temperature-dependence of the dielectric polarisation of PbTiO3 is shown in Figure 1.12.

1.6 How to observe a phase transition


Using the examples of phase transitions we have discussed so far, we now consider how one can observe
the phase transitions and perform quantitative measurements.

Direct macroscopic measurements of the order parameter


For a ferromagnetic or ferroelectric phase transition, it is possible to measure the magnetisation
or dielectic polarisation of the sample as a function of temperature. These quantities give a direct
measurement of the order parameter, scaled by the number of moments. Examples of experimental

Figure 1.10: Crystal structure of the high-temperature


phase of the perovskite BaTiO3 , showing 8 sites occupied equally by the Ti atoms at the centre of the TiO6
pseudo-octahedra, with the Ba cations occupying the
corners of the unit cell in 12-fold coordination with
the oxygen atoms.

Polarisation (103 Cm3)

The commercial applications of the ferroelectric phase transitions arise from a number of properties.
As we will discuss later in this module, the dielectric susceptibility, and hence the dielectric constant,
increases significantly at temperatures close to the ferroelectric transition temperature. This gives rise to
the possibility to create ceramic capacitors with much smaller volumes than can be achieved with other
dielectric materials. The fact that the polarisation changes with temperature gives the possibility to create
temperature sensors and infra-red detectors. Many ferroelectric phase transitions are accompanied by
a spontaneous elastic strain, the size of which will be related to the size of the dielectric polarisation.
This gives a direct coupling between electrical and elastic (mechanical) properties, formally called
the piezoelectric effect. Application of stress can generate a change in the dielectric polarisation, and
application of an electric field can generate a mechanical response. This leads to applications such as
spark generation and transducers. Finally, the ability of an applied field to switch the direction of the
dielectric polarisation leads to the possibility to use thin-film ferroelectrics for digital memory devices.

0
100

200
300
Temperature (K)

400

Figure 1.11: Temperature-dependence of the dielectric polarisation in BaTiO3 , showing the three phase
transitions.

7 Here

we meet the temptation to be pedantic: surely the example of BaTiO3 also involves small displacements of atoms.
We will see later in the module that the distinction we will
make is that displace phase transitions involve an explicit
phonon mechanism rather than an orderdisorder process.

Chapter 1. Phenomenology

10

Polarisation (Cm2)

0.8

data for the the temperature-dependence of the dielectric polarisation of BaTiO3 and PbTiO3 were given
above in Figures 1.11 and 1.12 respectively.

0.6

Diffraction

0.4

0.2

0.0
300

400

500
600
Temperature (K)

700

800

Figure 1.12: Temperature-dependence of the dielectric


polarisation in PbTiO3 . Note the discontinuity at the
transition temperature.

For the antiferromagnetic and atomic site ordering transition there is no macroscopic quantity that
directly gives a measurement of the order parameter. However, we can nevertheless measure the order
parameter directly through Bragg diffraction.
The intensity of a reflection with Miller indices hk ` is proportion to | F (hk `)|2 , where the structure
factor F (hk `) is the Fourier transform of the electron (for X-ray diffraction) or nuclear (for neutron
diffraction) densities. In the case of x-ray diffraction, the electron density can be described as a set of
atom-based electron densities convolved with the the points representing the positions of the atoms,
and in the Fourier transform this is equal to a produce of the separate Fourier transforms. In the case of
neutron diffraction, the nuclei can be considered to be point particles, but the scattering is weighted by
the strength of the nucleusneutron interaction. Either way, the structure factor can be written as a sum
over all atoms in the unit cell:
F (hk`) =

f j exp


2i (hx j + ky j + `z j ) exp(8 hu2j i sin2 /2 )

(1.5)

where f j represents the Fourier transform of the electron density (in the case of X-ray diffraction) or the
strength of the scattering of the neutron from the nucleus, ( x j , y j , z j ) are the fractional coordinates of the
atom, and hu2j i is the mean-square displacement of an atom. This latter quantity is a level of detail we
do not need to consider here and therefore we will set hu2j i = 0 in the analysis here.
We consider the case of ordering of two atoms, denoted here as A and B, over the two sites in a
body-centred-cubic crystal, Figure 1.13. In the disordered phase each site is equivalent and half occupied
on average by both atomic species. When the structure orders, one site becomes preferentially occupied
by the A atoms and the other site by the B atoms to form a crystal with the CsCl structure type. We
consider the sites that preferentially order with the A atoms as label 1, and the sites that preferentially
order with the B atoms as label 2, and define A1 and B1 = 1 A1 as the probably that the site labelled 1
is occupied by A and B atoms respectively. Similar definitions exist for site 2, with the constraint that
A2 = B1 and B2 = A1 . The value of A will vary between values of 0.5 and 1.0 as the system orders
from the high-temperature disordered phase to the low-temperature ordered phase. Given that the

1.6. How to observe a phase transition

11

order parameter correspondingly changes in value from 0 to 1, a little thought will convince you that
A1 = (1 + )/2 and B1 = (1 )/2, and thus = A1 B1 . The structure factor for Bragg scattering
from this system can be written as




F (hk`) = A1 f A + B1 f B exp (2i (hx1 + ky1 + `z1 )) + A2 f A + B2 f B exp (2i (hx2 + ky2 + `z2 ))
(1.6)
where hk ` are the Miller indices for the reflection, x1 , y1 , z1 and x2 , y2 , z2 are the fractional coordinates
of atomic sites 1 and 2 respectively, and f A and f B are the scattering factors of atom species A and B
respectively.
In our specific example of a transition from bcc to CsCl structures, we can assign coordinates
x1 , y1 , z1 = 0, 0, 0 and x2 , y2 , z2 = 12 , 21 , 12 . Thus the structure factor can be written as

 

F (hk `) = A1 f A + B1 f B + B1 f A + A1 f B exp (i (h + k + `))
(1.7)
where we have replaced the site 2 occupancies with their corresponding values from site 1. We note
that the exponent on the right hand side has value +1 if h + k + ` has an even value, and value 1 if
h + k + ` has an odd value. In the even case, we have


F (hk `)even = A1 + B1 ( f A + f B ) = f A + f B
(1.8)

Figure 1.13: Top: Body-centred cubic structure, in


which each site is occupied 50% by atom type A and
50% by atom type B. Bottom: ordered phase of the
same system, giving rise to the CsCl-type structure.

This tell us nothing of interest. But the case where h + k + ` has an odd value gives the structure factor


F (hk `)odd = A1 B1 ( f A f B ) = ( f A f B )
(1.9)
Provided that there is a large enough difference between values of f A and f B , the order parameter can
be measured directly through measurements of the intensities of Bragg reflections for odd values of
h + k + `, noting that the intensity is proportional to | F (hk `)|2 and hence proportional to 2 . When
= 0, the structure is bcc for which the condition h + k + ` odd is a systematic absence. The intensities
of Bragg reflections can be measured relatively easily using x-ray or neutron diffraction; which one
chooses may depend on the relative sizes of f A and f B for the two types of radiation.8
This exact analysis is only applicable for site-occupancy order-disorder phase transitions, but the
general idea holds true for many types of phase transition. For example, a small reworking would make
this approach apply to atoms hopping between two or more sites, and for phase transitions in KH2 PO4

8 Elements

that have similar numbers of electrons will have


very small differences between the atomic scattering factors
for x-rays, but there are also many pairs of elements with
different atomic numbers that have similar neutron scattering
factors. In some cases the neutron scattering factors have
negative values, which can amplify the difference between
scattering factors of two elements.

12
2.5
d11
dij (1012 CN1)

2.0
1.5
1.0

d14

0.5
0.0
200

300

400

500

600

700

800

900

1000

Temperature (K)

Figure 1.14: Variation of two components of the piezoelectric tensor of quartz on heating through the phase
diagram. One component falls to zero at the transition
temperature.

Chapter 1. Phenomenology

and BaTiO3 atom type A could be the H+ or Ti4+ cations, and atom type B could represent a vacant site.
The ideas also have direct application to magnetic phase transitions, although only neutrons, which
have their own magnetic moment, couple directly to the atomic moment and hence diffraction from
magnetic structures is carried our using neutron diffraction. Magnetic diffraction has a somewhat more
complicated form for the structure factor than that for atomic diffraction, but the general principle we
have outlined above holds good.
Displacive phase transitions can also be observed by measurements of specific Bragg peaks that have
zero value of the structure factor in the higher-symmetry phase. Such reflections may correspond to
systematic absences due to certain symmetry elements glide planes and screw axes which are lost
through the phase transition, and the intensities of reflections that are systematically absent due to
these symmetries again give a measurement of 2 . Moreover, when any type of phase transition leads
to a multiplication of the size of the unit cell (as seen in antiferromagnetic and atomic site ordering
phase transitions, but also found in many displace phase transitions) reflections that will correspond to
fractional values of the Miller indices in the high-temperature high-symmetry will have integer values
in the low-symmetry phases and hence the possibility for non-zero intensities.
Of course, in the modern era it is possible to measure the entire diffraction pattern quickly (for
powder diffraction on synchrotron x-ray sources or even neutron sources diffraction patterns these
can be measured in minutes). In that case, it is easy to use modern structure refinement methods to
understand the complete set of atomic structural changes, and deduce the order parameter from the
whole structure. We will see examples of more complicated phase transitions whose properties are
understood from diffraction later in course.

Indirect macroscopic measurements of the order parameter


There are a number of properties of a material that are directly affected by a phase transition and which
will contain quantitative information about the order parameter. One important quantity, which we will
meet later, is the spontaneous strain that can be generated by a phase transition. Consider, for example, a
phase transition between a high-temperature phase with cubic lattice type and a low-temperature phase
with tetragonal lattice type. This might involve a small expansion of the c axis and small contraction of
the a and b axes. These small expansions and contractions define the spontaneous strain, and typically
this strain will be proportional to or 2 depending on the symmetry. Other physical properties, such
as elastic and piezoelectric coefficients, and refractive index, can change in value due to the phase

1.7. Order of a phase transition

13

transition, typically in a way proportional to 2 . In the best examples, the values of the coefficients, or
differences in coefficients, will vanish at the phase transition. An example is given in Figure 1.14.

1.7 Order of a phase transition


We have noted that the value of the order parameter, reflecting the size of the changes to the structure
due to the phase transition, will change continuously on heating up to the transition temperature
Tc . The way that the order parameter changes at the phase transition defines two types of phase
transition. First-order phase transitions are defined as phase transitions in which the order parameter
changes discontinuously at the actual transition. This illustrated in Figure 1.15. In contrast, continuous
phase transitions or second-order phase transitions are defined as phase transitions in which the order
parameter decreases continuously to zero at the transition, typically with the same sort of power law as
we identified earlier, Equation 1.2.9 The temperature-dependence of the order parameter associated
with a second-order phase transition is represented in Figure 1.15. Second-order phase transitions,
particularly for temperatures in the vicinity of Tc , have a fascinating science of their own, which will
become apparent and more explicit throughout the module.
We anticipate the further discussion by noting here that second-order phase transitions are frequently
accompanied by extreme behaviour of some physical and thermodynamic properties, frequently with
divergence of some properties. For example, the heat capacity may show a small step or a divergence. A
phase transition in which the order parameter is associated with the formation of a dialectic polarisation,
called a ferroelectric phase transition will see the divergence of the dielectric constant. Some phase
transition will see the vanishing an an elastic constant at the transition temperature, and hence the
divergence of an element of the elastic compliance tensor.

9 This

terminology originally arises from a scheme due to Paul


Ehrenfest, who proposed that an nth-order transition was
one in which the nth derivative of the free energy was discontinuous at the phase change (but all lower derivatives were
continuous). In practice only the distinction between firstorder and higher-order transitions turned out to be useful.

1.8 General aspects of the thermodynamics of a phase transition


We will make considerable use of thermodynamics to understand phase transitions. The important
concept is the free energy, although we will be a little cavalier for the moment about whether we mean the
Gibbs (constant pressure) or Helmholtz (constant volume) free energy.10 The power of the free energy is
that the global minimum in the free energy function for all variables defines the thermodynamic state of
equilibrium, and the phase transition defines how the state of equilibrium changes as we scan through
changes in external variables such as temperature or pressure.

10 Most

experiment is performed at constant pressure, and most


theory is performed at constant volume!

Chapter 1. Phenomenology

14

Second-order

Order parameter

First-order

Tc

Tc
Temperature

Figure 1.15: Representation of the variation of the order parameter at first-order (left) and second-order (right)
phase transitions.

We can imagine that the free energy F can be written as a function of the order parameter
whether this is actually true is not proven, but it sounds like a reasonable presumption. On this basis,
the function F ( ) will have a minimum at the equilibrium value of . For T > Tc , the equilibrium
value of = 0, and thus F ( ) will be a function with a minimum at = 0. This is illustrated for a
second-order phase transition in Figure 1.16. For T < Tc , the equilibrium value of = 0 , and thus
F ( ) will have minima at 0 . This is illustrated for a second-order phase transition in Figure 1.16. At
a second-order phase transition, the point F ( = 0) will switch from being a minimum to a maximum
at T = Tc .
It is pertinent now to define the susceptibility, . For any phase transition it can be imagined that
there is a field that couples in some way to , such that / = , and hence changes in are
proportional to with as the constant of proportionality. We can write the free energy F in partial
form as
dF = SdT + d
(1.10)
This defines the field as
=

dF
d


(1.11)
T

1.8. General aspects of the thermodynamics of a phase transition

15

2nd-order

1st-order

F()

F()
T > Tc
T > Tc
T = Tc

T = Tc

T < Tc

T < Tc

Figure 1.16: Representation of the variation of the free energy with order parameter at temperatures above, exactly
at, and below the transition temperature for second-order (left) and first-order (right) phase transitions.

The susceptibility then follows as


1 =

d2 F
d
=
d
d 2

(1.12)

We have noted that for a second-order phase transition at T = Tc the function F ( = 0) switches
from a minimum to a maximum, which means that d2 F/d 2 = 0 at T = Tc . This immediately implies
that as T Tc . The statistical mechanics aspects of the susceptibility are discussed in the
Appendix to this chapter, particularly how it is linked to fluctuations of the order parameter.
For a first-order phase transition, we expect the F ( ) function to have minima of equal depth at
both = 0 and = 0 exactly at the transition temperature, allowing for a discontinuous jump in the
equilibrium value of on changing temperature through the transition temperature.11 This is illustrated
in Figure 1.16.

11 Beware,

we are treating the function F ( ) as a symmetric


function of , which it frequently is but not necessarily so.
There are some changes in symmetry that require F ( ) to be
an asymmetric function.

Chapter 1. Phenomenology

16

1.9

Seeds of a theoretical model


This module will combine both empirical results for phase transitions and the underpinning theoretical
ideas. At this stage we have not done enough to develop a detailed theory, but it is interesting to
anticipate what sorts of theoretical framework we are likely to develop.
The important factor is that we have individual atoms taking part in the phase transition. We could
define a crystal as a set of individual atomic oscillators vibrating in their own potential energy wells. To
have the structure deform at low temperature, the potential energy wells need to have minima either
side of the mean atom position, and at low temperature the atom will be confined to one minimum or
the other. Lets define this mathematically as
1
1
E(u j ) = 2 u2j + 4 u4j
2
4

(1.13)

At high temperature, where the kinetic energy is higher than the height of the barrier between the two
minima in the potential energy well, the atom will vibrate back and forth. So you can imagine there will
be a phase transition. However, so far we do not have an adequate model for a phase transition. Since
all the atoms are independent, there is no way in which they can move together to break the symmetry.
Instead the mean position of each atom of each atom will gradually drift towards the centre point on
heating but without a sharp transition.
The missing factor is something that couples the displacements of neighbouring atoms. The simplest
form is a term of the form u j u j0 to couple the displacements of atoms of labels j and j0 . Extending the
energy to cover all atoms in the crystal and all neighbours, we now have


1
1
2
4
E = 2 u j + 4 u j + J u j u j0
(1.14)
2
4
j
j,j0
The bad news is that we are not yet in a position to make use of this reasonable-looking function. So
instead, lets go for something simpler, and set the constraint that the displacements u j can only have
values 1 corresponding to the positions of the minima of the function (the units dont matter at the
moment). Thus we no longer need the double-well potential energy since we only have the energies at
the two minima. Hence the relevant part of the energy is
E = J u j u j0
j,j0

(1.15)

1.10. Various types of phase transitions

17

This situation is formally the same as that of a collection of interacting magnetic moments that are
constrained to only two orientations. In fact, this model is sufficiently important that it has its own
name, the Ising model, named after the PhD student who first studied it. We will meet this model later,
but for now it suffices to note that the model was solved by Ising only for the one-dimensional case,
and the rather more difficult two-dimensional case was not solved for another 20 years. There is no
solution yet to the three-dimensional version.12 So we are forced to use some tricks to make progress.
One trick is to replace the neighbouring atoms by a copy of an average atom, so we will now write
E = J u j hui

(1.16)

where

hui =

1
N

uj

(1.17)

This model can be solved using standard thermodynamics, and will lead to a temperature-dependence
of the order parameter of the form
| Tc T |1/2

(1.18)

Compare this with the general equation we noted in Equation 1.2, where here = 1/2. We will learn
that this is an important exponent, and the value deduced here is typical for models with this sort of
approximation.
You may not be comfortable with using this trick; but in many cases it is the best that can be done.
This trick is so important that it has its own name, the mean-field approximation, reflecting the fact that
we have replaced the field of interactions seen by each atom by an average field. We will revisit the
mean-field approximation at several points in this module.

1.10 Various types of phase transitions


Phase transitions are found that involve atom positions, electrons, electronic states, momentum states
and magnetic moments.
Antiferromagnetic Ordering of the magnetic moments (atomic or electronic) in a way that does not
create a sample magnetisation.

12 And

some people believe that there will be no analytical


solution.

18

Chapter 1. Phenomenology

Displacive Phase transition involving small symmetry-breaking displacements of atoms.


Ferrimagnetic Ordering of two or more types of magnetic moments (atomic or electronic) to create a
sample magnetisation that can be reversed by application of a magnetic field.
Ferroelastic Displacive phase transition accompanied by a spontaneous strain that can be reversed by
application of an external stress.
Ferroelectric Displacive phase transition which breaks the centre of symmetry and gives rise to a
spontaneous dielectric polarisation. The direction of the polarisation can be reversed by application
of an electric field.
Ferromagnetic Ordering of the magnetic moments (atomic or electronic) to create a sample magnetisation that can be reversed by application of a magnetic field.
Incommensurate Phase transition in which the ordering wave vector is not related to the vectors that
define the reciprocal lattice. There are examples in magnetic, charge density and displacive
systems.
Liquidgas Phase transition between the liquid and gas states of matter.
Liquid crystal Ordering of the positions of long-chain molecules whose orientations are already ordered.
Orientational order/disorder Ordering of the orientations of molecules (usually highly symmetry small
globular molecules) whose positions are already ordered.
Reconstructive Phase transition between two distinctly-different crystal structures.
Site ordering Ordering of atoms of different types over equivalent sites in the crystal.
Superconducting Ordering of the momenta of electrons within the conduction bands of a metal to
create a state of zero electrical resistance.
Superfluid Ordering of the momenta of atoms within a fluid to create a state of zero viscosity.
Superionic Ordering of the positions of small mobile ions within a crystal defined by the positions of
the other atoms.

1.11. Summary

19

Verwey Ordering of different electronic states of one type of ion, e.g. ordering of Fe2+ and Fe3+ cations
in Fe3 O4 .

1.11 Summary
In this chapter we have introduced a number of concepts. Briefly these are
1. The role of symmetry, and in particular the way in which a phase transition is accompanied by a
change in symmetry.
2. Phase transitions occur at a fixed temperature called the transition or critical temperature, with
symbol Tc .
3. Magnetic phase transitions involve ordering of the orientations of magnetic moments, with
neighbouring moments ordering in the same direction (ferromagnetic) or in opposite directions
(antiferromagnetic).
4. The overall sample magnetisation created by a ferromagnetic phase transition can be reversed by
application of an external magnetic field.
5. Displacive phase transitions involve small symmetry-breaking displacements of atoms.
6. Ferroelectric phase transitions are the dielectric equivalent of ferromagnetic phase transitions, in
that they involve creation of a dielectric polarisation that can be reversed by application of an
external electric field.
7. Atomic site ordering phase transitions involve ordering of the positions of different types of atoms
on equivalent sites in the crystal.
8. The size of the symmetry-breaking displacements or ordering can be quantified by a dimensionless
variable called the order parameter.
9. Experimentally phase transitions can be measured by some macroscopic manifestation of the order
parameter, by measurements of the intensities of Bragg peaks that disappear on heating through
the phase transition, by measurement of some crystal property that depends on the value of the
order parameter, or by crystal structure analysis from Bragg diffraction.

Chapter 1. Phenomenology

20

10. We defined two types of phase transition first-order or second-order depending on whether
the order parameter changes discontinuously or continuously at the phase transition.
11. We assume that the free energy is a function of the order parameter, and identified how this
function should change with temperature through a phase transition.
12. We introduced the basic idea of a mean-field theory as a way to simplify theoretical analysis of a
model for a phase transition; the essential idea is that any atom interacts with an average rather
than instantaneous environment.
13. One of the important crystal structures in the study of phase transitions is the perovskite structure.

1.12 Appendix on the statistical mechanics approach to susceptibility


The partition function Z is defined as
Z=

exp(Ei /kB T ) = exp( Ei )


i

(1.19)

where Ei is the energy of state i, and = 1/kB T. The value of the partition function is that some
thermodynamic properties can be defined in terms of the partition function or its derivative. For
example, the average energy h Ei can be written as

h Ei = Z 1 Ei exp( Ei ) =
i

ln Z
1 Z
=
Z

(1.20)

and the free energy can be written as


F = ln Z

(1.21)

Furthermore, a number of relationships can be derived starting from the partition function; one of these
is the susceptibility.
A particular state of the system subject to applied field might have an order parameter of value i .
The energy of this state is equal to i . The partition function is therefore
Z=

exp( i )
i

(1.22)

1.13. Short questions to test your basic knowledge

21

It follows that

h i = Z 1 i exp( i ) =
i

and

h 2 i = Z 1 i2 exp( i ) =
i

1 Z
Z

(1.23)

1 2 Z
2 Z 2

(1.24)

The susceptibility (equation 1.12) can therefore be obtained as


=

h i
1 2 Z
1

=
2

Z
Z2

2

Comparison with equations 1.23 and 1.24 leads to the result




= h 2 i h i2

(1.25)

(1.26)

What this equation tells us is that the value of the susceptibility reflects the size of the fluctuations in the
order parameter, in a similar manner to the fact that the heat capacity reflect the fluctuations in energy.
Thus a divergent susceptibility reflects large fluctuations in the value of the order parameter around a
second-order phase transition.

1.13 Short questions to test your basic knowledge


1. State the difference between the paramagnetic, ferromagnetic and antiferromagnetic states.
2. Define the order parameter for a ferromagnetic phase transition.
3. How does the order parameter vary with temperature at temperatures just below a phase transition?
4. What is the primary difference between a first-order and second-order phase transition?
5. Why does the magnetisation of a ferromagnetic material show hysteresis in the presence of a
time-varying magnetic field?
6. What are domains and domain walls?

Chapter 1. Phenomenology

22

7. For a binary alloy that has a body-centred cubic structure above an ordering phase transition, and
a CsCl structure in the ordered state, write equations for the occupancy of both sites in terms of
the order parameter.
8. What is the defining feature of a ferroelectric phase transition.
9. What are the points of similarity between ferroelectric and ferromagnetic phase transitions?
10. Why can the intensities of certain Bragg reflections measured by x-ray or neutron diffraction give
a direct measurement of the size of the order parameter associated with a phase transition?
11. Define the susceptibility associated with a phase transition in terms of the free energy and the
order parameter.
12. How does the susceptibility change at a second-order phase transition?

1.14 Longer questions


1. One family of crystals that undergoes many important phase transitions is the perovskite family,
of general formula ABO3 , where A and B are cations. The parent crystal structure is primitive
cubic with fractional coordinates A at 12 , 12 , 12 ; atom B at 0, 0, 0; and oxygen atoms at 12 , 0, 0; 0, 12 , 0
and 0, 0, 12 .
(a) Draw a plan of the crystal structure and a three-dimensional perspective image of the ideal
cubic perovskite structure.
(b) Show the AO and BO bonds on the three-dimensional perspective image you have just
made.
(c) Identify the positions of points in the structure that have the property that all atoms at any
position r with respect to that point is also found at r. Such a point is called a centre of
symmetry.
(d) Identify the positions of axes parallel to x, y and z about which the crystal can be rotated by
90 and look like the same structure such an axis is called a 4-fold rotation axis.

1.14. Longer questions

23

2. In some phase transitions in the perovskite structure, the main changes are displacements of the
cations in the z direction relative to the positions of the oxygen anions.
(a) Describe the effects on the centre of symmetry and 4-fold rotation axes caused by positive
displacements.
(b) Is the phase transition ferroelectric? Why or why not?
(c) If the oxygen atoms have charge 2e, and the two cations have charge +3e each, where e
is the elementary charge with value 1.60218 1019 C, and if the cation displacements are
relative to the positions of the oxygen anions, what is the size of the polarisation P
both 0.1 A

generated if the three unit cell parameters are equal to 4 A?


3. Consider the domain walls shown in Figure 1.7. Assuming nearest-neighbour interactions which
give a negative energy when neighbouring moments are oriented in the same direction and a
positive energy when neighbouring moments are oriented in opposite directions, compare the
energies of the three types of domain wall shown.

Chapter 2

Magnetic phase transitions

2.1 A macroscopic view of magnetism


Magnetism in the solid state has been known for a very long time lodestones, naturally occurring
pieces of the magnetic mineral magnetite, were known to the ancient Greeks but requires some rather
sophisticated quantum mechanics to describe properly. Our interest in magnetism stems from the fact
that the study of magnetic phases led to some of the earliest results in the general theory of phase
transitions. In particular, some early models of magnetic solids are examples of mean-field theories,
an important class of model that reproduces many experimentally observed features of real phase
transitions while remaining computationally tractable.
To begin, we recall some key facts from electromagnetism and atomic physics. Unlike the case with
electric charge, there is no such thing as a magnetic monopole (as Gauss law of magnetism tells us); so
the basic building unit of a magnetic solid is the magnetic dipole. Recall a current I around a circular
loop of area A has a magnetic moment given by
=

I dA.

(2.1)

We define the magnetisation M as the total magnetic moment per unit volume. In the simplest case of a
linear material, the magnetisation of the material will depend on the applied field according to
M = H

(2.2)

where the constant of proportionality is known as the magnetic susceptibility. If we tabulate values of
the susceptibility , we find a fairly wide range of values, both positive (e.g., copper sulfate pentahydrate
has = +176 106 ) and negative (e.g., sodium chloride, = 13.9 106 ). If the susceptibility is
negative, so that the magnetisation opposes the applied field, we call the material diamagnetic; if it is
25

26

Chapter 2. Magnetic phase transitions

positive, so that the magnetisation lines up with the applied field, we call the material paramagnetic. The
origins of these phenomena are in quantum mechanics, and to discuss them would be beyond the scope
of this module. As a rule of thumb, paramagnetism is associated with unpaired electrons, which, as we
have seen, act as magnetic dipoles that line up with an external field; diamagnetism is a weaker effect
present in all materials, but only apparent if there are no unpaired electrons to cause the material to
become paramagnetic.

2.2

1 Total

electronic angular momentum because the nucleus, too,


has a spin that gives rise to hyperfine structure.

2 Not

exactly equal, because of corrections that come from the


theory of quantum electrodynamics, but near enough for our
purposes. We will ignore these corrections completely from
now on.

A microscopic view of magnetism


As equation 2.1 shows, it is natural to think of the magnetic moment as arising from the moving charge.
On the other hand, the current also implies a moving mass and hence an angular momentum. It is not
too difficult to show that the magnetic moment is also proportional to this angular momentum, and so
we could also view the moment as arising from the angular momentum. It turns out that as we move
into the quantum world, it is this description that will be most helpful.
In atoms, where we must use a quantum description, magnetism is associated with both the orbital
angular momentum L and the intrinsic angular momentum, or spin S. The terminology here comes from
the rough image that magnetism is caused by the currents of the electron either orbiting the nucleus
(orbital angular momentum) or spinning on its own axis (spin angular momentum). But this is not
an entirely accurate picture for instance, as far as we know the electron is a point particle without
any axis to rotate around! and it is best to think of these simply as quantum numbers describing the
wavefunctions of the electrons in an atom. The sum of these two components gives the total electronic
angular momentum J = L + S.1 Just as we argued for the classical case, the magnetic moment is
proportional to the angular momentum; we write
g J B
=
J
(2.3)
h
where g J , called the Lande g-factor, is a constant of proportionality that can be calculated from J, L, and
S. Broadly, its function is to take quantum effects into account (it would be 1 in the classical case).
The constant B = eh/2me = 9.274 1024 J T1 is the Bohr magneton, a sensible unit for atomic-scale
magnetism (it is approximately equal to the z component of the spin magnetic moment of an electron).2
The minus sign arises because of the negative charge on the electron.
A complication arises here: as usual when describing angular momentum in the formalism of
quantum mechanics, we can simultaneously specify only the magnitude of a vector and its component

2.3. Non-interacting atoms in a magnetic field: paramagnetism

27

along one specified direction, conventionally taken to be the z axis.3 So, for instance, the total angular
momentum J is typically described by the quantum numbers J, which must be an integer or half-integer,
and M J = J, ( J 1), ( J 2), . . .. These quantum numbers correspond to eigenvalues

|J|2 = J ( J + 1)h 2

(2.4)

J z = M J h .

(2.5)

3 That

is, the operators for the magnitude and z component of


the vector commute. Of course, by symmetry the operator
for the magnitude commutes with the component for any
component, but since the operators for different components
do not commute we have to choose a single special axis to
describe with a quantum number.

and

The values of these quantum numbers can be determined from their equivalents for the spin and orbital
components L, ML , S, and MS , which in turn depend on the corresponding values `, m` , and ms for each
individual electron in the atom (and the constant s = 12 for electrons). Exactly how we do this depends
on the system under consideration; for more details consult a book on atomic physics.4
The classical expression for the energy of a magnetic dipole in an external field B is E = B; if
we choose the z direction to be the direction of the external field we can rewrite this as E = ( z ) B.
This makes writing down the equivalent quantum expression very easy: from equations 2.3 and 2.5 we
have
E = g J B M J B.
(2.6)

2.3 Non-interacting atoms in a magnetic field: paramagnetism


As we have seen in the previous chapter, phase transitions are collective, not individual phenomena:
they depend on whole regions of atoms behaving in similar ways. We must therefore consider what
happens if we put a collection of atoms in an external field. At this point we will simplify things by
considering the specific case of a spin- 12 ion; that is, we will set S = 12 and L = 0, which gives J = 12 ,
and g J = 2. This describes the behaviour of a single electron, and also of the common ions Ti3+ and
Cu2+ . More importantly for our purposes, it is the simplest model of magnetism we can consider: if
J = 12 then there are just two possible values for M J = 12 . Recall that this quantum number represents
the component of the angular momentum, and hence of the magnetic moment, in the direction of the
external field. Thus a spin- 12 species in an external field must either align with the field or against it,
just as we saw for the simple magnetic model of section 1.2. Equation 2.6 simplifies to show that the
corresponding energy is
E = B B.
(2.7)

4I

quite like Molecular Quantum Mechanics by Atkins and Friedman.

Chapter 2. Magnetic phase transitions

28

5 Of

course we could repeat this calculation for general J.


Slightly less obviously, even this is not an entirely innocent
assumption: many magnetic materials, like iron, are metals, so that the electrons responsible for the magnetism are
delocalised and not associated with any particular atom.

Our starting point, then, will to let each atom in our model solid be one of these spin- 12 ions.5 If we
apply an external field to this solid, equation 2.7 suggests that it will be energetically favourable for our
material to become magnetised: that is, the magnetic moments will align with the external field. Our
question is to quantify this relationship: how much magnetisation will a given field produce?
From statistical mechanics we find that the probability of being in a state with energy E is proportional
to exp( E/kB T ); so the average value of the z component of the magnetic moment in our system is the
weighted average




B
B


B exp k BT B exp k BT
B B
B
B




hz i =
.
(2.8)
= B tanh
B
B
kB T
exp B + exp B
kB T

To express this in terms of macroscopic properties, consider what would happen at saturation, when
every magnetic moment is aligned with the external field. In this case we would simply have hz is = B .
But the total magnetisation M hz i, and moreover the constant of proportionality will be independent
of the degree of saturation. Thus


M
hz i
B B
.
(2.9)
=
= tanh
Ms
h z is
kB T

M/Ms

0.5
0

0.5
1

kB T

2
0
2
y = B B/kB T

Figure 2.1: Magnetisation as a function of external


field and temperature for a spin- 12 paramagnet.

Re-expressing this, in order to find the system magnetisation M we must solve the simultaneous
equations
M
B
= tanh y
y= B
(2.10)
Ms
kB T
Using a pair of simultaneous equations might seem an odd way of expressing the relationship between
M and B, but it will be the easiest to extend when we come to derive a similar relationship for a
ferromagnet.
Graphically, this relationship is shown in Figure 2.1, which should make qualitative sense: there is a
linear region where the external field is not too big, in which the alignment of the magnetic moments
is proportional to the external field; and then after a while we reach saturation, with all the moments
lining up completely with the external field, so that the magnetisation cannot increase no matter how
great the external field. Moreover, this is equally true whichever direction the external field is in (i.e.,
whether y is positive or negative).
Lets focus on that linear region and calculate the susceptibility . In this region B is small, so we
can make the approximation tanh y y. Furthermore, since the magnetisation will also be small we

2.4. Interacting atoms in a magnetic field: ferromagnetism

29

can write B = 0 ( H + M ) 0 H, as in a vacuum, and we also express the saturation magnetisation in


terms of the number density of magnetic ions n as Ms = nB . Putting all of this together gives us the
expression
n0 2B
=
.
(2.11)
kB T
As an example, consider the spin- 12 paramagnet titanium(III) fluoride, TiF3 . The unit cell volume is
3 and there are two magnetic titanium ions per unit cell. Thus we find n = 2/(116.35 1030 )
116.35 A
and hence = 4.5 104 at 300 K. This agrees well with the experimental value of 4.7 104 .
Interesting applications of magnetic materials tend to come not from paramagnetism or diamagnetism, but from arrangements such as ferromagnetism where the magnetic moments remain aligned
even in the absence of an external magnetic field. In order to model this we need to consider not
just interactions between individual magnetic moments and the external field (as with our spin- 12
paramagnet model), but also interactions between the magnetic moments.

2.4 Interacting atoms in a magnetic field: ferromagnetism


Attentive readers will have noticed that we have not yet done quite what we claimed to. In the previous
section we noted that phase transitions depended on atoms acting collectively yet this is not currently
possible in our model, which just considers each atom interacting separately with the external field.
We are now going to rectify this by making a simple addition to the model. We will suppose that each
magnetic site sees not only the external magnetic field but also an additional field proportional to the
average of all the other magnetic sites that is, to the total magnetisation:
Btotal = Bexternal + M

(2.12)

where is some constant of proportionality. Note in particular that this doesnt give any special status
to any particular pairs of atoms: each magnetic site is affected by every other site equally, meaning for
instance that nearest neighbour interactions are treated no differently from interactions between atoms
on opposite sides of a macroscopic sample. For this reason this is referred to as a mean-field model
because we dont consider the specific field at a particular magnetic site, just the mean field over
the whole sample. Despite this rather dramatic approximation we will see that this model (originally
developed by the French physicists Pierre Curie and Pierre Weiss) turns out to be remarkably successful.

Chapter 2. Magnetic phase transitions

30
Under this transformation our model becomes
M
= tanh y
Ms

y=

B ( B + M)
.
kB T

(2.13)

1
It is easiest to get a feel for how these simultaneous equations behave by solving them graphically.
M/Ms

0.5
0

0.5
1

T > Tc
T = Tc
T < Tc

0
y

Figure 2.2: In the absence of an external field, the


Weiss model predicts that the magnetisation will be
zero at all temperatures above some critical value Tc .
Below this temperature, our simultaneous equations
have two solutions (dashed lines) that approach Ms
as the temperature drops to zero.

Consider first the case where there is no external field, shown in Figure 2.2. At a relatively low
temperature, the solutions are given by the intersections of the black curve with the blue line. As we
would expect from a ferromagnet, there is one solution with most of the spins aligning in one direction,
and one with most of the spins aligning in the other (light blue lines). (There is also a solution with zero
magnetisation, but this is metastable: if exactly half the spins point in each direction, our mean-field
equation is satisfied, but if there is just one more spin pointing up than down, it will influence its
neighbours, and hence their neighbours, until most of the spins in the sample point up.)
At a relatively high temperature, we consider the red line instead. We can see that now the only
solution is zero magnetisation: so much thermal energy is available to the system that the spins are
randomly disordered.
The transition between these two regions is known as the Curie temperature: above this temperature,
the material is paramagnetic, but below it is ferromagnetic. This qualitative phase transition behaviour
can be seen in real ferromagnets, and is therefore a substantial success of our simple mean-field model.
Lets calculate the Curie temperature in terms of our other model parameters. The phase transition
will occur when the slope of the straight line is equal to the slope of the hyperbolic tangent function at
x = 0. Thus
kB TC
d
=
tanh x | x=0 = 1
B M
dx

(2.14)

giving
TC = B

n2B
Ms
=
kB
kB

(2.15)

where again we have set the saturation magnetisation to nB .


Just as for the paramagnetic case, we calculate the susceptibility by making the approximation that,

2.5. Critical exponents revisited

31

for small B, tanh y y. Thus

n0 2B
1
M
=

.
H
kB ( T TC )
T TC

1
(2.16)
0.5
(2.17)
(2.18)

Compare this result to the case for the paramagnet 1/T. This result is known as the Curie-Weiss law.
Now consider the case where we apply an external magnetic field. As before, we have to solve the
pair of simultaneous equations (2.13). If we change the temperature but keep the external field constant,
we end up with a constant intercept on the vertical axis (Figure 2.3). We can read off the solutions
(dashed horizontal lines) from the intercepts at different temperatures. We see that there is no longer a
sharp phase transition, nor is there a finite temperature at which the magnetisation reaches zero, unlike
the zero-field case. In an external magnetic field, some fraction of the magnetic dipoles will always
align with that field.

M/Ms

M
( B + M )
= B
Ms
kB T


1
B B
H
B
M
=
0 B

Ms
kB T
kB T
kB T

0.5
1

T > Tc
T = Tc
T < Tc

B
M
s

0
y

In the previous chapter we introduced the critical exponent ; repeating equation 1.2 for convenience,
recall that
(2.19)

In this context, the symbol has the precise meaning that


f (x) x

means

lim

x 0

ln | f ( x )|
= .
ln | x |

(2.20)

As we previously noted, for many purposes you can think of this as simply meaning proportionality; in
fact, however, it is a slightly weaker condition. In other words, while f ( x ) x implies f ( x ) x , the
reverse is not true.6
Lets evaluate the value of for this model in the absence of an external field (recall that in the
presence of an external field, there is no phase transition!). The trick here is to note that we are only

Figure 2.3: Some proportion of the magnetic dipoles


will always align with an external field; thus in the
presence of such a field there is no phase transition.

2.5 Critical exponents revisited

| M( T Tc )| | Tc T | .

6 Exercise:

find an example of this fact.

Chapter 2. Magnetic phase transitions

32

really interested in temperatures just below the phase transition (Curie) temperature. It is convenient to
define the reduced temperature
T TC
(2.21)
=
TC
so that we are interested in the case where is small and negative. Rearranging and using the value of
TC from equation 2.15 gives
Ms
T = TC ( + 1) = B
( + 1)
(2.22)
kB

and substituting back into equation 2.13 gives


M
= tanh
Ms

M
1

+ 1 Ms


.

(2.23)

Now just below the Curie temperature M will be very small (since this is a continuous phase
transition), so we can replace the tanh function by its Taylor expansion
Figure 2.4: Schematic diagram showing the magnetisation of our ferromagnet model as a function of temperature in zero external field.

tanh x = x 31 x3 + O( x5 )
giving
1
M
1
M

Ms
+ 1 Ms
3( + 1)3

(2.24)

M
Ms

3
.

(2.25)

Ignoring the unstable solution M = 0, we can rearrange to give



Tc

M
Ms

2

= 3 ( + 1)2

(2.26)

which, remembering that  1 by assumption, shows that


M ( )1/2 ( TC T )1/2
Figure 2.5: Phase diagram for our model of ferromagnetism, showing a line of first-order transitions ending
at the critical point.

(2.27)

and hence that = 12 for this model (Figure 2.4). This result was anticipated in equation 1.18, and is
in fact a general characteristic of mean-field models. The phase diagram of our model, showing the
regions available as a function of temperature and external field, is shown in figure 2.5.

2.6. Successes and failures of the mean-field model

33

2.6 Successes and failures of the mean-field model


We conclude this chapter by considering the ways in which the predictions of our mean-field model are
borne out or not by experiment and simulation.
By definition, when considering the energy of a single magnetic site, a mean-field model takes every
other site in the system equally into consideration. The extent to which this is a good assumption will
depend on the specific interactions between sites in the system we are trying to model. The exchange
interaction is rather short-range: since it depends on the extent of orbital overlap between the atoms
involved, it is usually assumed only to apply between nearest neighbours.7 In other materials such as
ferroelectrics, where the interactions between neighbouring sites are somewhat longer-range, one might
expect a mean-field model to fit somewhat better.
A related consideration is that at no point does this model take into account the dimensionality of the
system it describes: a one-dimensional chain is treated exactly the same as a three-dimensional solid.
Experimentally, however, such systems behave rather differently; in fact, low-dimensional magnetism
continues to be a very active field of research. As the dimensionality increases, or as the number of
neighbours each atom interacts with increases, we might expect models to approach the mean-field
limit. Indeed, if we consider spin- 12 systems in which each atom is only allowed to interact with its
nearest neighbours8 and compare the phase transition temperature to its mean-field value Tmf , we find
that the transition occurs at T = 1.76Tmf for a two-dimensional square lattice (four nearest neighbours),
T = 1.33Tmf for a three-dimensional simple cubic lattice (six nearest neighbours), and T = 1.23Tmf for a
cubic close packed lattice (twelve nearest neighbours).
Similarly, the critical exponent turns out to be 18 for a 2d square lattice and approximately 0.33 for
either a simple or close-packed cubic lattice in 3d; recall for comparison that it is 12 for a mean-field
model.9 The fact that both 3d models have the same critical exponent is intriguing: it demonstrates
the key role played by the dimensionality of the interactions. We can also ask how fast these critical
exponents converge to the mean-field value as we increase the dimensionality of the system. The
remarkable answer is that all the critical exponents have exactly the values predicted by mean-field
theory once a system has four or more dimensions, as we will see in the following chapter.10
The key point that unifies this discussion is that mean-field theory is incapable of taking into
consideration fluctuations in the order parameter. As a material is cooled towards a phase transition,
regions with the same value of the order parameter will begin to form and gradually increase in size,
so that a spin in the middle of a clump will have quite a different local environment to one on the

7 Or

occasionally second-nearest neighbours.

8 This

9 Note

10 Six

is the Ising model again!

that the 3d value is not exactly equal to 31 .

critical exponents are commonly discussed: we have


already met . We will see in the problems below, and the
remaining exponents , , , and next chapter.

Chapter 2. Magnetic phase transitions

34

boundary. Obviously, though, this effect cannot be expressed by our single constant . This is why, as
we have seen, mean-field theory predicts phase transitions at lower temperatures than local-interaction
models (and real life): clumping gives almost all of the energetic benefits of long-range order while
retaining almost all of the entropic benefit of the disordered state. In fact, as we will see, a remarkable
feature of real phase transitions is that every length scale, from very small clumps to long-range order,
is important at the transition point. Of the two general properties of models we have discussed, the
number of nearest neighbours is more important over short distances, while the dimensionality is more
important over the long range.
A more formal way of expressing the same idea is to note that the average of a function of some
random variable is not normally the same as the function evaluated at the average of that variable. In
other words, for a general function f ( ) depending on the order parameter, h f ( )i 6= f (h i). However,
only the quantity on the right is accessible to mean-field theory. To progress further, we need to be able
to know more about values of other than its average. We will find out how to do this, and revisit the
ideas discussed in this section, in Chapter 7.
In the meantime, however, we will consider some of the successes of mean-field theory. First, it
usually predicts the qualitative phase diagrams of three-dimensional systems correctly: indeed, the
phase diagram of figure 2.5 is correct for a 3d ferromagnet. As we have seen, this is particularly true
where each atom is able to interact with many others, because of the close packing or the long-range
interactions or both. Moreover, it greatly simplifies models that are potentially extremely complex,
regardless of the dimensionality, and can provide a starting point for more accurate calculations or
may even be the only tractable computational method available. For these reasons it is worth exploring
further. In the next chapter we will meet a phenomenological model that turns out to be a mean-field
theory, although expressed slightly differently to the Weiss model, and see how it can be used to describe
a very wide range of structural phase transitions. Perhaps the most important success of mean-field
theory is that it is empirically very useful, as we shall see.

2.7

Summary
1. The magnetisation of a material is its magnetic moment per unit volume; the susceptibility is the
constant of proportionality relating the magnetisation to the external field applied.
2. The energy of an individual magnetic ion in an external field is quantised and depends on the

2.8. Short questions to test your basic knowledge

35

quantum number M J .
3. In the Curie-Weiss model of ferromagnetism, each ion is assumed to see a field proportional
to the total magnetisation of the material, in addition to any external field applied. This is an
example of a mean-field model.
4. This model has a ferromagneticparamagnetic phase transition at the Curie temperature TC . The
susceptibility in the paramagnetic phase is proportional to ( T TC )1 (the Curie-Weiss law); thus
the critical exponent = 1.
5. Similarly, the critical exponent =

1
2

for this model.

6. On the other hand, mean-field theories are often substantially more computationally tractable
than any alternatives.

2.8 Short questions to test your basic knowledge


1. Define the terms hysteresis and susceptibility.
2. Define the terms remanence (also known as remnant magnetisation) and coercivity (also known as
coercive field).
3. Explain why the model in equations 2.10, which describes the behaviour of a paramagnetic solid,
does not undergo a phase transition.
4. Explain how these equations can be modified to model ferromagnetism (and hence to undergo a
phase transition).
5. What order is the phase transition observed when a ferromagnet in zero external field is cooled
from high to low temperature?
6. What order is the phase transition observed when a ferromagnet at low temperature is placed in
an external field that is then varied from positive to negative?
7. Is the Curie-Weiss transition temperature higher or lower than the true value? Explain why.
8. Is the Curie-Weiss critical exponent higher or lower than the true value? Explain why.

Chapter 2. Magnetic phase transitions

36

2.9

Longer questions
1. Not all magnetic materials are ferromagnets: many sorts of magnetic order are possible. The next
simplest case is an antiferromagnet; that is, a material in which it is energetically most favourable
for neighbouring spins to be antiparallel. In order to extend the Weiss model to this situation,
we divide the atoms into two groups: those that would be aligned up in the case of perfect
antiferromagnetic order, and those that would be aligned down (Figure 2.6). In the analogue
to equation 2.12, we suppose that each group sees the external field plus an additional term
proportional to the magnetisation of the other group.

Figure 2.6: An antiferromagnet can be conceptually divided into two groups of atoms with different alignments in
the case of perfect order.

11 T
N

because this is known as the Neel temperature, after


another French physicist, Louis Neel, who won the Nobel
Prize in 1970.

Repeat the analysis of section 2.4 for this case (still considering only spin- 12 atoms, unless you are
feeling particularly ambitious) to show that this model also gives a phase transition, and calculate
the temperature TN at which this will occur in terms of the model parameters (i.e., an analogue to
equation 2.15).11 Hence show that the susceptibility of an antiferromagnet in this model obeys

1
.
T + TN

(2.28)

Explain in physical terms why diverges at the phase transition for ferromagnets (equation 2.15)
but not for antiferromagnets (equation 2.28).
2. Another critical exponent that we will later find useful is , which describes the response of a
system at a phase boundary to an external field. In terms of a magnetic phase transition, it is

2.9. Longer questions

37

defined such that


M B1/

at T = TC .

(2.29)

Calculate for the Weiss model.12 (Hint: recall that by definition of , equation 2.20, we only
need to consider the behaviour as B 0.)
3. Sketch the hysteresis loop for a Curie-Weiss magnet at T/TC = 0.8. For simplicity, nondimensionalise by plotting M/Ms against B B/kB TC . Show as much detail as possible; in particular,
calculate numerical values for the coercivity (coercive field strength) Bc and remnance (remnant
magnetisation) M0 . Compare your diagram to Figure 1.6 and explain any differences.

12 Like

, its value turns out to be common to all mean-field


models.

Chapter 3

Landau theory

3.1 Introduction
In this section we will explore an intuitively-derived form of the free energy of a system that has a
phase transition. The results we obtain will be quite general, and in many cases will be found to be
consistent with experimental data. As we go through this module we will explain where our intuition
comes from, and what assumptions lie behind our results.
The work of this module is known as Landau Theory, after the famous Russian Physicist Lev Landau
(Figure 3.1), who first developed the ideas outlined here. The key insight of Landau was to recognise
that the thermodynamics associated with a phase transition could be captured through the definition of
the order parameter.
Whilst the theory to be developed in this section is developed from intuition rather than derived
from other theories, and whilst it contains a number of parameters whose values can be set by the
researcher from comparison with experimental data, the approach correctly captures the symmetry
details of the phase transition, and because of this much of what the theory predicts turns out to be
rigorously true.

3.2 Quantification of the free energy

Figure 3.1: Lev Landau, 190868, Nobel Prize 1962.

In Chapter 1, we looked at the qualitative shape of the free energy curves, Figure 1.16. We are in a
position now to make a guess as to their quantitative form.
Lets look at this empirically. We can write the free energy as an expansion in terms of the order
parameter :
1
1
F ( T, ) = F0 ( T ) + A( T ) 2 + B( T ) 4 +
(3.1)
2
4
39

Chapter 3. Landau theory

40

This expansion only contains even terms if we expect F ( ) = F ( ), which will frequently be the case
on symmetry grounds, and will always be the case for second-order phase transitions. If we truncate the
series at the 4th-order term, the coefficient B( T ) needs to be positive at all temperature so that F ( T, )
will be positive and increasing in value for large values of | |; it is generally taken that B( T ) is in fact a
constant value for all temperatures. The coefficient A( T ) can be of either sign. If A( T ) > 0, F ( T, ) has
p
a single minimum at = 0. But if A( T ) < 0, F ( ) has minima at = A( T )/B( T ).
Through this form of the function we can see how a phase transition can arise, if A( T ) has a
temperature dependence that means that it changes sign at the phase transition temperature Tc . In
this case, F ( ) will have the two minima for T < Tc , but a single minimum at = 0 for T > Tc . The
simplest representation that achieves this is to set A( T ) = a( T Tc ), where a is a constant value for all
temperatures, so that A( T ) > 0 when T > Tc , and A( T ) < 0 when T < Tc . We therefore write the free
energy function as
1
1
F ( T, ) = F0 ( T ) + a( T Tc ) 2 + b 4
2
4
1 Of

course, F0 ( T ) is not constant with temperature or pressure,


but it essentially only acts as a background to the free energy
and any derived property.

2 In

the spirit of not caring whether F represents the Helmholtz


or Gibbs free energy, we could equally work with internal
energy E instead of enthalpy H, but experimentally H is the
quantity that is measured.

where both a and b are assumed to have that values that are independent of temperature. Going forward,
will drop the constant term F0 ( T ) because it contains no information about the phase transitions.1
The free energy consists of two terms, the enthalpy2 H and an entropic term TS:
F ( ) = H ( ) TS( )

that because we are dropping F0 we are also dropping


the corresponding H0 and S0 , which is why we can have a
function that apparently has a negative amount of entropy.
Given than S 0 as T 0, we might guess that we could
now write S0 = 12 a02 , where 0 is the value of the order
parameter at T = 0. However, we will later show that our
model doesnt work as a correct thermodynamic theory in
the limit T 0, a fact that should not worry us much at the
moment. We will see in subsequent topics that the model we
have developed is a good representation of the free energy
of several mean-field models in the vicinity of Tc .

(3.3)

The enthalpy is the part of F ( ) that doesnt have the term linear in T:
1
1
H ( ) = aTc 2 + b 4
2
4

3 Recall

(3.2)

(3.4)

Thus the entropy is what is left, and follows as


1
S( ) = a 2
2

(3.5)

We see that the entropy decreases with increasing , which is what we expect when defines a degree
of order, as in a magnetic or atom site-ordering phase transition.3 The enthalpy (equation 3.4) has a
double-well function reflecting the fact that 6= 0 at T = 0.

3.3. Results for second-order phase transitions

41

Figure 3.2: Crystal structure of the perovskite SrTiO3 in the cubic (left) and tetragonal (right) phases, showing the
TiO6 octahedra as solid objects, the oxygen atoms as small red spheres, and the Sr cations as large green spheres.
The phase transition is accompanied by rotations of the octahedral about the crystallographic [001] axis, with
rotations having opposite sense in neighbouring layers.

3.3 Results for second-order phase transitions


Temperature dependence of the order parameter
We have already noted that the state of equilibrium is that corresponding to the minimum of the free
energy. Thus the equilibrium value of at any T is obtained from solving F/ = 0. Thus we obtain
( T > Tc )

= 0

( T < Tc ) =
a/b ( Tc T )1/2

(3.6)
(3.7)

We noted in Chapter 1 that this result, namely the exponent 1/2, is characteristic of any mean-field
theory. This works here because the free energy contains no contribution from spatial fluctuations in .

Chapter 3. Landau theory

42

4 Second-order

Diraction intensity

order-disorder phase transitions such as magnetic transitions, atomic site-ordering transitions, and liquid
gas transtions are notable exceptions.

400
300
200
100
0

80

90
100
Temperature (K)

110

Figure 3.3: Temperature-dependence


of the Bragg scat

tering from the 12 21 32 reflection from the tetragonal
phase of SrTiO3 , indexed with respect to the parent
cubic phase. The intensity values are proportional to
2 . The dashed straight line is a guide to the eye.

It turns out that many second-order phase transitions are accurately described by this theory,
particularly displacive phase transitions.4
One example is the displacive phase transition in the perovskite SrTiO3 . The atomic displacements,
which correspond to small rotations of the TiO6 octahedra about the z axis to give a change in symmetry
from cubic to tetragonal, are shown in Figure 3.2. It appears that the unit cell size is doubled in both
the horizontal directions of the figure. In fact, alternate layers along the z direction have rotations in
opposite directions, which gives a doubling of the unit cell inthis direction
too. Thus we can state

that the ordering wave vector for this phase transition is q = 12 12 12 . The order parameter for this
phase transition has been studied by measurements of the intensity of a Bragg peak that vanishes in
the high-temperature phase; this particular peak vanished because its Miller indices are not integers in
the setting of the parent cubic phase. The experimental data are shown in Figure 3.3. In Chapter 1 we
showed that the intensities of Bragg peaks that vanish at the phase transition are proportional to 2 .
Thus we expect, if our prediction of the temperature-dependence is correct, we expect the intensity to
fall linearly to zero on heating up to the phase transition. You can judge whether the data corroborate
the theory or not; to my mind it looks good for most temperatures but with some small deviation from
the theory at temperatures close to the phase transition. This might indicate some interesting physics
we havent yet taken account of.

Critical exponents
The relationship ( Tc T ) , where = 1/2 from Landau theory in particular and any mean-field
treatment in general, is characteristic of a type of equation we will often meet in this module for a range
of physical properties. The exponent is called a critical exponent, and we will see some interesting
physics associated with the critical exponents. As we progress in this section we will define a number of
critical exponents and derive their values within the framework of Landau theory.

Temperature dependence of the heat capacity


The heat capacity is defined as
C = T

S
T

(3.8)

loo

110

120

T[*K]

Velocities of longitudinal and shear waves propagating in a [!00] direction in SrTi03 fr


Liithi and Moran (1970). Key: (S) shows results at 50 MHz; at 30 MHz; ! at 16 MH
and + at 10MHz.
43

3.3. Results for second-order phase transitions

Fig. 1.17

Where we use the symbol to denote the change in the heat capacity due to the phase transition. Using
the previous equation for S( ) and ( T ) we obtain
(3.9)

a2
C ( T < Tc ) = T
2b

(3.10)

CP (J/mol/K)

C ( T > Tc ) = 0

I 73

0.55
J/mol/K
O.13cal/mol
K

. ~

50
46

This suggests that the heat capacity should show a small jump equal to a2 Tc /2b at the phase transition.
I
I
I
I
I
I
I
I
~8
7O0
702
70~
106
7O8
770
772
17~
100
104
108
112
Data for our example of the displacive phase transition in SrTiO3 are shown in Figure 3.4.
T(K) - The specific heat of SrTiO3 near the 110
K phase transition.
(After Franke and Hegenba
Temperature
(K)
This might seem gratuitous at this point, but in general we can write
1974.)
C ( T > Tc ) | T Tc |
C ( T < Tc ) | T Tc |

(3.11)I f we separate
o u t 3.4:
the Temperature-dependence
term involving the order pof
a rthe
a m eheat
t e r susceptibilities
zi(k)
Figure
capac(3.12)

ity da(K)
CP on heating through the tetragonalcubic phase
=(~B(K)+kBT~lF(K,qs,+kjl)12zi(k)A(K+qsi+k),
transition
in SrTiO3 . The
small discontinuity at the
d~
ik
transition temperature is highlighted.

(I.3.2

where = 0 = 0 are the critical exponents for the heat capacity. We do so here because such a power
where aB(K) represents the other terms which p r o d u c e a b a c k g r o u n d scattering. T
law with non-zero values of the critical exponents are found in a number of phase transitions. The value
= 0 is characteristic of a phase transition described by mean-field theory.

Temperature dependence of the susceptibility


The susceptibility was defined (equation 1.12) by the differential
1 = 2 F/ 2

(3.13)

From our definition of the free energy we obtain


1 = a( T Tc ) + 3b 2

(3.14)

This gives us the following results for temperatures either side of Tc :


1

T > Tc = a ( T Tc )

(3.15)

T < Tc = 2a ( Tc T )

(3.16)

We see that as T Tc , a result that is found in many second-order phase transitions. An


example is given in Figure 3.5; this shows the dielectric susceptibility measured around the ferroelectric

Chapter 3. Landau theory

44

phase transition in the perovskite LiTaO3 . At this point we also have a prediction that the variation of
1 with temperature is a factor of 2 larger below Tc than above. Experimentally is is found that there
is a difference, but the factor is often larger than 2, although you might be persuaded that in the case
of LiTaO3 the difference is not large. In the following chapter we will extend the thermodynamics to
include the effect of strain, which will naturally account for the factor being larger than 2.
We could generalise the result for the susceptibility as

103

T >Tc | T Tc|

T <Tc | T Tc|
20

(3.17)

(3.18)

104 /

In Landau theory, as in all variants of a mean-field theory, we have critical exponents for the susceptibility
= 0 = 1. In order-disorder phase transitions that do not follow mean-field theory, we typically have
> 1. It is generally the case that 0 = .

10

3.4

We can add a field term of the form to the free energy, where is the general field that can change
the value of . At T = Tc the free energy is equal to

0
700

800

900

Field-dependence of the order parameter at the transition temperature

1000

Temperature (K)

F (, T = Tc ) =

Figure 3.5: Temperature-dependence of the dielectric


constant (approximately equal to the susceptibility,
top) and its inverse (bottom) on heating through the
ferroelectric phase transition in LiTaO3 .

1 4
b
4

(3.19)

Minimisation of F ( ) with respect to gives the result


3 =

(3.20)

More generally we could write 1/ , where the critical exponent = 3 in a mean-field theory, but
can take other values when mean-field theory is not applicable.

3.5

Taking account of spatial variations


To this point we have assumed that the system under study is the same at all points in space across all
length scales. We can extend the analysis to allow for a variation of the order parameter across space,
by including a term that depends on the spatial gradient of the order parameter. If we now consider the

3.5. Taking account of spatial variations

45

high-symmetry phase and work to quadratic order, we can write the free energy of the crystal as an
integration over space:
F=

1
a( T Tc )
2

{ (r)}2 dd r +

1
f
2

{ (r)}2 dd r

(3.21)

where (r) is the spatial gradient of the order parameter. This equation is easier to handle if we
perform a Fourier transform and integrate over reciprocal-space variables. To do so we need to know
about the Fourier transform of a differential. The Fourier transform of any function y( x ) is

F [y( x )] =

1
2

F [y0 ( x )] =

1
2

y( x )eikx dx

(3.22)

y0 ( x )eikx dx

(3.23)

Writing y0 ( x ) = dy( x )/dx, we have

R
This can be integrated in parts (where the integral can be represented as udv, where u = eikx and
dv = y0 ( x )dx:
Z

1
1

F [ f 0 ( x )] =
y( x )eikx
ik
y( x )eikx dx
(3.24)
2
2

We anticipate that y( x ) 0 as x , giving the result

F [ f 0 ( x )] = ikF [y( x )]

(3.25)

Writing (k) = F [ (r)], we can rewrite equation 3.21 as


F=

1
2(2 )d

Z 


a( T Tc ) + f k2 | (k)|2 dd k

(3.26)

We can use equipartition to write


D
E
1
1
a( T Tc ) + f k2 | (k)|2 = kB T
2
2
and thus we have
D

E
| (k)|2 =

kB T
a( T Tc ) + f k2

(3.27)

(3.28)

Chapter 3. Landau theory

46

D
E
In the high-temperature phase, h| (k)|i = 0, and | (k)|2 gives the fluctuations in the value of
(k). Assuming that theDphase transition
occurs for k = 0 and here we have written the
E
D mathematics
E
this way it is clear that | (k)|2 as T Tc at k = 0. From statistical mechanics, | (k)|2 /kB T
is equivalent to the susceptibility (k), and we see the divergence of the susceptibility at the phase
transition for k =D 0 as was
E derived earlier, equation 3.16.

The function | (k)|2 is the Fourier transform of the correlation function g(r) for the variable (r):
D
E
g(r) = h (r) (0)i = F [ | (k)|2 ]

(3.29)

This is obtained from the fact that g(r) is the convolution of (r) with itself g(r) = (r) (r) and
from the mathematical identity that the Fourier transform of a correlation function is the product of the
Fourier transforms of the two functions within the convolution. Thus we obtain from equation 3.28
g (r) =

kB T exp(r/ )
f
r d 2

(3.30)

where
s
=

f
= 0
a( T Tc )

T Tc
Tc

 1
2

(3.31)

has dimensions of length. It gives the length scale over which the value of g(r) decays. Put another
way, given that g(r) gives a measure of the extent to which values of the order parameter are correlated
in size, gives the size of fluctuations. It is called the correlation length.
In terms of critical exponents, we write
T >Tc | T Tc|
T <Tc | T Tc|

(3.32)
(3.33)

where within Landau theory we have = 0 = 1/2. We also have the general result at T = Tc that
g ( r ) r 2 d
In this analysis the critical exponent = 0.

(3.34)

3.6. Validity of Landau theory

47

3.6 Validity of Landau theory


Based on what we have studied so far, we are able to form an argument concerning the validity of
Landau theory. We expect Landau theory to be appropriate if the fluctuations in the order parameter
are rather less than the size of the order parameter. Given that the fluctuations become larger close to
the phase transition, and given that the order parameter goes to zero close to the phase transition, we
anticipate that Landau theory might fail at temperatures close to the transition temperature. Actually
this point can be calculated, but rather than derive an equation for this (which would need to be
evaluated for different systems, something we wont do) we use an argument based on dimension.
We can also consider that Landau theory will break down when the free energy associated with
fluctuations of the order parameter become similar to the overall Landau free energy over the length
scale of the fluctuations. This length scale is determined by the size of the correlation length .
The free energy per unit volume can be obtained by twice integration the heat capacity, giving
F t 2

(3.35)

where t = | T Tc | /Tc .
We can associated an energy of kB T with a fluctuation, which exists over the length scale of . Hence
we can define a free energy per unit volume associated with fluctuations as
Ffluct

kB T
td
d

(3.36)

The important point concerns whether the ratio Ffluct /F diverges as t 0. The criterion that this should
not diverge is that
d > 2
(3.37)
This gives the idea of the upper upper critical dimension, d? = (2 )/, namely the dimension of space
above which Landau theory (or more generally, mean-field theory) works for all temperatures.5 Using
the values of the critical exponents derived here. we find d? = 4.6
This analysis has been appropriate for the isotropic case. When there are significant directional
effects, such as in uniaxial ferroelectrics where there are strong directional long-range dipole interactions,
the effects can be to reduced the value of d? , often down to a value of 3 but in some cases (particularly
for ferroelastic transitions) to d? < 3. The analysis also doesnt comment on the range of temperatures
over which mean-field behaviour breaks down even for cases where d? > 3. As noted at the start of

at dimensions exactly d = d? there are small logarithmic corrections to the mean-field behaviour.

5 Actually
6 There

is also the concept of the lower critical dimension, which


defines the dimension at which and below the fluctuations
are so important that the phase transition doesnt occur. For
the Ising model, this lower critical dimension is 1. Again
other aspects of dimension are important, particularly the
dimension of the order parameter. For the case where spins
can order along any of three dimensions (the Heisenberg
model) the lower critical dimension is 2.

Chapter 3. Landau theory

48

this section, this range can be calculated. In general, the range of temperatures over which mean-field
theory breaks down is partly determined by the length scale of interactions, since this will feed through
to the size of the correlation length. Therefore for systems in which direct or indirect interactions are
long-range, mean-field and hence Landau theory will work at temperatures very close to Tc .

3.7

Ferromagnetism, the mean-field approximation, and Landau theory


In the previous analysis of the ferromagnetic phase transition, Section 2.4, we started from a model with
no interactions between the magnetic sites, and then introduced this interaction in a mean-field way.
Often, however, we already have an idea of which specific interactions exist between sites, and want to
find the best mean-field approximation to this Hamiltonian. As an example of how to do this, we will
now tackle the same problem from this direction.
Specifically, we will solve the spin- 21 Ising model, which has the Hamiltonian
nn

H = J Si S j B Si
i,j

(3.38)

where the Si = 1 and the sum labelled nn is to be taken over nearest neighbours. We will need the
fact that if every one of the N sites has z of these nearest neighbours alternatively, if each of the N
atoms has z bonds then the number of nearest-neighbour interactions, or bonds, is given by
Nbonds = 21 Nz.

(3.39)

We need to find a mean-field approximation to the first term in our Hamiltonian (3.38); the second
can be used unchanged as it does not depend on the details of which sites are nearest neighbours. To
do this we write each Si as
Si = + Si

(3.40)

3.7. Ferromagnetism, the mean-field approximation, and Landau theory

49

where = hSi is the order parameter. Then our Hamiltonian becomes


nn

H = J ( + Si )( + S j ) B Si
i,j

(3.41)

nn

J 2 + (Si + S j ) B Si

(3.42)

= JNbonds 2 Jz (Si ) B Si

(3.43)

= JNbonds 2 ( Jz + B) Si

(3.44)

i,j

where we have assumed in (3.42) that Si S j is negligible on average (this is where we actually make
the mean-field approximation) and used (3.39). We have thus achieved our mean-field aim of removing
the sum over nearest neighbours from the Hamiltonian.
We ultimately want to calculate the free energy F using the statistical relationship
F = kB T ln Z ,

(3.45)

so our next step is to calculate the partition function:




JNbonds 2 ( Jz + B) i Si
exp

kB T
Si =1




N
Jz + B
JNbonds 2
= exp
exp kB T Si
kB T
Si =1 i




JNz 2
Jz + B N
= exp
2 cosh
2kB T
kB T

Z=

(3.46)
(3.47)
(3.48)

giving7

7 You

F = 21 JNz 2 kB TN ln 2 cosh

Jz + B
kB T


(3.49)

We know that there is no phase transition where B 6= 0, so lets investigate the B = 0 case more
thoroughly. We can relate this free energy to the Landau-type form simply by taking a Taylor expansion:
ln 2 cosh x = ln 2 + 12 x2

1 4
12 x

+ O( x6 )

(3.50)

may remember from statistical physics modules that, if


N systems each have individual partition function z, then the
total partition function is Z = z N provided that the systems are
distinguishable. Can you see why this is the case here?

Chapter 3. Landau theory

50
giving




Jz 2
Jz 4
1
+ 12 kB T
F/N
kB T ln 2
kB T
kB T



3
Jz
Jz
1
= kB T ln 2 + 12 Jz 1
Jz
2 + 12
4
kB T
kB T
2
1
2 Jz

particular the fourth-order coefficient does not vanish at


the transition temperature, showing that this is indeed a
second-order phase transition.

point of interest is that the part of the free energy


proportional to temperature is, as usual, the entropy; as
expected this is equal to k ln 2 for each site that can take one
of two possible spin orientations.

(3.51)
(3.52)

This now looks familiar: there is a second-order coefficient that becomes negative at the transition
temperature kB Tc = Jz and a a positive fourth-order coefficient.8 In other words, applying the mean-field
approximation to this Hamiltonian gives a standard Landau second-order phase transition.9

8 In

9 Another

1
2 kB T

3.8

First-order phase transitions


There is no reason to constrain the fourth-order coefficient in the expansion of the free energy to have a
positive value; all we require is that the highest-order relevant term needs to be positive so that F ( ) is
positive for large values of | |. Thus with a negative fourth-order term, we need to include a sixth-order
term, and can write
1
1
1
F ( ) = a( T T0 ) 2 b 4 + c 6
(3.53)
2
4
6
where, for reasons that will be clear shortly, we now use T0 instead of Tc in the prefactor of the
second-order term.

Transition temperature and the temperature-dependence of the order parameter


The temperature-dependence of F ( ) was previously shown in Figure 1.16. The striking result is that
there is a well-defined temperature at which F ( ) has three minima with F = 0. This is the temperature
at which the system undergoes the first-order phase transition.
We can calculate this temperature by noting that we have two equations. First we have the equation
for which F = 0 at the minimum point, and we have the condition that F/ = 0 at the same point.
Writing the transition temperature as Tc , we have the two equations, after dividing by factors of and
numerical factors:
1
1
a( Tc T0 ) b 2 + c 4 = 0
2
3
a( Tc T0 ) b 2 + c 4 = 0

(3.54)
(3.55)

3.8. First-order phase transitions

51

We can equate these two equations, subtracting the constant term a( T T0 ) and dividing by 2 to give
1
1
3b
b + c 2 = b + c 2 2 =
2
3
4c

(3.56)

This is the jump in the value of 2 at the phase transition. We substitute into equation 3.55 to obtain
Tc = T0 +

3b2
16ac

(3.57)

For temperatures below Tc the temperature-dependence of the order parameter can be obtained from
the minimisation of the free energy:
1 F
= a( T T0 ) b 2 + c 4 = 0

(3.58)

from which we obtain

p
b
b2 4ac( T T0 )
= +
(3.59)
c
2c
The example of the first-order ferroelectric phase transition was previously shown in Figure 1.12, where
the order parameter corresponded to the dielectric polarisation and could be measured directly.
2

Latent heat and heat capacity


We can also derive other thermodynamic properties. One of the most important ones is the latent heat
at the phase transition. This can be calculated from the change in enthalpy at the phase transition
1
1
1
3abT0
aT
H = aT0 2 b 4 + c 6 =
= 0 2
2
4
6
8c
2

(3.60)

The heat capacity can be obtained from the entropy, which as before is taken from the term in the
free energy with the temperature prefactor:
C=T

S
2
= T aTc
T
T

(3.61)

We anticipate that 2 changes more with temperature as we heat towards the transition temperature, so
that the heat capacity will increase gradually on heating.

Chapter 3. Landau theory

52
Susceptibility
The susceptibility at a first-order phase transition will vary as
1

T > Tc = a ( T T0 )
1

T < Tc

= a( T T0 ) 3b + 5c

(3.62)
(3.63)

This has a change in value of 1 from 3b2 /16c to 12b2 /16c at the phase transition.

3.9

The case when the free energy is allowed to have odd-order terms
There are some phase transitions where we do not require to maintain the symmetry of the system
under a change in sign of the order parameter, in which care we will have F ( ) 6= F ( ). This most
often occurs when the phase transition involves the loss of a 3-fold rotation axis.
The free energy in this case is allowed to have odd-power terms in the order parameter (except the
linear term; note the negative sign of the cubic term):
1
1
1
a( T T0 ) 2 b 3 + c 4
(3.64)
2
3
4
At the phase transition temperature, this function will have two minima rather than three, one at =
and another at a non-zero value of . As before, we can write a pair of equations corresponding to the
phase transition temperature for F = 0 and F/ = 0, dividing by factors of and numerical factors:
F ( ) =

1
2
a( T T0 ) b + c 2 = 0
3
2
a( T T0 ) b + c 2 = 0

(3.65)

The solution of these two equations gives a jump in the value of the order parameter from = 0 to
= 2b/3c at the phase transition. We can now obtain the transition temperature as
b
c
2b2
Tc = T0 + 2 = T0 +
a
a
9ac

(3.66)

1 = a( T T0 ) 2b + 3c 2

(3.67)

The susceptibility is given as

Unlike the case of the first-order phase transition evaluated earlier, there is no jump in the value of at
the phase transition.

3.10. Tricritical phase transitions

53

3.10 Tricritical phase transitions


There are many many materials where external variables such as pressure and internal variable such as
chemical composition can change the fundamental parameters that determine the properties of a given
material. These include the coefficients in the Landau expansion of the free energy, and one interesting
case is where there is a change in the value of the coefficient b that involves a change in sign. In such
cases, application of pressure or tuning of the chemical composition will lead to the condition b = 0, the
case that is exactly on the border between the phase transition being first-order or second-order. Such a
phase transition is called a tricritical phase transition.
We can write the free energy as
1
1
a( T Tc ) 2 + c 6
(3.68)
2
6
Minimisation of the free energy with respect to leads to the result
a1
4
(3.69)
=
( Tc T )1/4
c
The exponent = 1/4 is a factor of 2 smaller than obtained for second-order phase transitions.
The other interesting result is the heat capacity. As before, the entropy can be written as
r
1 2
1 a3
S = a =
(3.70)
( Tc T )1/2
2
2
c
The heat capacity follows as
r
S
1 a3
C = T
=
T ( Tc T )1/2
(3.71)
T
4
c
In contrast to the case of second-order phase transitions, for which Landau theory predicts a small jump
in the heat capacity at the phase transition, the heat capacity diverges at a tricritical phase transition.
We can derive values for other critical exponents (see the chapter problems), and compare the
results from Landau theory for second-order and tricritical phase transitions in table 3.10. Note that for
tricritical phase transitions the upper critical dimensionality d? = 3.
F=

3.11 Relevant aside: Definition of the strain and elastic tensors


For many practical examples, we need to account for the fact that many phase transitions are accompanied by an elastic strain, and within Landau theory we are able to understand how this can significantly

Table 3.1: Values of the critical exponents and upper


critical dimensionality from Landau theory of secondorder and tricritical phase transitions.

Exponent

Second order

Tricritical

0
1/2
1
3
0
1/2

1/2
1/4
1
5
0
1/2

d?

Chapter 3. Landau theory

54

Table 3.2: Examples of material tensors.

Rank

Tensor

0
1

Temperature, volume, pressure


Electric and magnetic fields,
dielectric polarisation, magnetisation,
applied force, normal to a surface,
thermal expansivity
Susceptibility, elastic strain, elastic stress,
refractive index, thermal conductivity
Piezoelectric coefficients,
electro-optic coefficients
Elastic constants

2
3
4

10 But

for non-linear behaviour it may be necessary to go to


higher order.

affect many properties of a phase transition. Before we can develop the formalism to include strain, we
need to understand that strain is a tensor quantity.

Tensors and coordinate transformations


Tensors are multi-dimensional sets of numbers that have some specific properties that allow us to change
the definition of the coordinate system without changing the practical outcome of a calculation. We
start by considering an anisotropic dielectric medium. Application of an electric field E = ( E1 , E2 , E3 )
will generate a polarisation P = ( P1 , P2 , P3 ), but in an anisotropic material there is no reason to assume
that P will be in the same direction as E, particularly if some directions in the material are more easily
polarised than others. So we therefore write what looks like a matrix equation:

P1
11 12 13
E1

(3.72)
P2 = 21 22 23 E2
P3
31 32 33
E3
The arrays P and E are called first-rank tensors, and the array is called a second-rank tensor. In
principle this can be extended to any rank we require, up to infinite-rank, but for crystalline materials
we only need to go to fourth-rank to describe linear quantities.10 Examples are given in Table 3.11.

Tensors and coordinate transformations


So far these look like vector and matrix quantities, but what characterises these as tensors is the process
by which we can rotate the set of axes. Our convention for the choice of axes only comes down to our
personal preference, yet the direction of P relative to the orientation of the crystal and the direction
of E must transcend our choice of axes. So our first task is to define the transformation operation
that converts a tensor between one coordinate system and another. We start with an orthogonal set of
axes defined by the vector set (x1 , x2 , x3 ), and a rotated set of axes with orthogonal vectors (x10 , x20 , x30 ),

normalised such that |xi | = xi0 = 1 for all i, as illustrated in Figure 3.6. We define a direction cosine as
aij = xi0 x j

(3.73)

which equals the cosine of the angle between the two vectors xi0 and x j , and is defined in a way to give
the correct signs.

3.11. Relevant aside: Definition of the strain and elastic tensors

x3

55

x3

x3

x3

x1

x2
x2

x1

x2

x2
Figure 3.6: Definition of two sets of orthogonal axes
in three-dimensional space.

Figure 3.7: Example of two sets of orthogonal axes


in two-dimensional space.

The matrix that transforms a vector in one space to the other is defined as the matrix of the direction
cosines:

a11 a12 a13

a = a21 a22 a23


(3.74)
a31 a32 a33
such that a vector in the first space is transformed to the other through the multiplication
q0 = a q

(3.75)

This is illustrated in two dimensions in Figure 3.7, where we show the vector P as defined in two spaces.
The transformation is given as

1
0
0
0

P0 = a P = 0 cos sin P2
(3.76)
0 sin
cos
P3

Chapter 3. Landau theory

56

Note that the transpose of the matrix a gives the transformation back to the original axes; thus a1 = aT .
Now we consider how this transformation works on a second-rank tensor. We define the first-rank
tensors P and Q, which are coupled through a second-rank tensor T via the equation P = T Q. The
same equation in the transformed coordinate system is written as P0 = T0 Q0 . We recall that P0 = a P,
and we can write P0 = a T Q. Similarly the equation P0 = T0 Q0 can be written as P0 = T0 a Q. We
thus can combine these two equations for P0 to write a T Q = T0 a Q and thus we have a T = T0 a.
It thus follows from the property of a that
T 0 = a T a 1 = a T aT

(3.77)

Put into expanded form, we have


Tij0 =

aik a j` Tk`

(3.78)

aki a` j Tk0 `

(3.79)

k`

and
Tij =

k`

Extension to higher order is straightforward; the case of the transformation of a third-rank tensor
will give the recipe
33

0
Tijk
=

ai` a jm akn T`mn

(3.80)

`mn
23

and for a fourth-rank tensor we have

32

0
Tijk
` =
22

22

x3

32
23

aim a jn akp a`q Tmnpq

(3.81)

mnpq

Stress and strain


Stress tensor

x2
33

Figure 3.8: Definition of the components of the stress


tensor, noting the direction of the force and the face
operated on.

Although both force and length extensions are vector quantities, stress and strain form second-rank
tensors. In the case of stress, we actually have two quantities to couple, namely the direction in which
the force is applied and the direction of the normal to the face of the material on which the force acts.
Thus a force can be applied both in the direction of the face normal tensile stress or at right angles to
the face normal shear stress. These stress terms are shown in Figure 3.8

3.11. Relevant aside: Definition of the strain and elastic tensors


The stress tensor is written as

11

= 21
31

12
22
32

57

13

23
33

(3.82)

Strain tensor
The strain tensor takes a bit more effort to understand. Lets start with some things we know. Tensile
strain is simply defined as the linear expansion of an object:

x1

x2

u1

where the strain is defined as


e=

u2

u
x

(3.83)

reflecting the fact that u = ex.


Shear strain might be defined as the change in angle subtended by the corner of an object:

u1
x1

u2
x2

Chapter 3. Landau theory

58

but here the displacement is now in the direction orthogonal to the vector we are considering.
We can create a combination of the two types of strain, focusing on three points, Figure 3.9.

uC

Point A is displaced according to


uA = (e11 x1 , e21 x1 )

uB

Point B is displaced according to


uB = (e12 x2 , e22 x2 )
And point C is displaced according to

x2
x1

uA

uC = (e11 x1 + e12 x2 , e21 x1 + e22 x2 )

This illustrates the way in which strain operates as a tensor. However, we also have the complication
that pure rotation can be defined also according to the same formalism:

Figure 3.9: System strained in a general way, with an


emphasis on the displacements of 3 points.

x2
x1
In the limit of a small rotation angle, we have
e11 = e22 = 0

e12 = e21 =

3.11. Relevant aside: Definition of the strain and elastic tensors

59

In the case that we combine pure rotations and pure shear stresses, we need a way to separate the two.
The way to do this is illustrated in Figure 3.10. It can be seen that we can define two combinations of
the stress components we have defined so far:

1
e + e ji
2 ij

1
ij =
e e ji
2 ij
eij =

(3.84)
(3.85)

The tensor components eij define the pure shear strain, and the components ij define the pure
rotation. Thus the fundamental materials tensor is e.
Elastic constant tensor
Stress and strain are coupled through the elastic constant tensor, defined as
= ce

(3.86)

This expands to
ij =

cijk` ek`

Figure 3.10: Mixture of shear strain and rotation, with


angles denoted by components of the shear stress
tensor.

(3.87)

k`

It should be clear that the elastic constant tensor is an example of a fourth-rank tensor.
The reverse relation defines the elastic compliance tensor
e = s

(3.88)

The harmonic strain energy is also given by the elastic constant tensor:
E=

1 T
1
e c e = cijk` eij ek`
2
2 ijk`

(3.89)

Rotation of axes
Whilst worrying about the orientation of axes something we humans determine might seem
unimportant, it turns out that our definition of shear strain and tensile strain also depends on our
arbitrary choice of axes. This point is illustrated in Figure 3.11. With the coordinate system x1 , x2 , x3 ,
the object is deformed by the shear strains e23 = e32 = e. However, viewed from the coordinate system

Figure 3.11: Demonstration that rotation of the system


of axes apparently converts between tensile and shear
strains.

Chapter 3. Landau theory

60

0 and e0 . Indeed, with a 45 rotation of the


x10 , x20 , x30 , the object is deformed by the tensile strains e22
33
coordinate system about x1 , the transformation is given as


0
0
0

1/ 2 1/ 2 0

1/ 2 1/ 2
0

e0 = 0
0

e
0

0
0
e

(3.90)

Remember that this is a tensor multiplication, not a matrix multiplication, as defined by equation 3.78.
Thus we have the following terms
0
e22

= a22 a23 e23 + a23 a22 e32 = e

0
e23
0
e32
0
e33

= a22 a33 e23 + a23 a32 e32 = 0


= a32 a23 e23 + a33 a22 e32 = 0
= a32 a33 e23 + a33 a32 e32 = e

(3.91)

giving

0
e = 0
0

0
e
0

0
e

(3.92)

This has demonstrated the conversion from shear strain to tensile strain via a rotation of the axes; the
point is that our definition of these two quantities depends on our viewpoint.

Voigt notation
By the time we get to fourth-rank tensors, carrying a large number of subscripts becomes cumbersome.
Worse, it masks the fact that symmetry can reduce the number of independent components of a tensor.
Consider the strain tensor. It is apparent from equation 3.84 that we have the relation that eij = e ji .
We can make use of this by writing the strain tensor with new indices:

e11

e = e21
e31

e12
e22
e32


e13
e1

e23 = e6
e33
e5

e6
e2
e4

e5

e4
e3

(3.93)

3.11. Relevant aside: Definition of the strain and elastic tensors

61

This would more commonly be written as a straight matrix equation:

1
2
3
4
5
6

c11
c21
c31
c41
c51
c61

c12
c22
c32
c42
c52
c62

c13
c23
c33
c43
c53
c63

c14
c24
c34
c44
c54
c64

c15
c25
c35
c45
c55
c65

c16
c26
c36
c46
c56
c66

1
2
3
4
5
6

(3.94)

By symmetry c ji = cij . Thus in the case of lowest symmetry, there are 21 independent component values
of the elastic constant tensor, a much smaller number that the total number of 34 = 81 components of
the full tensor.

Piezoelectricity
A material whose crystal structure does not contain a centre of symmetry11 will be susceptible to the
creation or a change of the dielectric polarisation by application of a stress. Similarly, application of
an electric field will generate an elastic strain. This property is called piezoelectricity. Since the electric
field and electric polarisation tensors are of rank 1, and the stress and strain tensors are of rank 2, the
coupling tensor, called the piezoelectric constant tensor must be of rank 3. Formally we write the two
equations as
P = d

(3.95)

e = dE

(3.96)

We introduce piezoelectricity here because in Chapter 1 we saw an example of how a piezoelectric


coefficient vanished at the phase transition due to the change in symmetry (we will better understand
the role of symmetry in determining the zero coefficients of a tensor in the next chapter), and the point
was that the value of the vanishing coefficient would have been proportional to 2 , thereby giving the
means to experimentally measure the order parameter indirectly.

11 The

symmetry by which any point in the structure with


coordinates x, y, z is also reflected at x, y, z.

Chapter 3. Landau theory

62

3.12 Phase transitions and elastic strain: case of quadratic coupling


Strain and the free energy
Now we come to the key application of the digression to consider the strain tensor. Many phase
transitions involve a change in lattice type we have met two different cubictetragonal phase transitions
in PbTiO3 (ferroelectric) and SrTiO3 (octahedral tilting displacive). In these cases, it is possible to define
a strain whose value falls to zero on heating up to the phase transition; this is called a spontaneous strain.

R: ELASTIC ANOMALIES IN SrTiO3

discussed
on of temcontext.

ntal rig to
[100] and
etragonal
ms to the
he critical
= Pbia into

F=

heim and
obtained
m Table 1
a slope of
pendence
y between

ted to be
value of
ature as a
screpancy
ns when a

1
1
1
1
a( T T0 ) 2 + b 4 + e 2 + ce2
2
4
2
2

(3.97)

First we minimise with respect to strain, and set the differential to zero to find the equilibrium value
of the spontaneous strain:


coth s2 .
Tc2

(53)

321

We can incorporate strain into the Landau theory developed in this chapter quite naturally, but we
will need to consider the place of symmetry. If we have a ferroelectric phase transition, the change in
symmetry includes the loss of a centre of symmetry. Elastic strain, however, will always preserve any
centre of symmetry, and thus the strain distortion through the cubictetragonal phase transition has a
higher symmetry than the fully symmetry change associated with the ferroelectric distortion. Similarly,
strain does not change the size of the unit cell, and thus the cubictetragonal phase transitions in SrTiO3
involves more symmetry change than is capture by the strain. As a consequence, we can add a new
term to the free energy that has the form e 2 , together with the elastic energy. For simplicity here, we
work with only one strain component. The new form of the Landau free energy is

Figure 3.12: Lattice parameters of SrTiO3 as a function


of temperature. The splitting of the values of the a
and c lattice parameters on cooling through the phase
transition is clear.

1
F
= 2 + ce = 0
e
2

1
e = c1 2
2

(3.98)

The key implication here is that the spontaneous strain follows the relationship e 2 ( Tc T )2
for a second-order phase transition. The example for SrTiO3 is shown in Figure 3.12.

3.12. Phase transitions and elastic strain: case of quadratic coupling

63

We now substitute this back into the Landau free energy function, Equation 3.97:
F

=
=
=
=

1
a( T T0 ) 2 +
2
1
a( T T0 ) 2 +
2
1
a( T T0 ) 2 +
2
1
a( T T0 ) 2 +
2





1 4 1
1 1 2
1
1 1 2 2
2
b
c + c
c
4
2
2
2
2
1 4 1 2 1 4 1 2 1 4
b c + c
4
4
8

1
1 2 1
b c
4
4
2
1 4
b
4

where

1
b = b 2 c1
2

(3.99)

(3.100)

The primary effect here is that the coefficient of the fourth-order term has been reduced in size. In fact,
we propose here (and demonstrate later) that this provides a mechanism for making a phase transition
first-order. The common terminology is to say that the coefficient of the fourth-order term has been
renormalised through the interaction with strain.
The value of the order parameter obtained by minimising the modified free energy function is
2 =

a
( Tc T )
b 2 c1 /2

(3.101)

Clearly the maximum value of is increased from what was possible without the coupling to strain.

Strain and susceptibility


We need to be very careful when we look at susceptibility, because we need to take account of the time
scales over which processes and measurements operate. Typically susceptibility is measured using a
time-varying applied field, which often varies over a time scale that is much shorter than the times
scales within which the elastic degrees of freedom can respond. Thus to calculate the susceptibility,
we need to use the equation developed without the strain coupling, equation 3.14, but then use the
equilibrium static value of the order parameter that is affected by the coupling to strain. Thus for T > Tc
where 2 = 0, we have
1

T > Tc = a ( T Tc )

(3.102)

Chapter 3. Landau theory

64
However, for T < Tc , equation 3.14 becomes
1

T < Tc

= a ( Tc T ) +
=

3ab
( Tc T )
b 2 c1 /2

2 + 2 c1 /2b
a ( Tc T )
1 2 c1 /2b

(3.103)

We have written it this way, in particular dividing both numerator and denominator by b, to highlight
the fact that the prefactor is necessarily larger than 2; recall that without strain coupling the prefactor
for T < Tc is exactly 2. This explains why the 2:1 ratio of prefactors in the measured susceptibilities is
so rarely observed in practice, and indeed why it can be broken so completely.

3.13 Phase transitions and elastic strain: case of linear coupling


Strain and the free energy
There are some phase transitions in which the change in symmetry associated with the order parameter
is identical to the change in symmetry caused by the accompanying strain. These will necessarily be
phase transitions in which the centre of symmetry is preserved through the phase transition, and where
the unit cell is not multiplied through the phase transition.
In this case, the term to be added to the free energy is of the form e, giving the result
F=

1
1
1
a( T T0 ) 2 + b 4 + e + ce2
2
4
2

(3.104)

Minimisation with respect to strain gives


F
= + ce = 0
e

e = c1

(3.105)

As before, we substitute this into the equation for the free energy, equation 3.104, to obtain
F

=
=

1
1
1
a( T T0 ) 2 + b 4 2 c1 2 + 2 c 2
2
4
2


1
2 c 1
1
a T T0
2 + b 4
2
a
4

(3.106)

3.13. Phase transitions and elastic strain: case of linear coupling

65

This shows that in the case of linear coupling, the effect is to increase the transition temperature in
the case of a second-order phase transition:


2 c 1
2 c 1
a T T0
a( T Tc )

Tc = T0 +
(3.107)
a
a

Susceptibility and elastic constants: ferroelastic phase transitions


Because the linear coupling to strain has led to renormalisation of the coefficients a or b in the Landau
free energy function, the susceptibility will not be changed other than by having it fall to zero at the
temperature of Tc instead of T0 (but note that linear coupling to strain does not also preclude the
possibility of quadratic coupling, so the analysis in the previous section will be applicable here too, but
separately so). However, we can expect to see a renormalisation of the elastic constants.
We can rearrange equation 3.105 as = ce/, and substitute into equation 3.106 to obtain
F

=
=

1 bc4 4
1
1 ac2
2
(
T

T
)
e
+
e ce2 + ce2
0
2 2
4 4
2


2 c 1
1 ac2
1 bc4 4
e
T

e2 +
0
2
2
a
4 4

1 ac2
1 bc4 4
2
e
(
T

T
)
e
+
c
2 2
4 4
1 2
+ e4
=
ce
2
The interesting quality is the renormalised elastic constant:

(3.108)

ac2
( T Tc )
(3.109)
2
which shows that the elastic constant falls to zero at the phase transition temperature.
This type of phase transition is called a ferroelastic phase transition, and has a close analogy with the
ferromagnetic and ferroelectric phase transitions we have met previously. The order parameter can be
identified by the spontaneous strain, which can be reversed by application of an external stress (strain
and stress are analogues of P and E for example). The stress/strain relationships can be described
as a hysteresis curve as we met for the other ferro-phase transitions. And the analogy of the inverse
susceptibility 1 falling to zero at the second-order phase transition is the elastic constant softening to
zero.
c =

Rev. Lett., 21, 16, 1968).

Cooling

Cooling

Wave vector
Fig. 12.19 Schematic representation of the
softening of the acoustic mode. The left-hand
figure shows the case where the acoustic mode
softens across the whole branch. The righthand figure shows the case where the softening is restricted to small wave vectors. In both
cases, the softening leads to a reduction in the
slope of the acoustic mode dispersion curve
in the limit k 0.
2.5

c44 (GPa)

2.0
1.5
1.0
0.5
0.0

750

850
900
Temperature (K)

950

Fig.3.14:
12.20
Temperature
dependence
the
Figure
Temperature
dependence
of the cof
44 elastic
elastic
constant
associated
with
the
soft c44
constant
of Na
CO
measured
by
inelastic
neutron
2
3
ferroelastic
phase transition
in Na
mea- at
scattering.
The continuous
decrease
to a2 CO
zero3 value
sured by inelastic neutron scattering. The data
the second-order phase transition is clear.
have been adapted from Harris et al., Phys.
Rev. Lett., 79, 4846, 1997.

The soft modes considered above have been optic modes, or of optic character.
Acoustic modes at k = 0 can also soften, but here the basic picture will be
somewhat
different since at k = 0 the frequency is already zero. In thisChapter
case, 3.
66
the softening involves the slope of the acoustic mode falling to zero, and this
corresponds to a softening of the elastic constant. Different ways in which this
Hexagonal
Monoclinic
can happen are shown in Fig. 12.19. The phase transitions caused by a softening
of an acoustic mode are the ferroelastic phase transitions discussed earlier in
this chapter. The gradient of the soft acoustic mode as k 0 gives the velocity
of sound, the square of which gives one of the elastic constants. The temperature
dependence of the soft c44 elastic constant in ferroelastic Na2 CO3 (see Section
12.1.2) has been determined from inelastic neutron scattering measurements of
the acoustic modes, and the results are shown in Fig. 12.20.

Landau theory

12.4.5 Incommensurate phase transitions


Most soft-mode
involve
modes
with
wave vectors
at k =
Figure 3.13: phase
Crystal transitions
structure of the
hexagonal
(left) and
monoclinic
(right) phases
of 0
Naor
2 CO3 . The structure
consists
of columns zone
of face-sharing
NaO6 octahedra
(Na it
areisthe
yellow spheres),
by the CO3 groups (C
at one of
the Brillouin
boundaries.
However,
possible
for thecross-linked
instability
and O are black and red spheres respectively), with the second Na atom lying in a large cavity. The phase transition
to occur
at a wave
vector
that ispositions
at a more
general
in the
zone.
involves
shifting
of the relative
of the
columnsposition
with rotations
are Brillouin
the CO3 groups.
A schematic representation of the modulation of a crystal due to a frozen-in
normal mode with a general wave vector is given in Fig. 12.21.
a ferroelastic
ThisExample
type of of
phase
transitionphase
is thetransition
incommensurate phase transition introNaChapter
transition
at 750 K. The structures
of the two phases are shown
duced in
2, andaisferroelastic
so calledphase
because
the wavelength
of the modulation
2 CO3 undergoes
in
Figure
3.13.
Inspection
of
the
structure
shows
that
the
phase
transition
involves
vertical displacements
associated with the soft mode is incommensurate with the underlying crystal
columns(in
of face-sharing
with
rotations
of the cross-bracing
CO3 groups. The shear
6 octahedra,
repeat of
distance
metals theNaO
wave
vector of
the
incommensurate
soft mode
the crystal is
fairlywave
obvious
from of
thethe
structure
it isSection
also apparent
thatOne
the shear strain does
will beofassociated
with
vector
Fermiimages;
surface,
5.4.7).
define all that is happening at the phase transition and is thus a useful quantity to associate with the
example of an incommensurate phase transition is quartz: the famous
order parameter.
phase transition
actually has an incommensurate phase existing over a small
The ferroelastic phase transition in Na2 CO3 is accompanied by a softening of the c44 shear elastic
temperature
range
between
two better-known
phases.
The mechanism
for in Figure 3.14.
constant,
which
has beenthe
measured
by inelastic neutron
scattering.
Results are shown
this involves
a softening
ofvalue
an optic
and
anatanti-crossing
interaction
with
It can be
seen that the
of c44mode
falls to
zero
the phase transition,
exactly
as our models have
an acoustic
mode (anti-crossing was described in Section 8.5.7). The effects
indicated.
of the anti-crossing
increase
with
k, since
thebeyond
optic and
acoustic
have we will outline
Incidentally, for
theoretical
reasons
that are
this module
(butmodes
using methods
briefly
later
in
a
later
lecture)
it
has
been
shown
that
ferroelastic
phase
transitions
should
different symmetry at k = 0, so the result is that the frequency of the acoustic obey mean-field
mode is driven to zero at a non-zero value of k. This is illustrated in Fig. 12.22.

3.14. Practical example: Landau theory for the ferroelectric phase transition in PbTiO3
Specific heat study and Landau analysis of PbTiO3 single crystals

3962

3957

67
G A Rossetti Jr and N Maffei

(a)

(b)

Figure 3.15: Heat capacity of PbTiO3 as a function


of temperature, showing the latent heat at the phase
transition.

Figure 5.3.16:
CurieWeiss
plot ofsusceptibility
the reciprocal relativeof
dielectric
permittivity
versus temperature for
Figure
Inverse
PbTiO
3 as a
PbTiO3 crystals.
function
of temperature. The line is provided to
show the extrapolation to the temperature T0 .

4. Conclusions

The specific heat of pure, float-zone-grown PbTiO3 crystals was measured at temperatures
between
325 and 1250 K. In the cubic phase above 800 K the average specific heat
theory, and hence Landau theory will be exactly
applicable.

S XS

C P (cubic) = (125 1) J mol1 K1 was found to be very close to the DulongPetit classical
limit. A Landau analysis of the excess specific heat subject to the assumption that the higher
order terms are temperature independent showed that the predictions of the mean-field theory
3.14 Practical example: Landau theory for
the
phase
transition
3t = 0.0130.049.
were
wellferroelectric
obeyed over the reduced
temperature
range of in
= PbTiO
(Tt T )/T
The expected proportionality between the spontaneous strain (c/a) 1 and excess entropy
over a larger
interval of transition
= 0.0060.11.atThe
first-order
S XS was maintained
PbTiO3 is the classical ferroelectric phase transition,
undergoing
a first-order
760
K. Thecharacter of the
Table 3.3: Data for the ferroelectric phase transition in
m3m 4mm phase change was described using a latent heat L = (1550 100) J mol1 ,
transition involves small displacements of the
cations
and
anions
direction,
causing
XS along the [001] 1
1
PbTiO3 .
an entropy discontinuity "S = (2.0 0.1) J mol K , a CurieWeiss constant Ccw =
K, and a temperature
difference (Tt T0 ) = (12ofthe
3) K with Curie
(1.3types.
0.1)
105showed
the transition from cubic to tetragonal lattice
We
the temperature-dependence
Figure 1. Plot of the excess specific heat (a) and excess enthalpy (b) ofWeiss
PbTiO3temperature
crystals as
T0 = (748 2) K. These parameters, together with the Landau coefficients Property
Value
functions
of temperature.
The inset in (a) shows
the first
derivative
of the excess specific
heat
polarisation
(proportional
to the
order
parameter)
in Figure
1.12; the jump in polarisation at the phase
and
the
magnitude of the specific heat, were found to be in good agreement with the values
versus temperature curve near the phase transition.
previously reported for micron-sized solgel derived polycrystalline PbTiO3 materials. These Tc
transition is given in Table 3.3.
760 K
results contrast with the large range of values for Ccw and (Tt T0 ) obtained from dielectric
T
725 K
The heat capacity is shown in Figure 3.15.measurements,
As a first-order
phase transition, it shows a latent heat as
data [14]:
where experimental difficulties may complicate the determination of the relative 0
$
! "
#2
XS measurements;
P
at
T
0.39 Cm2
a sharp
spike
in
the
the
integral
of
this
spike
gives
the
total
latent
heat,
as
given
in
Table
and
800
K.
permittivity
at
the
temperatures
of
interest
between
T
4
(T

T
3T
S
)
c
t
t
0
F(S XS )

1 +1 =
(2)
3
2L
(Tt T0 )
H at Tc
1550 J/mol
3.3.

where L is the latent


heatinverse
at the first-order
change and
!S XS is theAcknowledgments
corresponding
c11
The
of thephase
dielectric
susceptibility
is shown in Figure 3.16. Since it is a first-order phase
discontinuity in excess entropy, Tt !S XS = L.
XS
1
transition,
susceptibility
the
transition
temperature,
instead
undergoes
a jump Space Agency c12
The
authors
are grateful
to the US Officebut
of Naval
Research
and to the Canadian
A plot of F(S
) versus Tthe
gives
T0 at F(S XS ) =doesnt
0 and thediverge
slope is (Tat
t T
0 ) , from
XS for their support of this work.
obtained.
The
broad
temperature
range
over
which
H
varies
as
shown
which Tt may be
c44
in value. The extrapolation of the high-temperature data down to a zero value gives an estimate of the
figure 1(b) and the non-equilibrium nature of the scanning calorimetry experiment made the
Lattice
parameter at Tc
value
of
T
,
which,
from
the
data,
appears
to
lie
around
35
K
below
the
transition
temperature.
0
direct evaluation of L impractical.
Rather than evaluating L using arbitrary integration limits
References
on !C P (T ), which may result in a large error [14], we instead take L as the heat evolved
Extrapolated lattice parameter
above the temperature T1 . In the scanning experiment, this temperature defines[1]thePark
maximum
S E and Shrout T R 1997 J. Appl. Phys. 82 1804
[2] Shirane G, Pepinsky R and Frazer B C 1956 Acta Crystallogr. 9 131

c lattice parameter at 300 K


P at 300 K

at 300 K

230 GPa
140 GPa
80 GPa

3.96 A

3.94 A

4.16 A
0.75 Cm2

Chapter 3. Landau theory

68

We are now in a position to do some calculations. We write the Landau free energy function treating
the polarisation P as the order parameter:
F=

1 4 1 6
1
+ cP
a( T T0 ) P2 bP
2
4
6

(3.110)

where we have a minus sign on the quartic term because this is a first-order phase transition, and we use
the normalised b rather than the bare b as the coefficient. We note that for a first-order phase transition
the polarisation, as the order parameter, will have a jump in value described by equation 3.56 and as
given by the data in Table 3.3:


P2

T = Tc

3b
= 0.152 C2 /m4
4c

b
= 0.203 C2 /m4
c

(3.111)

Holding onto units is going to get complicated, particularly when things start to get squared, so we are
going to drop the units Cm2 from our equations this doesnt matter for this exercise provided we are
consistent as we go step by step. From equation 3.60 we can write the latent heat at the phase transition
as
aT0
H | T =Tc =
P2 = 1550 J/mol

a = 28.1 J/mol K1
(3.112)
2
where we have used the data reported in Table 3.3 for T0 . Using the two temperatures T0 and Tc
extracted from Figure 3.16 and also given in Table 3.3, we can use equation 3.57 to write
Tc T0 =

3b 2
= 35 K
16ac

b 2
= 5248 J/mol
c

(3.113)

(equation 3.111) we can obtain separate values for


Since we now have separate values for b 2 /c and b/c
b and c
b = 25.88 103 J/mol
c = 127.59 103 J/mol
(3.114)
We now need to consider the elasticity. We can obtained values for the elastic constants from
measurements of the acoustic mode dispersion curves, measured by inelastic neutron scattering. Lowfrequency dispersion curves for the cubic phase of PbTiO3 are shown in Figure 3.17. The elastic constants
can be obtained directly from the slopes of the acoustic modes; calculated values are given in Table 3.3.

3.14. Practical example: Landau theory for the ferroelectric phase transition in PbTiO3

69

Lattice dynamics of cubic PbTiO3 by INS

Downloaded By: [University of Oxford] At: 01:51 24 January 2010

(a)

Figure 1.

(b)

353

(c)

Low-frequency phonon dispersion curves in PbTiO3 at 775 K along several high-

Figure
3.17: Low-frequency
phonon
dispersion
curves
thethe
paraelectric
phase
of PbTiO
measured
inelastic
symmetry
cubic directions.
Full
symbols
standforfor
transverse
modes
from3 this
study.byOpen
neutron scattering. The blue and red lines show the acoustic modes (transverse and longitudinal respectively), and
squares longitudinal acoustic branch, stars and open triangles neutron scattering data
the green curve shows one of the transverse optic modes. The dashed curves show data obtained at a temperature
from Ref. [7]. Dashed and solid lines are guides for the eyes.
closer to the transition temperature.

encountered there simply because the structure factor for q > 0:4 r.l.u. rapidly
decreases towards the Brillouin zone boundary.
Data presented in this study differ from the earlier work in two aspects:
.

The TO dispersion reported in [7] is by about 2 meV lower than our data,
while the TA dispersion is somewhat higher than in our experiment.

Chapter 3. Landau theory

Lattice parameter ()

70

4.1

4.0

3.9

a,b
400

600
Temperature (K)

800

Figure 3.18: Lattice parameters of PbTiO3 as functions


of temperature. The splitting of the values of the a and
c lattice parameters on cooling through the phase transition is clear. The dashed line gives an extrapolation
of the data from the high-temperature cubic phase.

The elastic strains can be obtained from measurements of the lattice parameters, which are shown in
Figure 3.18. It is not surprising that the c lattice parameter shows a relatively large positive spontaneous
strain, given that all ionic displacements are along the z direction. Similarly, it is not surprising that the
spontaneous strain along the x and y axes, as shown by the change in the a lattice parameter, is small.
Because of this, we will only use the larger strain, e3 in the next part of the analysis.
We need to convert the elastic constants from units of energy/volume (GPa) to energy/mole. The
conversion is
c33 = c11 = 230 GPa a3 NA = 8.60 106 J/mol
(3.115)
We work now with the un-renormalised Landau free energy function, where we have a positive
value of b and we include the strain energy and the coupling between order parameter and strain
explicitly:
1
1
1
1
1
F = ( T T0 ) P2 + P4 + cP6 + 3 e3 P2 + c33 e32
(3.116)
2
4
6
2
2
From the earlier relationship linking strain to order parameter we have
1
1 2
e3 = 3 c33
P
2

3 = 2c33

e3
P2

(3.117)

Taking data at 300 K from Figures 1.12 and 3.18 as given in Table 3.3, and calculating the strain e3 from
the lattice parameters, we have
e3
3

= (4.16 3.94)/3.94 = 0.0558


0.059
= 2 8.7 106
= 1.708 106 J/mol
0.762

(3.118)
(3.119)

We are now in a position to calculate the bare value of b in the Landau free energy function. We need
to be careful at this point regarding the sign of the quartic term; in our un-renormalised Landau free
energy function (equation 3.116) the quartic term has a positive coefficient, but in our starting point for
this section where we used the standard renormalised function (equation 3.110) we followed convention
for first-order phase transitions and worked with a negative quartic term. If we retrace the steps that
led to equation 3.100 for the link between b and b, and keep proper track of the signs, we have
1
(1.708 106 )2
1
b = b + 23 c33
= 2.59 104
= 3.082 105 J/mol
2
2 8.6 106

(3.120)

3.15. Comments on Landau theory

71

This is a positive value, because the coupling term is very large, and this confirms the point we made
earlier that it is the coupling to strain that can make many phase transitions look like first-order

transitions. However, we note also that the value of b is more than an order of magnitude smaller
than b, meaning that that phase transition has only just been changed to first-order, as evidenced by the
relatively small size of P at the transition temperature compared to the value of P at 300 K.

3.15 Comments on Landau theory


Landau theory has been developed here as it is anywhere in a rather ad hoc method, thinking of how
the general form of the free energy has to look, then casting it into a simple Taylor expansion. The free
energy is an incredibly powerful tool, because if we have an analytical (or even numeric) form for the
free energy of any system we can calculate its equilibrium properties from the appropriate minimisation
condition and differentials. Hence the Landau free energy function has been used here to calculate the
order parameter, susceptibility, heat capacity and elastic strain. The important point about the Landau
free energy function is that it gets the symmetry aspects correct (well, it does if you dont mess up!),
which means that relationships between derived functions and temperature are always qualitatively
correct. And for some systems, they are quantitatively correct too!
There are three ways in which the Landau approach can be criticised. First, in what we have done
(which follows the traditional formulation) Landau theory no information about fluctuations, which
means that it is equivalent to a mean-field theory. For order-disorder phase transitions in particular,
this leads to significant errors at temperatures close to the phase transition, which is ironic because it
was close to the phase transition where Landau was hoping the theory would work best. But there are
systems, particularly many displacive phase transitions, where Landau theory works spectacularly well
for most of the temperature range. We have discussed this point above.
The second criticism is that the Landau free energy function is an expansion around the transition
temperature, which means that nothing has been built in to account for behaviour at temperatures close
to absolute zero, where the third law of thermodynamics comes into play. In particular, we expect many
functions to have a zero value of their derivative with respect to temperature close to absolute zero,
such as /T, but this is not captured in Landau theory. Hence the Landau free energy is not actually
a thermodynamically consistent function.
The third criticism is that it assumes that the entropy scales as 2 and that the internal energy

Chapter 3. Landau theory

72

12 One

might say that this is still a double-well function, with


an infinite rise in the enthalpy at the maximum value of
, but such a function has non-analytical derivatives at the
minimum.

(or enthalpy) is a double-welled function of . For many order-disorder phase transitions, including
ferromagnetic phase transitions, in a mean-field sense the internal energy will only scale as 2 up to
some maximum value of (fully ordered).12 And if you do an expansion of the entropy, which will
look something like (1 + ) ln(1 + ) + (1 ) ln(1 ) you find that the free energy contains terms of
the form T n , where n is the set of even numbers. Thus the quartic term in the expansion of the free
energy, and indeed all higher-order terms, have a prefactor proportion to temperature rather than being
independent of temperature.
But for displacive phase transitions, Landau theory works great, and is frequently the starting point
for gaining an understanding of a range of data obtained for a phase transition.

3.16 Summary
1. The free energy can be written as a power series of the order parameter , with terms that reflect
the symmetry changes associated with the phase transition. This can be decomposed into enthalpic
and entropic terms, with both written in terms of the order parameter.
2. The equilibrium state is that for which the free energy has minimum value. This condition can be
used to find the equilibrium value of the order parameter in the Landau free energy function.
3. From the free energy we can obtain equations for the temperature dependence of the order
parameter, heat capacity and susceptibility at temperatures around a second-oder phase transition,
which can be written as a power of the temperature difference | T Tc |, where the exponent is
called the critical exponent.
4. Landau theory for a second-order phase transition predicts values for the critical exponents of
= 0 for the heat capacity, = 1/2 for the order parameter and = 1 for the susceptibility.
5. Landau theory predicts that the slope of the susceptibility with temperature is twice as large for
T < Tc that for T > Tc .
6. At the transition temperature for a second-order phase transition, the variation of the order
parameter with applied conjugate field has a power-law relation, 1/ , where Landau
theory predicts = 3.

3.16. Summary

73

7. Landau theory can be extended to include spatial variation of the order parameter, which enables
the theory to describe fluctuations. These have a size characterised by the correlation length , with
temperature dependence | T Tc | . Landau theory predicts = 2. At T = Tc fluctuations
decay as r2d , where = 0 in Landau theory.
8. Landau theory can be applied to first-order phase transitions with a negative quartic term, requiring the inclusion of the sixth-order term to ensure that the free energy increases at large values of
the order parameter. We have derived the temperature-dependence of the order parameter, heat
capacity and susceptibility on heating through a first-order phase transition.
9. When symmetry permits odd-order terms to exist in the Landau free energy function, the
transition is necessarily first order. There is no jump in the value of the susceptibility at the
transition temperature.
10. Tricritical phase transitions are defined as the intermediate case between first and second order
transitions. Within Landau theory this case can be represented by a zero coefficient of the quartic
term. In this case, critical exponents from Landau theory are = 1/2, = 1/4 and = 1.
11. Stress and strain are defined as two-dimensional tensors, written as 3 3 symmetric matrices.
12. Depending on symmetry, the Landau free energy can be expanded to include terms that couple
strain to the order parameter, together with an elastic energy term. Whether the coupling term
includes strain to linear or quadratic order depends on whether the strain has the same or different
symmetry to that of the order parameter.
13. Minimisation of the free energy with respect to both the strain and order parameter yields the
relationship between the two; in the two cases, strain varies either linearly or quadratically with
the order parameter.
14. In the case that strain has the same symmetry as the order parameter, substitution of the strain in
the free energy modifies the quadratic term and acts to increase the transition temperature.
15. In the case that strain has a different symmetry as the order parameter, substitution of the strain
in the free energy modifies the quartic term and acts to decrease the value of the coefficient. If the
effect is large enough it can make the quartic term negative, changing the transition from second
to first order.

Chapter 3. Landau theory

74

16. Coupling between the strain and order parameter leads to the slopes of the susceptibility curves
differing from the 2:1 ratio either side of the transition temperature.

3.17 Appendix: Correlation functions


Simple spin model
Consider a snapshot of a one-dimensional ferromagnet in its paramagnetic phase:

r
where the atomic magnetic moments which we will call spins are denoted by the arrows and can
point up or down, and where we denote cluster boundaries by the vertical dashed lines. In this figure r
is the spatial coordinate. Remember that in one dimension we do not actually get a phase transition
at a non-zero temperature, but this is not important for this analysis. As you move along the crystal
you encounter a small cluster where the spin points in one direction, followed by a cluster where the
spins point in the opposite direction, and back again. We can denote the density of cluster boundaries,
namely the number of cluster boundaries per unit length, by c , and the distance between spins as a, so
that the average size of a cluster is 1/c and the number of spins in a cluster is equal to 1l/ac .
The information that a correlation function tells us concerns the way that on average any spin
sees the local ordering of its neighbours. Specifically here, the correlation function will tell us about
the way in which the local magnetisation changes in space. If we have complete disorder, either
because we are at very high temperature or because there is no coupling between the spins, there
will be no correlation between neighbouring spins and so the average moment at any distance away
from one spin will be zero. However, spins do interact and temperature is not infinite, so there is a
force that favours one spin aligning with its neighbours, but above any phase transition this transition
does not propagate for ever. Therefore we can say that at long distances the correlation between any
one spin and any spin a long way away will be zero. That is, if we denote a spin at a distance r
as S(r ), we expect that hS(0) S(r )i = 0, where the average is over all origin points. On the
other hand, we have hS(0) S(0)i = S2 . With a given density of cluster walls, in a chain of N spins
there will be Nac cluster walls. Each of these has a pair of spins of opposite orientation, whereas the

3.17. Appendix: Correlation functions

75

remaining N Nac pairs have the same orientation. Therefore the sum of all products of neighbouring
spins will be ( N Nac )S2 Nac S2 = N (1 2ac )S2 . Normalising by the number of spins we have
hS(0) S( a)i = (1 2ac )S2 . We can write hS(0) S(r )i = g(r )S2 , where g(r ) is called a spatial correlation
function. From what we have written, we have values g(0) = 1, g( a) = 1 2ac , and g() = 0.
We can extend this analysis, and consider the crystal to be a continuum (that is, we assume 1/c > a).
In such a case, the probability that there is a cluster wall over a small spatial step dr is c dr. This is
equivalent to the probability of a flip in the sign of the magnetisation over this step. The probability
that at any point r the magnetisation is positive is P+ (r ) and the corresponding probability that the
magnetisation is negative is P (r ) = 1 P+ (r ). The change in P+ (r ) over the step dr is equal to

dP+ (r ) = P+ (r )c dr + P (r )c dr = 1 2P+ (r ) c dr

(3.121)

where the first term is the probability of the flip into the negative direction if the spin is already in the
positive direction (hence the negative sign) and the second sign is the probability of the flip from the
negative to positive direction. Hence we can write

dP+ (r )
= c 1 2P+ (r )
dr

(3.122)

A similar equation exists for P (r ). The solution has the form


P+ (r ) = A + B exp(r/ )

(3.123)

The boundary conditions for starting with a positive magnetisation are P+ (0) = 1, P+ () = 1/2, which
imply that A = B = 1/2. By comparing the differential
B
A P + (r )
dP+ (r )
= exp(r/ ) =
dr

(3.124)

with equation 3.122 we also have = 1/2c


We form the correlation function from the relationship

hS(0) S(r )i = S2 P+ (r ) P (r )


= S2 21 (1 + exp(r/ )) 12 (1 exp(r/ ))

= S2 exp(r/ )

(3.125)

Chapter 3. Landau theory

76
Thus
g(r ) = hS(0) S(r )i /S2 = exp(r/ )

(3.126)

This result is independent of whether we started with a positive or negative magnetisation.


is called the correlation length, and gives the length scale over which magnetisation in two places
of the crystal are correlated. Note that it is given by the inverse of the density of cluster walls; larger
clusters associated with larger values of correspond to a smaller density of cluster walls.
We should make one point here about terminology. Technically g(r ) is called an autocorrelation
function, because it is constructed as the correlation of one function S(r ) with itself. In the general
case we could have built the correlation function from two different functions to give information
about their correlation over distance. Secondly, we note that it is more often the case that correlation
functions are defined in terms of time rather than position, which in science takes us into the realm of
spectroscopy, but more generally time correlation functions are used in the study of spectra of many
different types of signal in, for example, the areas of engineering and medicine.

Correlation functions and Fourier transforms


If we imagine that our crystal, and hence our function S(r ), extends a long way in space (practically
to infinity), the mathematical operation to obtain the correlation function g(r ) can be written in the
discrete spin case as
1
g (r ) =
S(r i ) S(r i + r )
(3.127)
NS2
i
where N is the number of spins, and where r is restricted to the set na with n as an integer. Written as a
continuum equation, and thereby replacing the sum by an integral, we have (without being concerned
with normalisation)
Z
1 +
g (r ) = 2
S(r 0 ) S(r 0 + r )dr 0
(3.128)
S
Given that we think of the spins as discrete objects, we could generalise by considering that the order
parameter varies with space and can be locally defined. Thus the correlation function can instead be
written as
Z
g (r ) =

(r 0 ) (r 0 + r )dr 0

(3.129)

3.17. Appendix: Correlation functions

77

Let us take a general case of a correlation function involving two functions:


u (r ) =

Z +

v(r 0 )w(r 0 + r )dr 0

(3.130)

The Fourier transform is defined as


U (k)

=
=

ZZ

tv(r 0 )w(r 0 + r ) exp(ikr )dr 0 dr



Z
Z
w(r 0 + r ) exp(ikr )dr dr 0
v (r 0 )

(3.131)

We make a change in variable, z = r 0 + r, which means dz = dr, and so we can write



Z
Z
0
0
w(z) exp(ik (z r )dz dr 0
U (k) =
v (r )

v(r 0 ) exp(ikr 0 )dr 0

= V ? (k) W (k)

w(z) exp(ikz)dz
(3.132)

In the case of the autocorrelation function, writing G (k ) as the Fourier transform of g(r ), and S as the
Fourier transform of S(r ), we have
G (k ) = | (k)|2

(3.133)

The case G (k = 0) corresponds to the susceptibility. On cooling towards the phase transition, the
correlation length diverges, which means that G (k) becomes narrower, eventually becoming a deltafunction at the actual transition temperature.
The spin model has very sharp boundaries between clusters, but we could imagine cases where the
variation of (r ) is much smoother than for the case of spins with two states. For example, we could
imagine magnetic cluster boundaries where the spins can rotate gradually from one state to another, or
a displacive phase transition where the change in atomic displacements from one cluster to another
occurs over a number of unit cells. In this case, we can imagine that the function (r ) is similar to the
case of a line of spins except that it is convolved with a function that describes the gradual changes at
the cluster wall. For example, if this function is a box function, (r ) will vary linearly at the cluster wall.
Alternatively the function could look more like a Gaussian, giving a smoother variation of (r ) at the
cluster wall. Either way, provided that the size of the cluster boundary is smaller than , the Fourier

Chapter 3. Landau theory

78

transform of g(r ) will by multiplied by the square of the Fourier transform of the function defining the
cluster boundary.13 So when we do the reverse transform, we have that g(r ) is now convolved with the
the function defining the boundary convolved with itself. Because this function is rather sharper than
g(r ), the effect is most likely to be seen only where g(r ) changes most rapidly, namely for small r. Thus
we expect that at small r we will find g(r ) to be slightly rounded, with a flat top instead of a cusp at
r = 0.

3.18 Short questions to test your basic knowledge


1. Under which condition must the Landau free energy function contain only terms of even power
in the order parameter?
2. Show that for a Landau free energy function with positive quadratic and quartic coefficients, the
order parameter varies as ( Tc T )1/2 .
3. Derive expressions for the susceptibility and changes in heat capacity associated with a secondorder phase transition.
4. Derive expressions for the susceptibility and changes in heat capacity associated with a tricritical
phase transition.
5. Define the correlation length and state how it varies with temperature on approaching a phase
transition.
6. Define the critical exponents , , , and .
7. How is the Landau free energy function for a second-order phase transition modified to describe
a first-order phase transition?
8. Explain the circumstances, giving examples, for the occurrence of terms of the form e or e 2 in
the Landau free energy function. What other term must be added to the function to account for
the effects of strain on the free energy?
9. What effect does the case of a coupling term of the form e have on the transition temperature?

13 This

follows
Fourier trans
their separat
identical to t

3.19. Longer questions

79

10. What effect does the case of a coupling term of the form e 2 have on the coefficients of the various
terms in the free energy?
11. Explain how coupling to strain can change a phase transition from second to first order.

3.19 Longer questions


1. Some crystals show normal thermal expansion along one axis and negative thermal expansion
along an orthogonal axis; this is often the case for crystals that appear to have strongly hinged
networks of groups of atoms, as in the example in Figure 3.19. In this case the coefficients of
thermal expansion are 1 = 2 = +1.5 104 K1 , and 3 = 1.2 104 K1 .14
(a) Draw a slab of a crystal at some orientation to the x-z axes, denoting the angle between the
normal to the slab and the z axis as . Define a new coordinate system x 0 -z0 such that z0 is
normal to the slab.
(b) Write down the transformation matrix a in terms of .
(c) Write equations for the coefficients ij0 remembering now to transform from Voigt notation
into the full tensor notation for i, j = 1, 3 in terms of the original coefficients 11 and 33 ,
noting that by definition 13 = 31 = 0.
(d) Suppose we want the crystal cut into a slab in such a way that thermal expansion in the
0 = 0. Calculate the value of that will
direction normal to the slab face is zero, namely 33
achieve this.
2. Derive an equation for the temperature at which the the free energy for a first-order phase
transition as given by equation 3.53 first develops a minimum at a non-zero value of on cooling
from high temperature.
3. Show that there is no jump in the susceptibility at Tc for a phase transition that is described by
equation 3.64.
4. Show that the upper critical dimension for a tricritical phase transition is 3. This involves
explaining the value of the critical dimension for the case of the tricritical phase transition within
the framework of Landau theory.

Figure 3.19: Crystal structure of Ag3 Co(CN)6 , showing


CoC6 octahedra, and NAgN linkages between the
octahedra to create a hinged structure.

14 Note

that these are possibly record values, and you can read
about this system in Science, volume 319, pp 794797, 2008.

Chapter 3. Landau theory

80
5. Consider the spin-1 Ising model, which has the Hamiltonian
nn

H = J Si S j Si2 B Si
i,j

where the spins S can take the values 1, 0, 1 and the sum labelled nn is over all pairs of nearest
neighbours i and j. Develop this model using the mean-field approximation. Show that for certain
values of this system undergoes a tricritical phase transition, and find those values. Find the
temperature of the tricritical transition in terms of . Give a physical explanation for why this
model should have a tricritical phase transition.
Comments and hints There are several possible physical interpretations of this model. One is, as the
name suggests, a spin-1 magnet in which there is a preferred direction of magnetisation. If > 0 then
S = 1 is energetically favourable and magnetisation occurs preferentially in the z direction; if < 0 then
S = 0 is favoured by this term so magnetisation occurs preferentially in the xy plane. A second interpretation
is a spin- 12 magnet where S = 1 represents spin-up and spin-down orientations and S = 0 represents
defects; thus in this interpretation is the energetic cost of creating a defect.
There are also several possible ways of approaching this problem. One would be to follow the same route we
used to derive the Curie-Weiss model in section 2.4: initially set J = 0 and calculate the average value of the
order parameter hSi, then replace B by B + J hSi and solve graphically in the same way. Another would be,
as in section 3.7, to calculate the partition function and hence the free energy from the Hamiltonian, and
compare this to Landau theory expansions.
All you need to do for an 100%-correct answer is to find the tricritical value of in one of these ways, and to
explain it in terms of one of these interpretations. But I would encourage you to play with different values of
the parameters and ways of approaching and interpreting the problem: try it in different ways and see how
the model behaves in different circumstances, and write down what you discover for bonus points!
You should feel free to make use of appropriate software (Mathematica, Matlab, python. . . ) in your solution
to avoid tedious algebra.

Chapter 4

Symmetry

4.1 Introduction to symmetry


We have already seen that there is an intimate connection between phase transitions and the symmetry
of materials: specifically, that many phase transitions arise from symmetry breaking. This chapter
deals in more detail with the formal means of describing materials symmetry, and in particular how
understanding it can help us to construct models of their phase transitions.
Formally speaking, a symmetry of a function defined on some space is a transformation of that space
that leaves the function invariant. To take a simple example, f ( x ) = cos( x ) is a function defined from R
to [1, 1]. Suppose we define the transformations
T1 ( x ) = x

(4.1)

T2 ( x ) = x + 2;
(4.2)


then we find that, for all x, f ( x ) = f T1 ( x ) and f ( x ) = f T2 ( x ) . Thus T1 and T2 are both symmetries
of f .
For an example more relevant to condensed matter physics, consider the structure shown in
figure 4.1. This structure would look identical if we rotated it by 90 about the z axis; in other words,
fourfold rotation is a symmetry operation of this material. In the same way as before, this is represented
mathematically by the transformation
T ( x, y, z) = (y, x, z)

(4.3)

(check for yourself that this represents a 90 rotation about the z axis). We will return to this example
throughout this chapter.
Recall that any crystal structure can be described in terms of the appropriate three-dimensional
Bravais lattice. It turns out that only a finite number of symmetry operations are compatible with the
81

Figure 4.1: A tetragonal perovskite-like system.

Chapter 4. Symmetry

82

possible lattices. These are probably familiar to you, but for completeness we will now briefly consider
each of these.

4.2

1 For

an example of a symmetry that does not obey this criterion, consider translation.

Point group symmetry operations


There are five possible point symmetry operations point operations because each of them leaves at
least one point of three-dimensional space fixed.1 Unfortunately there are two separate sets of notation
for these: we will use the Hermann-Mauguin notation, also called international notation, because it is better
suited to describe crystal structures. Be aware, though, that an alternative, Schonflies notation, is also
very common, particularly in the molecular physics and spectroscopy communities.
One way of visualising these operations is to draw a sphere diagram. The idea is to show how an
arbitrary point on the surface of a sphere would be transformed by the relevant symmetry operation.
We draw this in two dimensions by projecting these points onto a plane passing through the centre of
the sphere. The distinction between points above and below this plane is maintained by showing one as
a filled circle, the other as an empty circle (Figure 4.2).

Identity (1)
Figure 4.2: Construction of a sphere diagram showing
a centre of inversion. The two red atoms on the
surface of the sphere are related by inversion. Their
positions are projected onto the grey plane: the one
above the plane is shown as a filled circle, the one
below as an empty circle.

The identity is the symmetry operation that does nothing at all, transforming
( x, y, z) into ( x, y, z). This is an obvious symmetry of every single structure,
but we include it for completeness and because it is needed as part of the
formal mathematical description of a group (which we will discuss in the next
section).

Centre of inversion (centre of symmetry) (1)


A centre of inversion maps every point to the corresponding point through
the centre and the same distance out the other side. The simplest example
is a centre at (0, 0, 0), which maps ( x, y, z) into ( x, y, z). Indeed, the
convention is to put the origin of our cell axes at a centre of inversion wherever
this is possible.

4.2. Point group symmetry operations

83

Mirror plane (m)


A mirror plane performs the familiar operation of reflection. For instance, a
mirror plane in the xz plane reflects ( x, y, z) into ( x, y, z). The diagram on
the right shows a mirror plane in the plane of the page.

Rotation axis (2, 3, 4, 6)


A rotation axis of order n spins the structure around 360 /n, so that n turns
are needed to return to the starting point. Only axes of order 2, 3, 4, and 6
are compatible with lattice symmetry.2 For an algebraic example consider a
twofold axis along the y axis, which takes ( x, y, z) to ( x, y, z). Shown on
the right is a fourfold axis.
6)
4,
3
Rotoinversion axis (3,
A rotoinversion axis of order n, just like a rotation axis, spins the structure
around 360 /n; but this is then followed by an inversion. It is possible to have
a rotoinversion axis when neither an inversion centre nor a rotation axis of

the same order are present: for an example consider the sphere diagram for 4,
shown to the right. We dont include 2 in this list since this is equivalent to a
mirror plane m. Again, if you cant see this immediately a sphere diagram is
very helpful. For an algebraic example, a 4 axis along the y axis takes ( x, y, z)
to (y, x, z).
Put together, these operations give a choice of 32 possible point groups, or collections of symmetry
operations, consistent with a crystallographic lattice. These are conventionally divided into seven crystal
systems (Table 4.1), which have symmetries compatible with the corresponding lattice system (Table 4.2).4
Reading the point group notation can take a little while to get used to. One important thing to remember
is that the notation / (usually pronounced upon) means perpendicular to, so that 4/m is a fourfold
rotation axis perpendicular to a mirror plane, while 4mm is a fourfold rotation axis lying in two mirror
planes (i.e., the axis is located at the intersection of the mirror planes).
The point group symmetry of the phases involved in a phase transition can tell us a lot about
the nature of the transition. Two particularly important examples are that ferroelasticity arises from a

2 This

is why the discovery of quasicrystals, which have 5fold symmetry in their diffraction patterns, caused such a
controversy and was later responsible for Dan Shechtman
being awarded the 2011 Nobel Prize!

3 Warning

for those used to Schonflies


notation: this is not the

same operation as improper rotation in that system, which


is rotoreflection rather than rotoinversion. Thus, for instance, a
3 axis is the same as an S6 axis, not an S3 .

4 The

observant reader will notice that the seven crystal systems are not quite the same as the seven lattice systems.
Specifically, a hexagonal lattice is compatible with either
trigonal or hexagonal symmetry; while trigonal symmetry
can correspond to either a hexagonal or a rhombohedral lattice. This distinction, however, will not be important in this
module.

Chapter 4. Symmetry

84

Table 4.1: The 32 crystallographic point groups, divided into the seven crystal systems. The 11 centrosymmetric
groups, also called Laue classes, are shown in boxes; the 10 polar groups are circled.

Triclinic

none

Monoclinic

one 2-fold axis

Orthorhombic

three 2-fold axes

Trigonal

one 3-fold axis

Tetragonal

Hexagonal

one 4-fold axis

one 6-fold axis

222

mm2

32

3m

for transitions between trigonal and hexagonal point


groups. Again, this will not be important in this module!

4.3

only patterns of infinite extent can have such


symmetries.

3m
4

4mm

622

6mm

4/m

4m2

4/mmm

6/m

6m2

6/mmm

23 m3

432 43m
m3m

transition between two different crystal systems,5 and that ferroelectricity arises from a transition to a
polar point group that can support a net polarisation of the unit cell.

5 Except

6 Obviously

four 3-fold axes

mmm

422

(
Cubic

2/m

Space group symmetry operations


When we consider ideal crystal structures, which repeat over and over in three dimensions, there are
further symmetry operations to consider. We now refer to space symmetry because no point need be
fixed by these operations.6
Most obviously, translation by any lattice vector is now a symmetry operation. In addition, for
centred lattices (see Table 4.2), translation by a centring vector is possible. For example, in the body-

4.3. Space group symmetry operations

85

Table 4.2: The 14 Bravais lattices, divided into the seven lattice systems.

simple

base-centred

body-centred

face-centred

Triclinic
Monoclinic

= = 90

Orthorhombic

= = = 90

Rhombohedral

a = b = c, = =

Tetragonal

a = b, = = = 90

Hexagonal

a = b, = = 90 , = 120

Cubic

a = b = c, = = = 90

centred (I) cubic phase we can translate by any linear combination of the three lattice vectors (1, 0, 0),
(0, 1, 0), and (0, 0, 1) and the centring vector ( 12 , 12 , 21 ). In fact, there is no mathematical difference
between lattice and centring vectors; this distinction is purely for our own convenience. We can
always express a centred cell in terms of primitive lattice vectors in the I cubic case these are
 1 1 1

( 2 , 2 , 2 ), ( 12 , 21 , 12 ), ( 21 , 12 , 21 ) in which case there is no need to specify a separate centring
vector.7
To draw space group operations, we use a similar device to the sphere diagrams we used for point
group operations. This time we just use a single type of atom, conventionally drawn as an open circle.
We project the atoms and unit cell onto the plane of the paper and label each site with its z coordinate.
In particular, we often use the two special symbols + and , which mean a small, unspecified distance

7 Exercise:

check that you can generate all four of the vectors


mentioned above from linear combinations of these three
primitive lattice vectors.

Chapter 4. Symmetry

86
above and below the plane respectively.

For 73 space groups, called symmorphic groups, all the possible operations are generated by combinations of the point group operations listed in the previous section and translations. However, there
are also 157 non-symmorphic groups in which two further symmetry operations may feature. What these
have in common is that, while repeated application of any point group operation will return you to the
identity, repeated application of these operations leads eventually to a pure, non-zero translation.

Screw axis (21 , 31 , 32 , 41 , 42 , 43 , 61 , 62 , 63 , 64 , 65 )


A screw axis is the combination of a rotation and a translation parallel to the
axis of rotation. The notation nm means a rotation of 360 /n followed by a
translation of m/n; thus 31 is a rotation of 120 followed by a translation of 13
of a cell vector, and (31 )3 is a translation by one unit cell in the direction of
the rotation axis.
The diagram on the right shows a 21 axis in the horizontal direction.

Glide plane (a, b, c, d, e, n)


Similarly, a glide plane is the combination of a reflection and a translation
parallel to the plane of reflection. The notation refers to the direction of
translation: a, b, c respectively mean translations in the directions of the three
unit cell vectors, and d, e, n refer to other possibilities which wont concern us
here.
The diagram on the right shows an a glide plane in the plane of the paper.

Putting these all together gives the 230 space groups. Each of these is represented in HaumannMaugin notation by a capital letter referring to the lattice type (Table 4.2) followed by a symbol that
looks very similar to the symbol for the point group of the crystal (Table 4.1), except that the space
group symmetry operations above may replace their point group equivalents. So, for instance, I41 /a is
a space group that we shall encounter later in this chapter: it is based on the tetragonal point group
4/m and a body-centred (I) tetragonal lattice.

4.4. Groups and their representations

87

4.4 Groups and their representations

Table 4.3: A representation of the point group 4/mmm.

At this stage we will pause to justify our use of the word group. We mean this in the strict mathematical
sense, which describes a set G obeying the following axioms:

1 !
1 0
0 1

1. We can multiply any two elements g, h of the set to form another element, which we write gh.

2x
1
0

2. Multiplication is associative: for all g, h, k G, ( gh)k = g(hk ).

0
1

1
0

4. Each element g G has an inverse, written g1 , such that gg1 = g1 g = e.

1
0

0
1

0
1

mx !
1 0
0
1

1
0

1
0

0
1

1
0

2x y !
0
1
1
0

mz !
1 0
0 1

4 3 !
0 1
1
0

m x +y !
0
1
1
0

m x y !
0 1
1 0

my !
0
1

0
1

2x +y !
0 1
1 0

4
!

43

2z

2y !
1 0
0
1

3. There is an identity element e G such that, for all g G, eg = ge = g.

A little thought shows that the symmetries of any object will indeed obey these axioms. Following
axiom (1), we can define the product of two symmetries as the composition of the associated transforms:
that is, we simply perform the first symmetry operation, then the second. It is then easy to check that
axiom (2) holds. The identity of the group is simply the identity operation 1, which leaves every point
in place, satisfying axiom (3). Finally, it is also easy to check that every operation can be reversed to
give its inverse, satisfying axiom (4): for instance, the inverse of the operation 4, rotation by 90 , is 43 ,
rotation by 270 to arrive back where we started.
As an example, lets consider the point group of the tetragonal structure shown in figure 4.1. Listing
the symmetries carefully, in the same order as we discussed in the previous section, we have: the identity
a mirror plane m perpendicular to z; mirror planes m perpendicular to x
1; a centre of symmetry 1;
and y; mirror planes m perpendicular to x + y and x y; a fourfold axis 4 in the z direction; a twofold
axis 2 in the z direction; twofold axes 2 in the x and y directions; twofold axes 2 in the x + y and x y
directions; and a fourfold rotoinversion axis 4 in the z direction, for a total of 16 symmetry operations.
Check that you can see all of these in the figure.8 These operations comprise the point group 4/mmm.
We define a subgroup, not surprisingly, as a subset of a group that itself obeys the group axioms. In
fact most of them will trivially continue to be true because the parent group obeys them; the main thing
to check is that the subset is closed under the group multiplication operation. Formally, H G is a
subgroup if and only if for any g, h H, gh H; this is often written H < G. The key significance of
this definition is that very often there is a group-subgroup relationship between the space groups of
the structures on either side of a phase transition. Indeed, this is sometimes called the first criterion of
Landau theory.

4 !
0 1
1
0

Table 4.4: A character of the point group 4/mmm.

8 Of

1
2

4
0

2z
2

43
0

2x
0

2y
0

2x +y
0

2x y
0

1
2

4
0

mz
2

4 3
0

mx
0

my
0

m x +y
0

m x y
0

course, some of these symmetry operations are necessarily


present because of others: obviously any 4 axis will coincide
with a 2 axis, because 42 = 2 (an apparently bizarre consequence of our notation!). In fact all you need to generate this
group is the fourfold axis, the inversion centre, and either
one mirror plane parallel to the 4 axis or one twofold axis
perpendicular to it. Exercise for the keen: demonstrate this.

Chapter 4. Symmetry

88

Now we can apply group symmetry operations to any function defined on R3 , not just our crystal
structure. As an example, lets apply these operations to an arbitrary point ( x, y) in the xy plane. For
instance, applying the symmetry operation 4 to this point gives (y, x ): written as a matrix equation
this becomes
!
!
!
0 1
x
y
=
(4.4)
1 0
y
x
showing that we can in some sense represent the symmetry operation 4 by the 2 2 matrix on the left of
the equation. We can continue this process for each element of the group, with results shown in Table 4.3.
This gives a representation of our point group: a set of matrices that obey the same multiplication rules
as the group itself. In fact, since these are 2 2 matrices, we say that this is a representation of degree 2.
We denote the matrix corresponding to some group element g by ( g).
The subject of representation theory is a branch of mathematics in its own right, and we will
unfortunately only have time in this module to quote a few key results.
We first note that, given a representation and any invertible matrix T, we can construct another
representation according to
( g) = T 1 ( g) T.

(4.5)

We know is also a representation, because for any group elements g1 , g2 ,


( g1 ) ( g2 ) = T 1 ( g1 ) TT 1 ( g2 ) T

= T 1 ( g1 ) ( g2 ) T
=T

( g1 g2 ) T = ( g1 g2 ) .

(4.6)
(4.7)
(4.8)

We call and equivalent representations.


Now consider again the representation in Table 4.3. If we wanted to, we could have defined a
representation 0 of degree 3 that included the effect of each operation on the z component of a position
vector: so, for instance, the operation 4 would have matrix

0
(4) = 1
0

1 0

0 0 .
0 1

(4.9)

4.4. Groups and their representations

89

This is a perfectly valid representation. It turns out, however, that every matrix in this representation
has the form

(4.10)
0 ;
0 0
thus we say that this representation is reducible into the second-degree representation , as defined in
Table 4.3, and a first-degree representation that we can call . We write
0 =

(4.11)

where the operation is called the direct sum. It is reasonably obvious that the direct sum of representations will always give another representation; for example, is a representation of
degree 7.
In fact it turns out that almost all the information from any given representation is contained in the
trace of the matrices; that is, the sum of their diagonal elements. These are so important that they are
given the special name of characters. The character () of is given in Table 4.4. It is not too difficult to
show that equivalent representations have the same character, and that for any two representations ,
( ) = ( ) + ( );

(4.12)

that is, the character of a direct sum of representations is the sum of the corresponding characters.
Another interesting observation from Table 4.4 is that symmetry operations which are in some sense
similar to one another have the same character: for instance, (4) = (43 ) and (m x ) = (my ). To
formalise this similarity, we say that two elements g1 , g2 of a group are conjugate if, for some element
h of the group,
h 1 g1 h = g2 .
(4.13)
The set of all elements conjugate to some element g is called the conjugacy class of g. Then our
observation, which is true in general, becomes that conjugate elements have the same characters, or
equivalently that characters are constant on conjugacy classes.
Finally, we simply state without proof the important result that the number of irreducible characters9
of a group is the same as the number of conjugacy classes. Putting all of this together, we have the
concept of a character table of a group, which is a square matrix in which the characters of the group
appear in rows. The characters are listed down the left-hand side (typically with labels referring to

9 That

is, characters of irreducible representations.

Chapter 4. Symmetry

90

Table 4.5: The character table for the point group 4/mmm.

10 Note

the second row, which shows the multiplicity, or


number of elements in each conjugacy class. This is not
formally part of the character table but it is so useful that it
is often included.

4/mmm

2z

2x

2x +y

mz

mx

m x +y

mult.

A1g
A2g
B1g
B2g
Eg
A1u
A2u
B1u
B2u
Eu

1
1
1
1
2
1
1
1
1
2

1
1
1
1
2
1
1
1
1
2

1
1
1
1
0
1
1
1
1
0

1
1
1
1
0
1
1
1
1
0

1
1
1
1
0
1
1
1
1
0

1
1
1
1
2
1
1
1
1
2

1
1
1
1
2
1
1
1
1
2

1
1
1
1
0
1
1
1
1
0

1
1
1
1
0
1
1
1
1
0

1
1
1
1
0
1
1
1
1
0

x 2 + y2 , z2
Jz
x 2 y2
xy
( xz, yz), ( Jx , Jy )
z

( x, y)

their symmetry whose meaning we wont worry about) and the conjugacy classes along the top. Thus
the character table for 4/mmm is as shown in Table 4.5.10 Careful comparison shows that the character
shown in Table 4.4 is the one conventionally labelled Eu . Note too that the first character is somewhat
trivial, with every value equal to 1: this is clearly possible no matter which group we consider, and is
called the totally symmetric representation.
The significance of the character table is that we can decompose any function of position in our
solid the electron density, the wavefunction, or whatever else were interested in into terms that
transform according to individual characters. Some particular functions are listed in the final column of
the character table. For instance, the function x2 + y2 transforms according to the totally symmetric
representation A1g : any operation of the group will leave this function completely unchanged. On the
other hand, the function xy transforms as B2g : the value of this function remains the same if, say, we
but becomes its negative if we apply 2x or 4. Functions that transform into
apply the operations 2z or 1,
one another are listed in brackets; thus in this point group, x and y do not transform independently, but
together transform as Eu .

4.5. Symmetry of the order parameter

91

4.5 Symmetry of the order parameter


After our discursion we return now to Landau theory. We have already discussed the symmetry
requirements of our free energy expansion, arguing for instance that we would (often) expect positive
and negative values of to be equally likely as justification for including no terms of odd power in our
expansion of the free energy (equation 3.1). We can now express this intuition in more precise terms by
saying that in Landau theory, the free energy expression always has the full symmetry of the high-symmetry
phase. In other words, the terms in the free energy expansion must all transform according to the totally
symmetric representation.
Returning to our example in Figure 4.1, we consider a phase transition that involves only the
movement of the central (black) ion, whose coordinates we denote x, y, z, measured from the centre of
the cell. According to the character table 4.5 and our symmetry principle, the only possible terms in our
free energy expansion are comprised of x2 + y2 and z2 . So our free energy in this case will begin with
the terms
F ( ) = F0 + 21 A1 ( T )z2 + 12 A2 ( T )( x2 + y2 ) +

(4.14)

This looks a bit more complicated than eq. 3.1: we seem to have two order parameters! Note, however,
that we have no reason to suppose any particular relationship between A1 and A2 . Looking back at
Figure 4.1, we can see that movement of the central ion in the xy plane is likely to have a different
energetic effect to movement in the z direction. Now we know that above TC , both A1 > 0 and A2 > 0
(otherwise the equilibrium free energy would not be F0 ); and at TC at least one of A1 and A2 must be
equal to zero (otherwise there would be no phase transition). Because of our assumption that A1 and A2
are independent, we can conclude that exactly one of A1 and A2 becomes zero at the phase transition.
This one corresponds to the primary order parameter that we need to consider.
So we have two possible types of phase transition to consider, depending on whether 2 = z2 or
= x2 + y2 (meaning that itself lives somewhere in the ( x, y) plane). If we take z as the order
parameter, we have, by the same analysis as in the previous chapter,
2

F (z) = F0 + 12 a( T Tc )z2 + 41 bz4

(4.15)

(compare equation 3.2).


On the other hand, if we take ( x, y) as the order parameter, we have two possible fourth-order terms,

Chapter 4. Symmetry

92
x4 + y4 and x2 y2 . Thus our free energy expansion is

F ( x, y) = F0 + 12 a( T Tc )( x2 + y2 ) + 41 b1 ( x4 + y4 ) + 12 b2 x2 y2 .

(4.16)

Just as before, we can minimise with respect to the order parameter to find the equilibrium values of x
and y. We set both derivatives to zero:

F
= x a( T TC ) + b1 x2 + b2 y2 = 0
x

F
= y a( T TC ) + b1 y2 + b2 x2 = 0
y

(4.17)
(4.18)

Solving these equations simultaneously gives three possible solutions:


1. x = y = 0. By assumption, this is stable above TC and unstable below.
p
p
2. x = 0, y = a( TC T )/b1 or y = 0, x = a( TC T )/b1 . Substitution into equation 4.16
shows that these four solutions have the same free energy.
3. a( T TC ) + b1 x2 + b2 y2 = 0 and a( T TC ) + b1 y2 + b2 x2 = 0. Subtracting one of these expressions
from the other gives

(b1 b2 )( x2 y2 ) = 0.

(4.19)

Just as we argued for A1 and A2 , we have no reason to believe any special relationship between b1
and b2 , so can discount the case where these happen to be equal. Hence this gives the solution
p
x = y = a( T TC )/(b1 + b2 ). Again these four solutions have equal free energy.
Just as before, we could go on to calculate the conditions on a, b1 , b2 for each of these solutions to be
preferred, as well as the corresponding susceptibilities and other relevant quantities.
We note finally that the two possible order parameters, ( x, y) and z, each transform according
to a single irreducible representation of the point group of the high-symmetry phase. This is a key
assumption of Landau theory, sometimes called the second criterion of Landau theory. The representation
of the order parameter is known as the active representation. Because it indicates the symmetry lost in
the phase transition, it is never the totally symmetric representation of the high-symmetry phase but
always the totally symmetric representation of the low-symmetry phase.

4.6. Symmetry of the spontaneous strain

93

4.6 Symmetry of the spontaneous strain

(a)

Last chapter we considered the idea of spontaneous strain, a component of the strain tensor that changes
on going through the phase transition. This is also known as a secondary order parameter. Can we apply
similar symmetry arguments to determine the form of this contribution to the free energy?
In fact we have already seen that at least two different situations are possible: linear-quadratic
coupling or bilinear coupling. For an example of the first we return to our tetragonal example for a final
time. Lets consider the first case we discussed, where the order parameter is the motion of the central
cation (or, equivalently, the polarisation) along the z direction. This is the hypothetical transition from

P4/mmm to P4mm in BaTiO in real life this proceeds in one step from the cubic space group Pm3m
3

all the way to P4mm (the transition just below 400 K in Figure 1.11). This is associated with spontaneous
(tensile) strain in the z direction: that is, e3 in Voigt notation. For convenience in this example we will
just write e.
An informal version of our symmetry argument would run as follows: we cannot have an e term in
our free energy, because the +z and z directions are equivalent in the high-symmetry phase. Thus it
would make no sense for polarisation in one direction to cause an expansion along the z axis, and in the
other to cause a contraction! More formally, we see that since transforms as z and e as z2 , e would
transform according to z3 , which according to the character table has the character A2u . Since this is not
the totally symmetric character, e is not a possible term in the free energy.
On the other hand, e 2 transforms as z4 which has the required character A1g . So in this case the
lowest-order coupling term in the Landau free energy expression will be proportional to e 2 . Thus we
have the situation described in equation 3.97 and the analysis following it. In this case, Landau theory
predicts that the critical temperature is not affected by the spontaneous strain. However, the value of
b has effectively been changed by the strain, which as discussed last chapter may cause an otherwise
second-order transition to become first-order. Furthermore, given our usual mean-field relationship
| T Tc |1/2 , we predict from equation 3.98 that the spontaneous strain e should be proportional to
| T Tc |.
For an example of bilinear coupling, we consider another material: the phase transition in LaNbO4
from I41 /a to I2/a is shown in Figure 4.3. The primary order parameter for this transition is shearing
in the xy plane, described in Voigt notation by the strain e6 . (An rough way of thinking about this is the
deviation of the angle from 90 .) However, there is also a secondary strain: in the monoclinic phase
there is no longer any reason for the a and b lattice parameters to be equal, and indeed they gradually

(b)

Figure 4.3: The (a) high-temperature, tetragonal and


(b) low-temperature, monoclinic structures of LaNbO4 .
La atoms are shown in blue, Nb in green, and O in
red; the NbO4 tetrahedra that distort as a result of
the phase transition are also indicated in green. The
structure is shown projected onto the xy plane of the
tetragonal phase. The unit cell is shown in black;
the shift of origin between the two phases is purely
conventional.

Chapter 4. Symmetry

94

Table 4.6: The character table for the point group 4/m.

4/m

4+

4 +

Ag
Bg

1
1
1
1
1
1
1
1

1
1
1
1
1
1
1
1

1
1
i
i
1
1
i
i

1
1
i
i
1
1
i
i

1
1
1
1
1
1
1
1

1
1
1
1
1
1
1
1

1
1
i
i
1
1
i
i

1
1
i
i
1
1
i
i

Eg
Au
Bu
Eu

11 This

is much easier to see on a diagram than to describe in


words. Stare at Figure 4.3 until you feel confident you can
see why the two shearing directions are not symmetrically
equivalent.

x2 + y2 , z2 , Jz
xy, x2 y2

( xz, yz), ( Jx , Jy )
z

( x, y)

begin to differ from one another. This is equivalent to an elongation in the x y direction: in Voigt
notation e1 e2 .
Again, lets make an informal argument before justifying it with group theory. This time the highsymmetry phase is not symmetrical with respect to positive and negative values of e6 : one corresponds
to shearing in the direction in which the tetrahedra are already rotated, the other to shearing in the
opposite direction.11 It is thus intuitively plausible that, if shearing in one direction leads to an increase
in e1 e2 (that is, causes a to increase relative to b), then shearing in the opposite sense should lead
to the opposite effect. Formally, by consulting the character table for the high-symmetry point group
4/m (Table 4.6), we see that e6 (e1 e2 ) transforms as xy( x2 y2 ), which carries the totally symmetric
representation A g so that this is an allowable term in the free energy.
Thus in this example our Landau free energy expression contains a term proportional to e (where,
to emphasise the relationship to our general formalism, we have set e = e1 e2 and = e6 ): This time
our analysis will follow that of equation 3.104. That is, the effect of the spontaneous strain is to stabilise
the low-symmetry phase and hence to increase the transition temperature by an amount proportional to
the square of the coupling constant .

4.7. Group-subgroup relationships across phase transitions

95

4.7 Group-subgroup relationships across phase transitions


As we alluded to earlier, there is often a group-subgroup relationship between the space groups of
a material on either side of a phase transition.12 We can distinguish two possibilities. In the cases
we have considered so far, the unit cell vectors have remained the same across the phase transition,
while some point symmetry operations have been lost. Subgroups that behave in this way are called
translationengleiche, translation-preserving. The alternative is to keep the point symmetry operations
but lose some translational symmetry operations; in this case the subgroup is called klassengleiche,
class-preserving. In the context of phase transitions, a transition to a translationengleiche subgroup is
called ferroic, since as we have seen such transitions may give rise to properties such as ferroelectricity
or ferroelasticity; a transition to a klassengleiche subgroup is called non-ferroic since these transitions do
not give rise to such phenomena.
Of course, it may be possible to lose both point and translational symmetry elements on going
from a space group G to one of its subgroups H. However, a theorem of group theory says that in this
case there is always some intermediate subgroup I such that H < I < G, where one of the subgroup
relationships is translationengleiche and the other is klassengleiche. So we can always divide up symmetry
loss neatly into these two categories.
As an example, we consider the phase transition in SrTiO3 , shown in figure 4.4. This phase transition
involves rotations of neighbouring TiO6 octahedra in opposite directions, with a corresponding change
to I4/mcm. Considering the effects on the unit cell, we can see that the a and
in symmetry from Pm3m

b cell vectors each increase by a factor of 2 while the c vector doubles. This also changes the crystal
system from cubic to tetragonal. Thus both translational and point symmetry operations are lost in this
process. According to the theorem mentioned above, we can consider these in separate stages. First,
to P4/mmm, a translationengleiche subgroup. This is the transition from cubic to
we move from Pm3m
tetragonal symmetry, so that the a and b axes can behave differently from the c axis. In the second step,
we move from P4/mmm to I4/mcm, a klassengleiche subgroup. In the first of these space groups every
octahedron must behave identically while in the second space group every other octahedron behaves
identically, giving the low-temperature phase shown in figure 4.4.
We will not go very far into the representation theory of space groups. The main result well need is
that each wavevector q in the Brillouin zone of a lattice corresponds to a representation of the group
of translations of that lattice in which translation by some lattice vector T is represented by the 1 1


matrix exp(iq T) . Thus the active representation of a phase transition to a klassengleiche subgroup

12 An

exception is the case of reconstructive phase transitions,


where many chemical bonds are broken and the structure is
rebuilt from scratch. In this case it is less likely that there is
a group-subgroup relationship between the two space groups
involved, although more subtle relationships (such as both
space groups being subgroups of some common group) may
exist.

Figure 4.4: Two phases of SrTiO3 : high-temperature,


cubic on the left; low-temperature, tetragonal on the
right. Sr ions are shown in green, TiO6 octahedra
in blue. For ease of comparison, both structures are
shown referred to the low-temperature unit cell.

Chapter 4. Symmetry

96

can be labelled with a q vector sometimes called the propagation vector (especially in magnetic contexts).
This is simply the wavevector of the mode that freezes in at the phase transition. In other words, it is
the wavevector q associated with the variation of the local contribution to the order parameter in the
low-symmetry phase:
= site exp(iq r)
(4.20)
all sites

(this repeats equation 1.4). In our example of SrTiO3 , we can take the rotation angle of the octahedra
as our site variable . We see that, if the octahedron at (0, 0, 0) twists one way ( = 0 ), those at (1, 0, 0),
(0, 1, 0), and (0, 0, 1) will twist the other ( = 0 ); those at (1, 1, 0), (0, 1, 1), and (1, 0, 1) will match the
one at the origin ( = 0 again), and so on.13 This can be summarised by setting q = ( 12 , 21 , 12 ). Check for
yourself that substituting this into equation 4.20 together with the values weve just described gives a
uniformly positive contribution to the order parameter.

13 Note

that we are using the coordinate system of the highsymmetry phase, which is generally the right way to describe
these wavevectors.

4.8

Summary
1. A symmetry of a materials structure is a geometric transformation that leaves the structure
invariant. Many phase transitions arise from broken symmetry: that is, some symmetries of the
high-temperature phase do not remain in the low-temperature phase.
2. The symmetries of any given structure form a mathematical group.
3. Crystal structures may contain both translational and rotational (also known as point) symmetry.
4. There are five pure point symmetry operations: the identity, inversion, mirror reflection, proper
rotation, and rotoinversion (rotation followed by inversion). If we count rotoinversion as a sort of
rotation, then all of these can be described as rotations.
5. These five operations can be combined into 32 crystallographic point groups. Of these, 11 are
centrosymmetric and 10 are polar. The 32 groups can be futher divided into seven crystal systems.
6. The possible categories of translational symmetry are defined by the fourteen Bravais lattices,
which are divided into seven lattice systems. Each lattice system (roughly) corresponds to a crystal
system, which describes the possible point symmetry groups compatible with the corresponding
lattices.

4.8. Summary

97

7. Combining point and translational symmetry gives two further possible symmetry operations:
glide reflection and screw rotation. Both consist of a point operation followed by a translation by
a rational fraction of a lattice vector.
8. The possible point and translational symmetry operations can be combined into 230 crystallographic
space groups.
9. Group operations can be represented by matrices. A set of matrices that obey the same multiplication table as a group is called a representation of that group, and the set of traces of those matrices
is called a character.
10. For any group, there are only a finite number of irreducible representations that cannot be broken
down into representations of smaller degree. Any function in 3D space can be written as a sum of
functions that transform in the same way as (according to) the irreducible representations of a
given group.
11. The free energy of a system that undergoes a phase transition must have the full symmetry of the
high-symmetry phase; i.e., it must transform according to the totally symmetric representation.
This criterion can help us decide what terms coupled to the spontaneous strain are possible.
12. In Landau theory, the order parameter will always transform according to a single, different
representation called the active representation.
13. A translationengleiche subgroup keeps the translational symmetry of the parent group but loses
some rotational symmetry; a klassengleiche subgroup keeps the rotational symmetry but loses
some translational symmetry. All group-subgroup relationships can be divided into these two
components.
14. A transition to a translationengleiche subgroup is called a ferroic transition. If this involves a change
between two different crystal systems, this is a ferroelastic transition. If it involves a change from a
non-polar to a polar point group, it is a ferroelectric transition.
15. A transition to a klassengleiche subgroup is called non-ferroic. This can be labelled by a propagation
vector q.

Chapter 4. Symmetry

98

4.9

Short questions to test your knowledge


1. What is a symmetry? Why is it appropriate to describe the symmetry of a material in terms of
group theory?
2. Define the following symmetry operations: identity, inversion, rotation, rotoinversion, screw
rotation, mirror reflection, glide reflection, translation.
3. Define translationengleiche and klassengleiche subgroups. Why need we only consider these two
types of group-subgroup relationship?
4. What is a ferroelectric phase transition? What are the conditions for one to occur?
5. What is a ferroelastic phase transition? What are the conditions for one to occur?
6. What is the totally symmetric representation? What does it mean to say that the free energy
transforms as the totally symmetric representation of the high-symmetry space group? Explain
why this must be the case.
7. Explain the effect of the criterion in the previous question on the symmetry of the spontaneous
strain in a phase transition.

4.10 Longer questions

1. Draw sphere diagrams for the point groups 2/m, mm2, and 4m2.
2.

(a) Confirm, following footnote 7, that you can generate the three lattice vectors (1, 0, 0), (0, 1, 0),
and (0, 0, 1) and the centring vector ( 12 , 12 , 12 ) for a body-centred lattice by taking linear


combinations of the primitive lattice vectors ( 12 , 12 , 12 ), ( 12 , 12 , 21 ), ( 12 , 21 , 12 ) .
(b) In the same way, find three primitive lattice vectors that can describe the lattice vectors
(1, 0, 0), (0, 1, 0), and (0, 0, 1) and the centring vector ( 21 , 12 , 0) for a base-centred lattice. Show
that your answers are correct.

3. List the four symmetry elements of the monoclinic point group 2/m (your sphere diagram from
the previous question may help). By convention, we align the twofold axis along the y direction.
Write down the characters corresponding to

4.10. Longer questions

99

(a) polarisation in the x direction


(b) polarisation in the y direction
(c) the strain element e1
(d) the strain element e4 .
(You should be able to do this by inspection, but can check by looking up a character table, for
instance online at http://www.cryst.ehu.es/rep/point.html.)
4. Suppose that e4 is the primary order parameter for a phase transition from 2/m to the triclinic
What will be the lowest-order coupling term in the Landau free energy between
point group 1.
= e4 and
(a) the spontaneous strain e1 ?
(b) the spontaneous strain e6 ?
5. The mineral VO2 undergoes a phase transition from space group P42 /mnm to P21 /c on cooling
below 343 K. Is this transition likely to lead to ferroelectricity in the low-temperature phase? What
about ferroelasticity?
6. The mineral NaNbO3 undergoes a phase transition from the ideal cubic perovskite structure to a
tetragonal structure with space group P4/mbm (Figure 4.5). This is similar to the tetragonal phase
of SrTiO3 in Figure 4.4, except that there is no cell doubling in the z direction, so that adjacent
octahedra along z rotate in the same direction.
(a) Copy Figure 4.5 and annotate it to show (i) a fourfold axis (the conventional symbol for this
is ); (ii) a glide plane (the conventional symbol is
).
(b) What q vector is associated with this phase transition?

Figure 4.5: The tetragonal phase of NaNbO3 . Na


atoms are shown in yellow, Nb in green, O in red.

Chapter 5

D isplacive phase transitions and soft modes

5.1 Displacive phase transitions


Displacive phase transition in quartz
We have already met a number of displacive phase transitions: the ferroelectic phase transition in
PbTiO3 , the rotational phase transition in SrTiO3 , and the ferroelastic phase transition in Na2 CO3 . Here
we want to introduce two more, namely those in the quartz and cristobalite polymorphs of silica, SiO2 .
Actually cristobalite is the stable form of silica at very high temperatures, but it is very easily quenched
below its stability point to effectively appear as a separate stable form at ambient temperature.
The crystal structure of the high-temperature -phase of quartz has hexagonal symmetry, which is
lowered to trigonal on cooling to the low-temperature -phase, with the transition temperature of 858 K.
The two structures viewed down the common [001] axis the axis with the 6-fold and 3-fold axes are
shown in Figure 5.1. The key characteristic of the structure is the existence of tetrahedral SiO4 groups
of atoms that are linked together through the corner oxygen atoms. The structure plots in Figure 5.1
and later in this chapter will highlight these topological groups of atoms. In fact it appears that the
SiO bond length doesnt change as a result of the phase transition, which instead is achieved through
rotations of the SiO4 tetrahedra as is clear from Figure 5.1.
The order parameter associated with the phase transition in quartz is defined through the major
displacements of the atoms, which can be projected onto a rotation of the SiO4 tetrahedra. The rotation
angle is shown as a function of temperature in Figure 5.2, where we have plotted the quartic power of
the rotation angle. The phase transition in quartz is weakly first-order for reasons that we will revisit
later in this chapter and thus the behaviour of the order parameter is close to that of a tricritical phase
transition.
The phase transition in quartz is accompanied by significant changes in the lattice parameters;
101

Figure 5.1: The crystal structures of the hightemperature hexagonal (top) and low-temperature trigonal (bottom) phases of quartz, viewed down the [001]
axes. The SiO4 tetrahedra are shown as shaded objects
rather than individual atoms.

Chapter 5. Displacive phase transitions and soft modes

102

5.00

5.48

4.98

5.46

4.96

5.44

4.94

5.42

4.92

5.40

80000

a ()

60000

40000

20000

0
0

200

400

600

800

1000

T (K)

Figure 5.2: The angle of the symmetry-breaking rotation of the SiO4 tetrahedra in quartz vs temperature,
plotted to the fourth power to highlight the near
tricritical character of the phase transition.

4.90
0

c ()

(Tilt Angle)4 (deg4)

5.38
200 400 600 800 1000
Temperature (K)

Figure 5.3: Temperature dependence of the lattice


parameters of quartz. Although the phase transition
does not induce a symmetry-breaking strain, nevertheless there is a large spontaneous strain associated
with the phase transition.

because these changes do not change the symmetry at all, they scale as the squared power of the order
parameter. The temperature-dependence of the lattice parameters of quartz are shown in Figure 5.3.
It should be noted the two phases of quartz contain the same number of atoms (three formula
units). Thus in terms of thinking of the modulation of the structure by imposition of a wave of atomic
displacements, every unit cell behaves the same way and the modulation wave vector is qc = [000].
Figure 5.4: The crystal structures of the hightemperature cubic (top) and low-temperature tetragonal (bottom) phases of cristobalite, viewed down a
common axis that is the [001] axis in the tetragonal
phase. The SiO4 tetrahedra are shown as shaded objects rather than individual atoms.

Displacive phase transition in cristobalite


The crystal structure of the high-temperature -phase of cristobalite is F-centred cubic, and is based
on the diamond structure. The silicon atoms occupy the main sites of the diamond lattice, which are
tetrahedrally coordinated to nearest-neighbour sites. The full silica structure is achieved by having the
oxygen atoms occupy the sites half-way between all nearest-neighbour silicon atoms. As in quartz, this
leads to a structure that can be described in terms of corner-linked SiO4 tetrahedra. And as in quartz,
the phase transition to the tetragonal phase at a temperature of around 400 K involves rotations
and displacements of the SiO4 tetrahedra without any noticeable effect on the size and shape of the

5.1. Displacive phase transitions

103

tetrahedra. The structure of the phase is compared with that of the phase in Figure 5.4.

The displacive phase transition in cristobalite is strongly first order, with significant reduction in
volume. As for quartz, the phase transition involves rotations of nearly-rigid SiO4 tetrahedra. The
square of the rotation angle is plotted as a function of temperature in Figure 5.5, where the strongly
first-order nature of the phase transition can be seen. The lattice parameters are plotted as functions of
temperature in Figure 5.6. As in the case of quartz, the spontaneous strains scale as the square of the
order parameter.
In our discussions on quartz and cristobalite we have commented on how the phase transition
appears to involve rotations and displacements of SiO4 tetrahedra. Experiments that probe atomic
structure over short distances show clearly that there are no distortions of these tetrahedra through the
phase transition, and that they genuinely appear to move as nearly-rigid objects. This observation is the
basis for one way to view phase transitions, and we will explore this towards the end of this chapter.

0.6
Displacements2 (2)

Unlike the phase transition in quartz, the size of the unit cell (in terms of number of atoms in the
primitive unit D
cell) doubles
on cooling through the phase transition. In this case, the modulation wave
E
11
vector is qc = 2 2 0 .

0.4

0.50
0.45

0.2

0.40
0.35
300

0.0
0

400

500

200
400
Temperature (K)

600

Figure 5.5: Sum of the symmetry-breaking displacements of the atoms in cristobalite vs temperature, plotted to the squared power. The solid curve is the fitted
Landau function order parameter.

Displacive phase transitions in molecular crystals

Malononitrile, C3 H2 (CN)2
The phase transition in malononitrile is peculiar in one regard. Over the small range of temperatures
between 295 K and the melting temperature, the temperature range of the -phase, the crystal of
malononitrile adopts a monoclinic structure. On cooling through the phase transition to the -phase,
small rotations of the molecules cause a lowering of the symmetry to triclinic; the centre of symmetry is
preserved, but the structure loses its screw axis and glide plane symmetries. The number of molecules

7.15
Cell parameter ()

We have seen above examples where groups of atoms move together through a displacive phase
transition, and previously in ferroelectric phase transitions in the perovskite family cases where ions
move individually against each other. Molecular crystals are cases where the groups of atoms that
constitute molecules clearly move together. We cite two molecular crystals that show displacive phase
transitions.

a (cubic)

7.10
7.05

2a (tetragonal)
7.00
6.95
6.90

c (tetragonal)
300

400 500 600


Temperature (K)

700

Figure 5.6: Temperature dependence of the lattice parameters of cristobalite. The phase transition involves
a large change of volume in addition to the strains
associated with the lowering of symmetry from cubic
to tetragonal.

104

Chapter 5. Displacive phase transitions and soft modes

Rotation squared, scaled strain

60
50
40
30
20

Figure 5.7: The crystal structures of the high/low-temperature monoclinic / phases (left) and intermediatetemperature triclinic phase (right) of malononitrile, C3 H2 (CN)2 , viewed down the common [010] axes.

10
0

150

200
250
Temperature (K)

300

Strain x 1000, squared

in the unit cell remains fixed, meaning that the modulation wave vector is qc = [000]. The crystal
structures are shown in Figure 5.7.

6
5
4
3
2
1
0

150

200
250
Temperature (K)

300

Figure 5.8: The square of the symmetry-breaking rotation of the malononitrile molecule (black filled circles)
and spontaneous strain e1 as function of temperatures
(top), and squares of the symmetry-breaking strains
e4 (black filled circles) and e6 (bottom). In all cases
the data are plotted with the curve ( T T1 )( T2 T ),
where T2 > T1

On cooling below 141 K there is a second phase transition to the -phase; the peculiar thing about this
phase transition is that the -phase has the same monoclinic structure as the high-temperature -phase.
It is very unusual for a phase transition to have the lower-temperature phase as the higher-symmetry
structure, and very unusual for a sequence of two phase transitions in which the high-temperature
and low-temperature phases to have the same symmetry. Such a sequence of phase transitions is
called a sequence of re-entrant phase transitions. The order parameter will only be non-zero within the
temperature range of the intermediate lower-symmetry phase.
For malononitrile, the order parameter can be described as the symmetry-breaking rotation angle,
and its temperature dependence has been determined from neutron diffraction and is shown in Figure
5.8, together with the symmetry-preserving straine1 and the squares of the symmetry-breaking strains e4
and e6 . In this system, e1 is proportional to the square of the order parameter, whereas the spontaneous
strains charge the symmetry in the same way as the order parameter and thus are proportional to the
value of the order parameter. Although the symmetry changes are consistent with a ferroelastic phase
transition, the strains are too small for an elastic instability to be the origin of the phase transition.

5.1.

where the errors are standard deviations and no corrections have been made for film
shrinkage or other systematic errors. The systematic absences were consistent
with the
Displacive phase transitions
105

0".

__--1

'.

e -

'

\!

2. The phase rhombohedral


change in s-triazine
alongCthe
of the high-temperature
Figure 5.9: The crystal structure ofFigure
the high-temperature
phase ofviewed
sym-triazine,
H3axis
, viewed
3 N3a
down the 3-fold axis (left), and a hexagonal
view from an
direction showing
of low-temperature
the crystal and themonoclinic cell. The
cellorthogonal
which is identical
to the 6 the
axisshear
of the
rotations of the molecules (right). full and broken lines refer to the high- and low-temperature structures respectively and 15
the

Sym-triazine is a planar symmetric molecule. The crystal structure of the high-temperature phase
is rhombohedral, and the phase transition at 198 K involves the lattice changing to monoclinic with
rotations of the molecules. In this case, there is a significant shear of the unit cell that breaks the
symmetry in the same way as the molecular rotations, and the phase transition is correctly described as
a ferroelastic phase transition. The crystal structures of the two phases are shown in Figure 5.9. The
temperature dependence of the molecular rotation and shear angles are shown in Figure 5.10.

Incommensurate phase transitions


We have characterised many phase transition in terms of an order wave vector qc . For ferroelectric
phase transitions, qc = 0. For phase transitions that involve, say, a doubling of the unit cell, qc will

Angle (deg)

Sym-triazine, C3 N3 H3

chain dotted lines outline the reduced low-temperature cell. The thick lines (full and broken)
are
show the molecular planes. The shear angle (e) and the molecular rotation angle (4)10
indicated.
5
0

50

100
150
Temperature (K)

200

Figure 5.10: Molecular rotation angle (open circles)


and shear strain (closed circles, represented by an
angle) of sym-triazine as functions of temperature.

Chapter 5. Displacive phase transitions and soft modes

106

h
i
have a value at the edge of the Brillouin zone, say qc = 12 21 12 in SrTiO3 . It is possible for some
phase transitions to have a value of qc that is not related to the periodicity of the lattice. Such phase
transitions are called incommensurate phase transitions, because the period of the modulation vector
is not commensurate with the underlying periodicity of the crystal. A cartoon representation of a
modulation with an incommensurate wave vector is shown in Figure 5.11, and the diffraction from an
incommensurate phase showing satellite reflections is shown in Figure 5.12 Whilst this idea may seem
to be somewhat exotic, incommensurate phase transitions are more common than you might imagine.

Displacive phase transitions in general: some open questions

Figure 5.11: Cartoon representation of an incommensurate modulation of a simple lattice.

What all these examples show is that a displacive phase transition involves a change in the symmetry
of a crystal structure through small symmetry-breaking displacements of the atoms from their mean
positions in the higher-symmetry phase. In this sense there is no sense of structural disorder as we have
seen in the ferromagnetic or site-ordering phase transition.
Based on what we understand as the origin of phase transitions in disordered systems coming from
the balance between entropy (from disorder and favoured at high temperature) and enthalpy (that
favours the ordered state at low temperature), it is not clear at face value what drives displacive phase
transitions. It is fair to assume that the symmetry-breaking displacements lower the energy of the
structure, and we will explore why this might the so in one of the exercises at the end of the chapter.
But we still have a number of questions that we will explore in this chapter:
1. Why can displacive phase transitions occur?
2. What is the mechanism by which displacive phase transitions occur?
3. Why one instability and not another?
4. What is the nature of the high-temperature phase?
5. What determines the transition temperature?

Figure 5.12: Diffraction from the incommensurate


phase of Ca2 Al2 O5 , showing satellite reflections.

6. Why does Landau theory work so well for displacive phase transitions?
We will tackle all these questions in this chapter, albeit not in this order.

5.2. Phenomenology of the soft mode model of displacive phase transitions

107

5.2 Phenomenology of the soft mode model of displacive phase transitions


Soft modes and dielectric divergence
Before the study of phase transitions became significant, and in a rather different context, three physicists
Lyddane, Sachs and Teller showed that for a simple cubic material there is a relationship between the
dielectric constants at zero and infinite frequency of applied field e(0) and e() respectively and the
angular frequencies of the transverse and longitudinal optic phonons at zero wave vector TO (q = 0)
and LO (q = 0) respectively:
2 (q = 0)
e (0)
= LO
(5.1)
2 (q = 0)
e()
TO

2 T T0

(5.2)

This is not surprising, given that the LST relationship suggests that 2 1/e(0) ' 1 ; this result is
clearly consistent with mean-field theory and Landau theory, and we have already suggested that many

Frequency2 (THz2)

The infinite-frequency dielectric constant arises from the ability of an applied electric field to polarise
the material through displacing the spatial arrangement of electronic charge relative to the atomic nuclei.
Although it isnt strictly an infinite-frequency field, it is designated as such because it has a much faster
variation than the atomic nuclei can respond to, with the assumption that electrons can, because of their
low mass, respond instantaneously to the fast electric field.
We have previously seen that the static (zero-frequency) dielectric susceptibility and hence dielectric
constant (e(0) = + 1) diverges at a second-order ferroelectric phase transition. Since e() will have a
constant value, a divergent value of e(0) will mean a divergent value of the left-hand side of equation
5.1. The LO mode frequency depends on the various force constants, and cannot diverge, therefore the
only way for the right-hand side of equation 5.1 to diverge is for the value of TO (q = 0) to fall to zero
at the phase transition.
Looking back at the phonon dispersion curves of the ferroelectric material PbTiO3 in the hightemperature cubic phase shown in Figure 3.17, there is a transverse optic mode whose frequency falls
dramatically to zero at wave vector q = [000]. Figure 5.13 shows the temperature dependence of the
frequency of this mode, showing the decrease to zero on cooling towards the phase transition, followed
by the increase in frequency on cooling below the phase transition. It can be seen that the temperature
dependence for T > Tc is approximately described as

15

10
2)

0
200

400

600
800
Temperature (K)

1000

1200

Figure 5.13: Temperature dependence of the square


of the soft mode frequency in PbTiO3 , measured by a
combination of inelastic neutron scattering and Raman
spectroscopy.

108

Chapter 5. Displacive phase transitions and soft modes

displacive phase transitions are well-described by Landau theory.


This leads to a suggestion that a displacive phase transition will be accompanied by a phonon whose
frequency will fall to zero at the transition temperature this phonon is called the soft mode. Although
the LST relationship is specifically concerned with a phonon of wave vector q = 0 in a cubic crystal, It
is not unreasonable to anticipate the idea will also apply to phase transitions in crystals of other crystal
classes, and for other wave vectors.

Soft modes and ferroelectric phase transitions

a)

A phonon has an associated and well-defined set of atomic displacements, which (for reasons we will
see later) are called the phonon eigenvector. The modulation of these displacements from unit cell to unit
cell is defined by the wave vector q of the soft mode. First we are going to consider the case q = 0.

b)

Frequency

T > Tc
T > Tc
T Tc
T < Tc
Wave vector

Figure 5.14: Schematic representations of a ferroelectric soft mode behaviour: a) behaviour of the phonon
dispersion curves; b) atomic displacements.

The acoustic phonons at q = 0 are defined as the modes where all atoms are displaced equally; since
there are three directions, there are three acoustic modes corresponding to atomic displacements in each
of these three directions. On the other hand, the optic phonons have out-of-phase atomic displacements.
In a simple crystal with two atoms in the unit cell, the optic modes at q = 0 correspond to the two
atoms moving in opposite directions, weighted so as to ensure the centre of mass does not move.
Figure 5.14 shows a cartoon representation of a ferroelectric phase transition from a simple cubic
material with 2 atoms in the unit cell, together with a schematic representation of the softening of the
transverse optic phonon branch.
In the case of a cubic perovskite lets take PbTiO3 as an example there will be 15 phonons for
each wave vector. At q = 0, these are divided into modes with separate motions along x, y and z, and
we consider modes whose motions lie along z. In PbTiO3 , the four optic modes in this group include
a mode where the Pb2+ cation is displaced relative to the TiO6 octahedron, which corresponds to the
soft mode, a mode where two oxygen atoms move in opposite directions to give zero polarisation, and
modes where the TiO6 octahedra distort in concert with a displacement of the Pb2+ cation. The atomic
motions in the soft mode break the centre of symmetry and give rise to a net dielectric polarisation if
frozen into the structure. And indeed, the phase transition can be described exactly as the freezing in of
the phonon displacements below the transition temperature, with an amplitude that grows on cooling
and which corresponds to the value of the order parameter.

5.2. Phenomenology of the soft mode model of displacive phase transitions

109
.

A number of crystals undergo phase transitions which involve soft modes with wave vectors at Brillouin
zone boundaries. In these cases the soft phonons can be either acoustic or optic modes, but because of
mixing of eigenvectors the distinction between the two types of mode at zone boundary wave vectors
may not clear. The different types of soft mode, and the atomic displacements, are shown schematically
in Figure 5.15. One of the results of a zone boundary soft mode phase transition is that the unit cell of
the low temperature phase is doubled in one or more directions. In some cases neighbouring unit cells
of the high temperature form develop dipole moments, but as these are in opposite directions the unit
cell of the low temperature has no net moment.
Undoubtedly the best example of a zone boundary phase transition is the cubictetragonal transition in the perovskite SrTiO3 , which we have seen before. The soft mode involves rotations of the
interconnected TiO6 octahedra about [001], with neighbouring octahedra in the (001) plane rotating in
opposite directions. The soft mode has wave vector qc = [1/2, 1/2, 1/2]. The variation of its frequency
with temperature has been measured by both neutron scattering and Raman spectroscopy, and its
temperature dependence, and is shown in Figure 5.16.
As in the case of PbTiO3 , the soft mode frequency in SrTiO3 rises again on cooling below the phase
transition. At the critical wave vector qc = [ 12 , 21 , 21 ] there are three degenerate modes corresponding to
rotations of the TiO6 octahedra about the x, y and z axes. Below Tc the mode that rotates about the z
axis gains an increasing frequency as the rotation angle increases, because this rotation can only occur
with a first-order change in the TO bond length. On the other hand, rotations about the other two axes
remain relatively easy, and so the other components of the soft mode do not increase much on cooling
below Tc .

Soft modes in other materials


Soft modes are also seen accompanying displacive phase transitions in molecular crystals. In malononitrile, whose structure we have previously seen, the soft mode is at zero wave vector. But in this case,
there are two phase transitions, and the soft mode is expected to fall to zero at both temperatures.
This is seen in Figure 5.17, which gives measurements of the soft mode frequency obtained by Raman
spectroscopy. The upper transition is too close to the melting point for measurements above this
temperature to have been possible.

a)

Wave vector

Wave vector

b)

T > Tc

T < Tc

Figure 5.15: Schematic behaviour of a) zone boundary


acoustic and optic soft modes; b) atomic displacements
showing doubling of the unit cell and cancelling induced dipole moments.

2.0
Frequency2 (THz2)

Zone boundary (antiferroelectric) phase transitions

1.5
1.0
0.5
0.0

100
200
Temperature (K)

300

Figure 5.16: The temperature dependence of the soft


mode in SrTiO3 , as measured by a combination of
inelastic neutron scattering and Raman spectroscopy.

Chapter 5. Displacive phase transitions and soft modes

110

Frequency (cm1)

Frequency (cm1)

15
10
5
0

50

100
150
200
Temperature (K)

250

300

Frequency

Figure 5.17: Temperature dependence of the soft mode in malononitrile, CH2 (CN)2 , showing two temperatures at which the
soft mode falls to zero frequency.

cooling
cooling

Wave vector
Figure 5.19: Schematic representation of the softening
of the acoustic mode. The left-hand figure shows the
case where the acoustic mode softens across the whole
branch. The right-hand figure shows the case where
the softening is restricted to small wave vectors. In
both cases, the softening leads to a reduction in the
slope of the acoustic mode dispersion curve in the
limit k 0.

100
80
60
40
20
0

750

800
Temperature (K)

850

Figure 5.18: Temperature dependence of the


soft mode in quartz on heating up to the displacive phase transition, measured by Raman
spectroscopy.

The soft mode in quartz was actually the first to have been observed experimentally, but at the time
was not understood as such. The atomic motions associated with the soft mode involved rotations of
the SiO4 tetrahedra, and the flexibility of the structure allows these phonon modes to propagate without
distortions of the tetrahedra. The soft mode has been measured but only in the low-temperature phase;
the results are shown in Figure 5.18.
The interesting thing about quartz is that the whole phonon branch along the direction from q = 0
to q = a? softens uniformly.

Ferroelastic phase transitions


The soft modes considered so far have been optic modes, or of optic character. Acoustic modes at q = 0
can also soften, but here the basic picture will be somewhat different since at q = 0 the frequency is
already zero. In this case, the softening involves the slope of the acoustic mode falling to zero, and this
corresponds to a softening of the velocity of sound and of an elastic constant (or combination of elastic
constants). Two ways in which this can happen are shown in Figure 5.19. The softening of an acoustic
mode drives a ferroelastic phase transition as discussed in the previous chapter.

J. D. Axe, M.

Iizumi,

Watiojtal LnbojatoI

' and
v,

G. Shirane

Uptojt, lV('It' YoI'I

l /973

(Received 27 F-ebruary 1980)

ncommensurate structur il tr5.2.


insform
ition at T= 128 Koffollowed
i
Phenomenology
the softby mode
model of displacive phase transitions
111
transformation at T, =93 K. Below To, the soft-phonon mode associatis "split" into modes representing fluctuations in &mplitude and phase
iture measurement
ameter. We have studied the dispersion
temper
dependence
Figure and
5.20
shows
of the soft transverse acoustic mode in s-triazine, together with the
K. We have also seen intensity anomalies which can be expl tined by
effective
elastic
constant
associated
with the gradient /q. The softening in this case is confined to
nsverse-acoustic-phonon
branch and the low-lying "ph tson" branch.

18

"

5.0-

Z.B.

high temperature

nce.

Structure
on spectra of
infrared, ' and
ve been attributthe situation is
ects and the lack
eem to be unation.
utron scattering
instability in
transformations.
order transformahombic Pnam
one with wave
he modulation
~+
to q, = ,
rate at lower
1

a,

22

cooling

4.0

Frequency

a growing inalline translationstortions of the


with that of the
in quasi-onelex salts or salts
he layered c's result from a
an electronic
h the lattice is
stortion requires
describing its amhe consequences
-wave distortion
Qo is energetically
naturally to the
mentary excitaquencies which
mit and an
se phase fluctuamodes"4) should
here is at present

temperatures.
Figure 1 shows the phonon dispersion
for the lowest-lying transverse modes propagating in
Incommensurate
phase
transitions
the [I00] direction at several
temperatures
above T.
[The X2 and X3 modes are continuous at the zone
Most
soft-mode
phase transitions
modes with wave vectors at qc = 0 or at one of the Brillouin
convenient toinvolve
and it is therefore
use an exboundary
Brillouin
in which the softtended
(doubled)
zone,
zone boundaries. However, it is possible for the instability to occur at a wave vector that is at a more
mode instability now appears in the vicinity of
general
position
in the
Brillouin
zone.studied
This type of phase transition is the incommensurate phase transition,
CDW metals
to the
] In contrast
q = , q ..
'
'
neutron
thusdiscussed
far by inelastic
K2Se04
as
above.
Herescattering,
we note that
in reciprocal space the length of the vector qc does not match any
0
shows essentially complete phonon softening, ~,
of
the reciprocal
lattice vectors.
as T
To, with relatively small damping, and consefor study
the excitaa good candidate
quently
An isexample
is K2 SeO
which
an incommensurate soft mode has been measured by inelastic
4 , in of
tions below To. The results of such experiments are
neutron
are shown in Figure 5.21, and the softening is quite clearly seen on
This results
the subject scattering.
of this paper. The
study complements
which
have
been
studies
the
scattering
light
cooling towards the phase transition.recently
reported by Wada et al. , Caville et al. , '" and Fleury,
Another
example of an incommensurate phase transition is quartz: the phase transition actually
and Lyons.
Chiang,

3.0

anti-crossing

2.0

1.0

Wave vector
0.5

I.C

qmax

FIG. 1.5.21:
of the soft-phonon
branch ofofK2Se04
Dispersion
Figure
Temperature
dependence
the soft
above To=128
K in on
&n extended-zone
from
Ref. tran8
scheme,the
mode
in K2 SeO
cooling towards
phase
4
(note the symmetry designation changes from X3 to X2 &t
sition,
measured by inelastic neutron scattering.
the zone boundary) ( q measured in units of a ).
340S

O1980 The American Physical Society

Figure 5.22: Schematic representation of the mechanism of the incommensurate softening of an acoustic
mode in quartz as a result of the softening of an optic branch.

Frequency (cm1)

the limit q 0.

16
12
8

C (GPa)

ha&'e'tt

0.2
0

4
0
0.0

0.4

0.2 0.4
q100/a*

200
280
Temperature (K)

0.6

Figure 5.20: The main figure shows the dispersion


curves of one of the transverse acoustic modes for
q k a? for temperatures 105 K (filled points) and 25 K
(open points) above Tc (188 K). The side figure shows
the temperature dependence of the effective elastic
constant.

Chapter 5. Displacive phase transitions and soft modes

112

has an incommensurate phase existing over a small temperature range between the two phases. The
mechanism for this involves a softening of an optic mode and an anti-crossing interaction with an
acoustic mode. The effects of the anti-crossing increase with k, since the optic and acoustic modes have
different symmetry at k = 0, so the result is that the frequency of the acoustic mode is driven to zero at
a non-zero value of k. This is illustrated if Figure 5.22.

5.3

Lattice dynamics theory of the soft mode


The potential energy of a crystal can be expanded in terms of the individual atomic displacements:
E = E0 +

1
2

2 E

0 u j, u 0

j ,0

j,j
,0

1
n!

j(1) , ,j(n)
(1) , ,(n)

u j, u j0 ,0 +

n E
u (1) (1) u j(n) ,(n)
u j(1) ,(1) u j(n) ,(n) j ,

(5.3)

The standard theory of lattice dynamics relies on what is called the harmonic approximation, which
involves taking only the quadratic term (note that the linear term sums to zero when the structure is
at equilibrium). The virtue of the harmonic approximation is that the equations can be solved exactly,
giving rise to vibrations with a well-defined set of angular frequencies at any wave vector. In turn
these vibrations can be quantised; the quanta are called phonons, and their statistics obey the standard
BoseEinstein distribution.
The higher-order terms are the anharmonic terms. Unlike for the harmonic approximation, there is
no exact solution for the anharmonic model. One way forward is to assume that the anhamonic terms
are not large, so that we can treat them as a small perturbation of the harmonic model. In particular, we
retain the concept of the normal mode.
The normal mode coordinate for a vibration of wave vector q is defined from the expansion of the
atomic displacement in terms of normal modes:
u( j`) = p

1
Nm j



e
(
j,
q,

)
exp
iq

j,` Q (q, )

q,

(5.4)

5.3. Lattice dynamics theory of the soft mode

113

where N is the number of unit cells, j labels the atom in the unit cell, ` labels the unit cell, q is a phonon
wave vector, defines the particular branch in the dispersion curves (recall that the number of branches
is equal to 3 times the number of atoms in the unit cell), and r j,` is the position of atom j in unit cell `.
On this basis, we can expand the full anharmonic Hamiltonian in terms of the normal mode
coordinates:

H=

1
1
2
Qq, Qq,
Q q, Q q, + 2 q,
2 q,
q,

1
3!
1
4!

0 00 q,q0 ,q00 Qq, Qq0 ,0 Qq00 ,00 (q + q0 + q00 )


(3)

q,q ,q
,0 ,00

0 00

,0 ,00

q,q ,q ,q000
,0 ,00 ,000

(4)

q,q0 ,q00 ,q000 Qq, Qq0 ,0 Qq00 ,00 Qq000 ,00 )(q + q0 + q00 + q000 )
,0 ,00 ,000

(5.5)

The (n) coefficients are related to the n-th derivative of the energy analogous to 2 for the harmonic
coefficient. The important point in the equation, which is implicit in the harmonic term, is that there is
a conservation law operating for the wave vectors of the normal mode coordinates in each term. This is
represented by the functions (q), where q is any combination of wave vectors, such that
(
1 if q = G
(q) =
(5.6)
0 if q 6= G
where G is a reciprocal lattice vector.
When this Hamiltonian is treated using second quantisation, it can be seen that the anharmonic terms
have a physical interpretation in terms of collisions between phonons. For example, the third-order
anharmonic term corresponds to events such as one phonon spontaneously decaying into two of lower
energy or two combining to form a third, even though this rather stretches the meaning of a collision.
The fourth-order anharmonic interactions include the case where two phonons interact to form two
different phonons, with changes in energy and wave vector.
Although the third-order interactions might be expected to be more significant than the fourth-order
interactions, the constraints on the possible interactions due to the conditions of conservation of energy
and wave vector are more restrictive for the third-order terms than the fourth-order terms. Because of

Chapter 5. Displacive phase transitions and soft modes

114

this, the fourth-order terms will be more important in our analysis. Accordingly our starting point is
the quartic Hamiltonian, where we can neglect the kinetic energy term:

H=

1
1
2
Qq, Qq, +
q,
2 q,
4!

0 00

q,q ,q ,q000
,0 ,00 ,000

(4)

q,q0 ,q00 ,q000 Qq, Qq0 ,0 Qq00 ,00 Qq000 ,000 (q + q0 + q00 + q000 )

(5.7)

,0 ,00 ,000

We now apply a mean-field approximation, by replacing pairs of normal mode coordinates by their
thermal averages. For simplicity, we proceed within the high-temperature limit:
D
E
k T
Qq00 ,00 Qq000 ,000 Qq00 ,00 Qq000 ,000 ' 2B q00 ,q000 00 ,000
(5.8)
q00 ,00
where the average is only non-zero if q00 = q000 by conservation of wave vector, and 00 = 000 by the
orthogonality condition. Thus the anharmonic Hamiltonian can be written as
D
E
1
1
(4)
2
H = q,
Qq, Qq, + q,q,q00 ,q00 Qq, Qq, Qq00 ,00 Qq00 ,00
2 q,
4 q,q00
00 00
,, ,

,00

1
k T
2
q,
Qq, Qq, + B

2 q,
4

00 q,q,q00 ,q00 Qq, Qq, /q2 00 ,00


(4)

q,q
,00

(5.9)

,,00 ,00

There are 6 ways of selecting independent pairs from the fourth-order term in the anharmonic Hamiltonian, which is why the numerical prefactor has been multiplied by a factor of 6. This equation now
looks like a harmonic Hamiltonian, since we have pairs of normal mode coordinates in both the last two
terms. We therefore can merge these terms to obtain

1
k
T
(
4
)
B
2
H = q,
+
q,q,q00 ,q00 / 2 (q00 , 00 ) Qq, Qq,
(5.10)
2 q,
2 q
00 ,00
00 00
,, ,

1
2
Qq, Qq,
= q,
2 q,

(5.11)

where we have defined a new set of renormalised phonon frequencies:


2
2
q,
= q,
+

kB T
2

q,q,q00 ,q00 /q2 00 ,00

00 00
(4)

q ,

,,00 ,00

(5.12)

5.3. Lattice dynamics theory of the soft mode

115

This applies to all phonons in the crystal: the frequency of each phonon is modified by its anharmonic
interaction with all other phonons in the crystal. This mechanism gives a temperature dependence to all
phonons, typically causing an increase in the value of 2 on increasing temperature.

Renormalised phonon theory and the soft mode


If we perform a lattice dynamics calculation on a high-temperature phase of a material that undergoes a
displacive phase transition, at least one of the values of 2 calculated in a harmonic lattice dynamics
simulation will have a negative value. Such a calculation corresponds to zero temperature, and the
calculated frequencies reflect the curvature of the energy with respect to deformations that correspond
to each normal mode. Usually the crystal is at a minimum of the potential energy, and so the curvature
is upwards in energy meaning a positive value of 2 . However, if there is a phase transition, the
structure of the high-temperature phase which is the high-symmetry structure will be unstable
with respect to the distortion that gives the low-symmetry phase, and the curvature of the energy with
respect to the normal mode that freezes into the structure will be downward. Hence the calculation will
give a negative 2 .
For phonons with a negative 2 at zero temperature, equation 5.12 suggest that with positive
average values of (4) the anharmonic terms add a positive contribution to 2 that offsets the negative
harmonic value of 2 until at some temperature the value of 2 becomes positive. At this point, the
high-symmetry structure is stable against fluctuations of the normal mode that freezes into the structure
at low temperatures.
Lets change nomenclature slightly and focus specifically on the soft mode. We denote 0 as the
harmonic angular frequency of the normal mode that is the soft mode, such that 02 has a negative
value. We also denote the pair of wave vector q and phonon branch of any other phonon as q (and
denote the pair with wave vector q as q), so that equation 5.12 can be re-written as
2 = 02 +

kB T
2

(4)

/q2

(5.13)

The value of 2 reaches zero at the temperature


kB Tc =

202
(4)

q q /q2

(5.14)

Chapter 5. Displacive phase transitions and soft modes

116

The temperature dependence of 2 is sketched in Figure 5.23.


For simplicity, we now make useDof the
E Einstein model. We define an average angular frequency h i
(
4
)
and an average coupling coefficient
(the averages can be weighted in order to ensure consistency).
We can therefore write the equation for the transition temperature as

High-temperature
phase stable

RTc =

202 h i2

3 (4)

(5.15)

The temperature dependence of the soft mode is then given as

Harmonic
value

Tc
T
C

Temperature

Low-temperature
phase stable

Figure 5.23: Schematic representation of the temperature dependence of the square of the frequency of a
soft mode. Below the temperature Tc the frequency
is imaginary and hence unstable. The frequency at
T = 0 K is the harmonic value.

2 =

5.4

|02 |
( T Tc )
Tc

(5.16)

Lattice dynamical theory of the low-temperature phase


Effect of the phase transition on phonon frequencies
The theory of the soft mode developed above is applicable to the high-temperature phase, and explains
how the high-symmetry phase can be stable when the potential energy surface is unstable. Because
the theory is only applicable to the high-temperature phase, it does not provide information about the
temperature dependence of the order parameter. This is best obtained using a lattice dynamic model of
the free energy, which we describe in this section. In this case, the soft mode becomes a static distortion
of the crystal the order parameter is the static equivalent of the normal mode coordinate of the soft
mode. In this case the potential energy part of the Hamiltonian has the form

H =
=

1
1
q2 Qq Qq + q Qq Qq 2
2
4
q


1
1
2
2

Qq Qq
q
q
2
2
q

(5.17)

The coefficient q denotes the fourth-order anharmonic coupling of the order parameter with any
(4)

phonon, equivalent to the coefficient q0 ,q , where q0 represents the wave vector and branch label of the
soft mode. The new Hamiltonian means that the static distortion modifies the phonon frequencies to
the form
1
q2 = q2 + q 2
(5.18)
2

5.4. Lattice dynamical theory of the low-temperature phase

117

If q is positive, the frequency is increased on cooling into the low-temperature phase, and this will have
the effect of decreasing the phonon entropy as required. We also remark that the the dependence of the
phonon frequency on the order parameter given in this relationship can be measured in a spectroscopy
experiment, giving an indirect method to determine the temperature dependence of .

Potential energy of the crystal


Before we consider the phonons further, we need to account for the change in potential energy of the
crystal due to the phase transition. We write this as

E
= 2 + 4 3 02 = 2 /4

(5.20)

An example of this energy function is shown in Figure 5.24, where calculations of E( ) for quartz are
fitted by a function of the form of Equation (5.19).

1.0
Energy (kJ/mol)

1
1
E ( ) = 2 2 + 4 4
(5.19)
2
4
We will show below that the coefficient 2 is equivalent to the square of the harmonic frequency of the
soft mode. The equilibrium value of the order parameter at zero temperature can be obtained from
minimisation of the potential energy:

0.5
0.0
-0.5

0.0

-1.0
-1.5
-1.0

Phonon free energy


The phonon free energy has the form
Fph = kB T ln 2 sinh h q /2kB T




(5.21)

We will make the high-temperature approximation, kB T > h q , to give



Fph = kB T ln h q /kB T
q

= 3RT ln (h h i /kB T )

(5.22)

We have replaced the sum over all modes by average values, so that we have from equation (5.18)

h i2 = h i2 +

1
hi 2
2

(5.23)

0.0
Order parameter

1.0

Figure 5.24: Variation of the potential energy of quartz


with changing order parameter. The points are energies calculated using a pair potential model, and the
curve is described by equation 5.19 with fitted values
of 2 and 4 .

Chapter 5. Displacive phase transitions and soft modes

118

The phonon free energy can be written as a Taylor expansion in the order parameter about the point
= 0:
2 Fph

1
Fph ( ) = Fph ( = 0) + 2
2

+...

(5.24)

=0

There is no linear term because F/ = 0 at equilibrium, and we will not consider the higher-order
terms. We have the following differentials:
Fph
3RT h i
=

h i
"

 #
2
Fph
3RT 2 h i
1
h i 2
=

2
2
h i
h i

(5.25)
(5.26)

From above we have the following differentials of the frequencies:


h i
hi
=

2 h i

2 h i
hi
hi h i
=

2
2 h i 2 h i2


h i
=0
=0

2 h i
hi
=

2
2 h i

=0

(5.27)
(5.28)

As a result we can write



2 Fph

2

=
=0

3 hi RT

(5.29)

2 h i2

and
Fph ( ) = Fph ( = 0) +

3 hi RT
4 h i

(5.30)

5.4. Lattice dynamical theory of the low-temperature phase

119

Full free energy and the Landau free energy function


We now combine the phonon free energy (equation 5.30) with the potential energy (equation 5.20) to
obtain
F ( ) = Fph ( ) + E( )

= Fph ( = 0) +

3 hi RT
4 h i

1
= Fph ( = 0) +
2
1
= Fph ( = 0) +
2

2 21 2 2 + 14 4 4

3 hi RT
2 h i2
!
3 hi R
2 h i2

2 + 14 4 4

22 h i2
T
3 hi R

!
2 + 41 4 4

(5.31)

This looks exactly like the Landau free energy function for second-order phase transitions,
F ( ) = F ( = 0) + 12 a( T Tc ) 2 + 14 b 4
with
Tc =

22 h i2
3 hi R

(5.32)

(5.33)

Since 2 is the negative of the square of the harmonic frequency of the soft mode, this equation for
Tc is exactly the same as the equation we obtained earlier from the model of the soft mode. We have,
however, achieved more than simply replicating the earlier result. We have shown how Landau theory
falls out of the lattice dynamics, which gives some justification for the use of Landau theory in the study
of displacive phase transitions. Furthermore, we have also obtained microscopic interpretations of the
coefficients in the Landau function.

Low temperature behaviour


The high-temperature approximation is clearly not appropriate at low temperatures. In principle we can
take the complete free energy function of equation 5.21 and expand this in terms of 2 as we have done
above. We would end up with something similar, albeit with more complicated differentials, and the
final equation would deviate from the Landau function at low temperature. Rather than do a complete

Chapter 5. Displacive phase transitions and soft modes

120

analysis of the low-temperature behaviour, we will restrict ourselves to T = 0 K. Here the free energy is
the potential energy plus the zero point phonon motion, giving
F (, T = 0) = 32 NA h h i 12 2 2 + 14 4 4

(5.34)

To obtain the equilibrium value of at T = 0 K we calculate the differential of F, using the results for
the differential of h i given above:

h i
F
3
2 + 4 3
= NA h
T =0
2



hi
= 3NA h
2 + 4 3
(5.35)
4 h i
Setting the differential to zero allows us to obtain the equilibrium zero-temperature value of the order
parameter, 0 :


hi
2 + 4 3 = 0
3NA h
4 h i

3NA h
02 = 2
(5.36)
4
44 h i

(Order paramater)2

1.0
0.8

Because we are not expanding around = 0, the equation for 0 contains h i rather than h i. This
result shows that the effect of zero-point motion is to lower the zero-temperature value of below that
given by minimisation of E( ) alone.
An example of the complete quantum mechanical version of this theory fitted to experimental data
for the order parameter is shown in Figure 5.25. This clearly shows a linear dependence of 2 close
to Tc , but with a marked curvature away from this linear relation at low temperatures and giving the
expected thermodynamic behaviour /T = 0 at T = 0.

0.6
0.4
0.2
0.0

100

200 300 400


Temperature (K)

500

600

Figure 5.25: Temperature dependence of the square


of the order parameter associated with the displacive
phase transition in the aluminosilicate CaAlSi3 O8 . The
curve represented a fit of the theory of this chapter to
the data, taking account of the quantum terms at low
temperature.

5.5

Phase transitions, soft modes, and structure flexibility: the Rigid Unit Mode model
The topology of framework structures
We have looked at a number of structures that can be defined as networks of linked polyhedral groups
of atoms. For example, the perovskite SrTiO3 can be described as a network of corner sharing TiO6

5.5. Phase transitions, soft modes, and structure flexibility: the Rigid Unit Mode model

121

octahedra, and the displace phase transition involves cooperative rotations of these octahedra without
significant distortions. Another example is silica, where we have met the quartz and cristobalite forms;
these any many related materials can be described as networks of corner-sharing SiO4 tetrahedra.
It is instructive to take one particular extreme view of network structures, namely that the forces
that hold the structural polyhedra together are very strong, and all other forces are relatively weak. In
the case of silica, this is not unreasonable. The SiO bond is known to be the strongest force in the
silica structure (stretching frequencies are are around 40 THz), and the forces associated with bending
the OSiO bond are about one quarter of the strength of the SiO bond (frequencies are around 20
THz). The forces that oppose pivoting of two SiO4 tetrahedra about the common oxygen atom are much
weaker (frequencies of these modes are around 1 THz). In the case of SrTiO3 , forces connected with
distortion of the TiO6 octahedra are not as extreme, but nevertheless it can be instructive to think of the
forces opposing rotations of two linked TiO6 octahedra as being relatively weak, and indeed inelastic
neutron spectroscopy confirms this.

Flexibility of framework structures


We can analyse the flexibility of framework structures using engineering principles first articulated by
James Clerk Maxwell. For any framework of connected objects, we count the number of constraints,
c, and the number of degrees of freedom, f , and if f > c the framework is flexible, and if c > f the
framework is rigid.
For the case of connected tetrahedra, for any tetrahedron f = 6, due to both translational and
rotational degrees of freedom. We count constraints by thinking of the tetrahedra as separate objects,
but subject to the constraint that the corner of one tetrahedron must have the same x, y, z coordinates
as of the linked corner. Thus we have 3 constraints per corner, which have to be shared by the two
tetrahedra. Each tetrahedron has 4 corners, and so c = 4 3/2 = 6. Thus we end up with the result
that f = c, so that each silica can be said to be exactly balanced between being rigid and flexible.
This approach to counting has not taken account of symmetry, and when we do so we find that
some of the constraints are degenerate. Consider a ring of 4 squares connected in three dimensions.
Each square has 3 degrees of freedom, so f = 12. The ring consists of 4 linked corners, so we have
c = 8. Hence f c = 4, corresponding to the number of ways we can deform the square; these are 2
translations, 1 rotation, and 1 shear. It is the shear mode that is interesting in this context. Now we
consider how to extend this structure. We add 2 additional squares to one side. This adds another 6

122

Chapter 5. Displacive phase transitions and soft modes

degrees of freedom and another 6 constraints. So we can keep adding squares in the same direction to
form a very long chain of connected 4-membered rings of squares. We denote the number of squares
along the chain as Nx , so that f = 6Nx and c = 6Nx 4.
Next we add a row of squares in the second dimension. This will add another 3Nx degrees of
freedom. Taking our prior argument at a simple level, we might imagine that we also add another
4Nx 4 constraints. This is alarming, because it implies that the number of constraints is quickly going
to exceed the number of degrees of freedom. However, there is something wrong here. So instead of
adding the whole row at once, let us add 2 squares from this row. To enable the corners to match, we
do need to add 6 constrains (balancing the additional number of degrees of freedom). Now we add
one more square to the side. This adds three degrees of freedom, but one fewer constraint that we
might think. So we have one constrain that pins the squares corner to the row below, but to join up the
new 4-member ring we only need one constraint instead of two, namely one that defines the angle or
rotation of the square required to form the corner. And as we complete the row, each new square adds
exactly 3 to both f and c.
In the end, for our array of squares, we have f c = 4 always, of which one of these degrees of
flexibility is an internal distortion of the structure. If we have an infinite network, we actually loose
the rotational of the whole network. The two translations correspond to the 2 acoustic modes at q = 0,
and the internal mode corresponds to the opposite rotations of the squares. But in the infinite network,
motions that cause a volume change will cost energy, so that internal mode is only low energy in the
fully extended network.
This argument can be extended to the three-dimensional network of connected octahedra as in the
perovskite structure. It is found that there are internal deformations possible, namely the opposite
rotations of connected octahedra about the x, y and z axes, but with each layer acting independently.
If we think of what this means in reciprocal space, we have lines of wave vectors along the edges of
the Brillouin zones (i.e. the lines connecting points ( 12 , 12 , 12 ) and symmetrically related lines. These
correspond to low-frequency phonons. Indeed, the phonon at q = ( 12 , 21 , 12 ) is the soft mode for the
cubictetragonal phase transition in SrTiO3 .
We can play the same trick for silica, but the flexibility we get will depend on the details of the
topology and the symmetry of the structure. What is found is that often there are whole planes in
reciprocal space where the phonon eigenvectors correspond to rigid-body rotations of the tetrahedra
without any distortion.

5.5. Phase transitions, soft modes, and structure flexibility: the Rigid Unit Mode model

123

We call these phonons Rigid Unit Modes, abbreviated as RUMs.

Rigid unit modes as soft modes for displacive phase transitions


Investigation of a series of displacive phase transitions in framework structure has shown that in all
cases the soft mode is indeed a RUM. It is the existence of the network flexibility that leads to RUMs
that allows phase transitions to exist.
What is not answered by this statement is the question of why the high-symmetry phase is not
stable. The answer to this is easy to imagine (and demonstrated by calculation). In most cases, the
high-symmetry phase is the fully-expanded version, and distortion (as noted in the example of squares
above) will lead to a volume reduction. Since the long-range forces in most materials (electrostatic and
dispersion) are attractive, the energy is lowered by reducing the volume. Of course, the reduction in
volume will at some point be offset by overlap of atoms leading to repulsion, but nevertheless, the gain
in negative energy caused by reducing the volume, albeit at the expense of lowering symmetry, will win.
Thus we see that the possibility of low-energy deformation of the structure as allowed by RUMs and
the gain in energy caused by the accompanying volume reduction will give rise to the lower-symmetry
structure as the low-temperature structure. The fact that there are many RUMs leads to the question
of why one RUM is the soft mode and not others. The answer will be subtle and specific to different
systems, but we can identify two issues. The first is whether we get large or small volume reductions
that accompany the RUM deformation (and hence greater or lower changes in energy), and whether the
necessary twisting of the linkage between two tetrahedra is large and whether it costs energy.
This argument explains why we do not see displacive phase transitions in more rigid systems, such
as systems composed of edge and face sharing octahedra. These systems cannot support RUMs, and
if the structural polyhedra are stable there will not be a displacive phase transition associated with
distortions of the polyhedra. Examples of stable systems are MgO and most other rocksalt structures,
TiO2 , Al2 O3 and related systems.

Why is there a phase transition?


The question comes down to asking why, given that the low-symmetry structure is energetically
favourable, why do we get a phase transition to the higher-symmetry structure? The answer to this
has to come down to the balance between energy and entropy the balance between E and TS in the

124

Chapter 5. Displacive phase transitions and soft modes

equation for free energy. Thus the question we have to ask is why is there an additional entropy in the
high-symmetry phase?
We have previously remarked that symmetry has an effect on the flexibility. Experimentally this is
seen in the data for the soft mode for SrTiO3 , where we see that the soft mode associated with the phase
transition increases considerably (and gradually) on cooling below Tc . This means that higher-symmetry
phases have more phonons that can be described as RUMs. Since RUMs have low frequency, they
have large amplitude, which means that their contribution to entropy is larger. Thus we can identify
the entropic contribution to the phase transition. Experimental studies have shown how the RUM
contribution to the thermal motion in silica phase increases on heating through the phase transition.

What does the structure of the high-temperature phase look like?


The high-symmetry phase looks, on average, like the crystal structure given by diffraction. However,
there will be an increase in thermal motion associated with the low-frequency RUMs, causing significant
rotations of the structural polyhedra. A snap shot of the atomic positions of the high-temperature phase
will show quite a lot of disorder associated with flexing of the structural network.
Experimental evidence points to the fact that this structural order does not constitute atoms jumping
between well-defined positions as you might get in an orderdisorder phase transition. Such a picture
is popular in some quarters, but is not sustained by evidence nor is it consistent with the network
flexibility. Instead, the polyhedra continue to rotate about their average orientations, albeit with large
amplitude.

What determines the value of the transition temperature?


Earlier in this chapter we have identified the displacive phase transition temperature as arising from
some fundamental parameters. We can express the RUM model in similar terms. We associate a variable
with each polyhedron, describing most probably a rotation. We consider each polyhedron to be linked
to its neighbour, with an energy cost associated with the rotations of the form + J i j , the sign being
chosen in order that the energy is lowered for opposite rotations.
We now associated an energy that drives the transition to have the form of a single-particle potential:
1
1
E( i ) = 2 2i + 4 4i
2
4

(5.37)

5.6. Summary

125

The quadratic term arises in part from the associated energy gain in lowering volume, and is propositional to the square of the soft mode frequency at the critical wave vector. The quartic term has to be
included in order for the system not to want to distort without end! One can think that the primary
origin of the quartic term will arise from the repulsive interactions.
Combining these two we have


1
1
(5.38)
E = 2 2i + 4 4i + J i j
2
4
i
i.k
In fact this is a well-studied model with well-established features. First we note that in a way that maps
onto our original theories of Tc we find that Tc J, where J is playing the role of 2 in the earlier
equations. So J enters the equation as the coupling between neighbouring tetrahedra, and therefore sets
the scale for the phonon frequencies. The constant of proportionality will be set by the fine details of the
crystal structure, as an integration over the phonon density of states. As before, the proportionality will
also involve the ratio 2 /4 . The point is that the transition temperature is set by the scale of J, which in
turn is set by the stiffness of the structural polyhedra.
One issue that has been addressed through variants of this simple model are the links between
orderdisorder and displacive transitions. Within the same model we have both limits. If the potential
energy wells defined by equation 5.37 are deep as compared to the energy of coupling set by J, the
transition is very much like an orderdisorder transition, whereas in the other extreme the transition is
very much like a displacive phase transition. The difference is characterised by the ratio of the height of
the barrier in the double well, E1 = 22 /24 , to the coupling energy when all the rotation angles are
at their minimum value, namely when = 2 /24 giving E2 = zJ2 /24 , where z is the number of
connected neighbours. This ratio is equal to
s=

E1

= 2
E2
2zJ

(5.39)

Based on what we have discussed, we expect J  2 , and hence we have demonstrated that RUM phase
transitions are expected to behave as displacive phase transitions, and hence all of the previous results
of this chapter apply exactly.

5.6 Summary
The key results of this chapter are

126

Chapter 5. Displacive phase transitions and soft modes

1. We have characterised a number of examples here and before, and the examples taken together
illustrate the main features of displacive phase transitions, such as the breaking of symmetry
through small displacements of atoms, the existence of both first and second-order transitions,
and the general coupling to spontaneous strains.
2. The Lyddane-Sachs-Teller relation predicts that a divergence of the dielectric susceptibility at a
ferroelectric phase transition will be accompanied by the frequency of a transverse optic mode
falling to zero value.
3. The soft mode is a phonon at the critical wave vector for a phase transition whose frequency
falls to zero on cooling towards the transition temperature. Soft modes are found to accompany
displacive phase transitions. We can have soft acoustic modes as well as soft optic modes.
4. Soft modes emerge from lattice dynamics extended to include fourth-order anharmonic terms.
The anharmonic terms act to stabilise the high-symmetry structure at temperatures above Tc .
5. From the soft mode theory we can calculate the value of Tc from fundamental parameters.
6. A model for the low-temperature phase based on the phonon free energy with an anharmonic
double-well function for the potential energy gives the temperature-dependence of the order
parameter, with the same expression for Tc .
7. The thermodynamic treatment of the low-temperature phase gives a final free energy that looks
like the standard Landau free energy, and as a result we show firstly that Landau theory should
be a good representation of the thermodynamics of displacive phase transitions, and secondly we
can related the parameters of the Landau function to fundamental parameters.
8. We introduced the Rigid Unit Mode model to understand phase transition in network structures,
and thereby were able to understand the origin of phase transitions in these systems.

5.7

Short questions to test your basic knowledge


1. Give the main characteristics of a displacive phase transition, with examples.
2. Write the Lyddane-Sachs-Teller relationship. What is the key implication for a ferroelectric phase
transition?

5.8. Longer questions

127

3. Why is the soft mode so called? What is the characteristic variation of the soft mode with
temperature, and how does this link to a displacive phase transition?
4. Sketch dispersion curves to show how the behaviour of the soft mode with temperature for a
ferroelectric phase transition.
5. Sketch dispersion curves to show how the behaviour of the soft mode with temperature for the
displacive phase transition in SrTiO3 .
6. Sketch dispersion curves to show how the behaviour of the soft mode with temperature for a
ferroelastic phase transition.
7. What do we mean by the term rigid unit mode?

5.8 Longer questions


1. In this question we will look at the possible origin of structural instability that could drive a phase
transition. We consider a perfect TiO6 octahedron.
(a) Show that as far as the Coulomb energy is concerned, the mid-point of an octahedron is a
maximum of the potential energy. One approach is to consider the two charged anions on
the opposite vertices of an octahedron, with a charged cation in the centre, and consider this
as a one-dimensional problem. Denote the distance from each anion to the centre as r, and
the displacement of the cation from the centre as . Hence show the energy will be equal to
Q+ Q
E=
4 0

1
1
+
r+ r


(5.40)

Show either directly by differentiation, or via a Taylor expansion, that energy has is a
maximum when = 0.
(b) Repeat this calculation but now considering a pair-wise repulsive term between cation
and anion in addition to the coulomb term. If this pair-wise repulsion is described by
V (r ) = C/r12 , repeat the analysis to show that the energy has its minimum value when = 0
for all values of C.

128

Chapter 5. Displacive phase transitions and soft modes


(c) Combine these two energies to show that the centre of the octahedron turns from a minimum
to a maximum when
156C
| Q+ Q |
> 12
2 0 r
r

(5.41)

At this point you will expect to see a ferroelectric phase transition.


2. The vibrations in a given material have an average frequency of 10 THz. The material undergoes a
phase transition, as a result of which the vibrational frequencies increase by an average of 0.2 THz
on cooling from the disordered state (order parameter = 0) to low temperatures (where = 1
by definition).
(a) Use the formula 2 = 02 + 12 2 (remembering the factors of 2) to calculate a value for the
parameter .
(b) If the energy change through the phase transition gives 2 = 1 kJ/mol, obtain a value for the
transition temperature Tc .
3.

(a) Describe the soft mode theory of phase transitions of ferroelectricity, explaining how it was
anticipated by the Lyddane-Sachs-Teller relation, outlining the link between the soft mode
and the static deformation in the low-temperature phase, and explaining how the theory is
applicable to other types of displacive phase transition.
(b) Explain how the lattice dynamical theory of phase transitions provides a link between the
phase transition temperature Tc and microscopic parameters that relate to the interactions
between atoms.
(c) Explain why the soft mode frequency increases rapidly on cooling below Tc .
(d) To what extent does the theory of displacive phase transitions covered in the lectures address
the open questions given earlier in the chapter?

Chapter 6

O rder disorder phase transitions

6.1 Orderdisorder phenomenology


Orderdisorder phase transitions
In many aluminosilicate minerals with crystal structures containing AlO4 and SiO4 tetrahedra there
may be two structural states, with the Al and Si cations ordered in definite tetrahedral sites in the
low-temperature phase, and disordered across the tetrahedral sites in the high-temperature phase.
One well-studied example is cordierite, Mg2 Al4 Si5 O18 , Figure 6.1. The orthorhombic phase has a
well-ordered distribution of Al and Si cations in five distinct tetrahedral sites, but the hexagonal form,
which is the phase that initially develops from the glass or liquid, has only two distinct tetrahedral sites
and Al/Si disorder. The process of ordering the Al and Si cations reduces the symmetry of the crystal.
In principle all aluminosilicate minerals can have ordered and disordered phases. In plagioclase feldspar,
CaAl2 Si2 O8 , both disordered and ordered forms are found as naturally occurring states, although it is
believed that the ordering temperature lies above the melting point. On the other hand, in the alkali
feldspar albite, NaAlSi3 O8 , the ordering temperature is as low as 500 K. In leucite, KAlSi2 O8 , and
gehlenite, Ca2 Al2 SiO7 , the ordering temperature is so low that ordered states have yet to be found.
Experimental data for cordierite and other aluminosilicates show that the onset of long-range order
occurs at a fixed point, whether in temperature or annealing time, and the degree of long-range
order changes continuously with temperature or time in the ordered state. However, there is also
a certain amount of short-range order in the disordered phase. This is shown particularly by NMR
measurements, which show that before the onset of long-range order there is the development of
non-random arrangements on a local scale. This short-range order is often described as Al-avoidance,
which is the tendency of the Al cations to avoid having neighbouring sites. Although Al/Si ordering is
the most commonly studied ordering process in minerals, it is not the only one. Ordering of atoms is
129

Figure 6.1:
Ordered structure of cordierite,
Mg2 Al4 Si5 O12 , showing the SiO4 (blue) and AlO4
(green) tetrahedra.

Chapter 6. Orderdisorder phase transitions

130

common in metal alloys (text book examples are the Cu4 x Alx systems), and most of the theoretical
developments have been motivated by the study of alloys. We can also cite the ordering of cations such
as Mg2+ and Ca2+ in dolomite, CaMg(CO3 )2 , as well as in some silicates, and the ordering of transition
metal ions and vacancies in sulphides.

Solid solutions and non-convergent ordering


Although the formation of long-range order is a dramatic process that has been studied in its own
right, the existence of short-range ordering in systems that do not have phase transitions is of equal
importance, not least because it can significantly affect the thermodynamic properties. One example
is the solid solution between different garnet phases, such as pyrope, Mg3 Al2 Si3 O12 , and grossular,
Ca3 Al2 Si3 O12 (Figure 6.2). In this case the Mg2+ and Ca2+ cations form arrangements that are different
from a random distribution.
Non-convergent ordering occurs when the onset of long-range order does not precipitate a phase
transition. One important example is spinel, where ordering of cations over the [4] and [6] sites does
not change the symmetry, although there are many aspects of the ordering process that have similarities
to conventional ordering phase transitions.
Figure 6.2: Crystal structure of a garnet, M3 Al2 Si3 O12 ,
showing the SiO4 (blue) and AlO4 (green) octahedra.
The M2+ cations (yellow spheres) are typically Mg2+ ,
Ca2+ or Fe2+ . Mixtures of cations are easily achieved
in this structure.

6.2

Mean-field theory of orderdisorder phase transitions: the BraggWilliams model


The Order Parameter
Consider the case of a system in which there is ordering of the positions of equal numbers of two
cations, A and B, and where in the disordered phase the cations occupy sites of equivalent symmetry.
In the ordered phase we can define the quantity that gives the fraction of A atoms and B atoms on the
right sites (these fractions will be the same). When there is complete order = 1, and when there is
complete disorder = 1/2. Conventionally it is preferred to quantify the degree of order with the order
parameter , whose value varies between 1 for complete order and 0 for complete disorder. For the
example considered here, is defined as
= 2 1

(6.1)

Most theories of the ordering phase transition are based on the idea that contains all the information
about the degree of order, and that the thermodynamic functions are analytic functions of . In this

6.2. Mean-field theory of orderdisorder phase transitions: the BraggWilliams model

131

section we show one such theory, that due to Bragg and Williams.
In the theory we develop here, the free energy is separated into two parts, the first giving the internal
energy of ordering, and the second giving the entropy. In both parts there is an essential approximation
that the degree of order can accurately be represented by the order parameter on a local scale as well as
the scale of the whole system.

Configurational entropy
We consider the entropy associated with ordering. The number of ways of distributing N A atoms over
N A sites follows the standard formula:
A =

N!
( N )! ((1 ) N )!

(6.2)

The same formula also gives the number of ways of distributing N B atoms over N B sites, B .
The entropy is given by the standard Boltzmann formula:
S = kB ln (A B )


N!
= 2kB ln
( N )! ((1 ) N )!

= 2kB [ln( N!) ln (( N )!) ln (((1 ) N )!)]

(6.3)

Mathematical identity
In these lectures we often need to evaluate pairs of equations of the form
=

N!

j N !


S/kB = ln = ln N! ln j N !

(6.4)

where N is a large number, and, by definition,

j = 1

(6.5)

We can use Stirlings formula for large N,


ln( N!) N ln N N

(6.6)

Figure 6.3: Atomic configurations corresponding to


top) complete order, centre) partial long-range order
with additional short-range order, bottom) no longrange order or short-range order.

Chapter 6. Orderdisorder phase transitions

132
The equation for the entropy can then be rewritten as


S/kB = N ln N N j N ln j N + j N
j

= N ln N N j ln j N j ln ( N )


= N j ln j

(6.7)

where a number of terms cancel out.

Configurational entropy with long-range order


From the equations above, the entropy reduces to
S = 2NkB [ ln + (1 ) ln(1 )]

(6.8)

We can then rewrite S in terms of , which after rearrangement gives


S = NkB [2 ln 2 (1 + ) ln(1 + ) (1 ) ln(1 )]

(6.9)

Calculation of the internal energy


We now consider the internal energy, and for simplicity assume that changes in the enthalpy on ordering
affect the nearest-neighbour interactions. Let us also assume that the atoms on each A site interact with
z neighbours on the B sites, and vice versa. For the interaction between one A site and a neighbouring B
site, we have the following probabilities P that the bonds will be of each type:
PAB = 2
PAA = (1 )

PBA = (1 )2
PBB = (1 )

(6.10)

If we represent the energy of the different types of bonds as EAA , EAB and EBB , the internal energy of
the system is given as
E = Nz [ PAA EAA + PBB EBB + ( PAB + PBA ) EAB ]
h


i
= Nz (1 )( EAA + EBB ) + 2 + (1 )2 EAB

(6.11)

6.2. Mean-field theory of orderdisorder phase transitions: the BraggWilliams model

133

where N is the number of A atoms and the number of B atoms (by convention here we denote N as
the total number of atoms of any sort, although in the examples of orderdisorder phase transitions it
is easier to divide the system into two sublattices). From the definition of the order parameter given
earlier we have
1 = 21 (1 )
= 21 (1 + )


(6.12)
1
2
2
+ (1 )2 = 21 1 + 2
(1 ) = 4 1
These expressions allow us to write the internal energy as a function of :



i
Nz h
1 2 ( EAA + EBB ) + 2 1 + 2 EAB
4
i
Nz h
=
( EAA + EBB + 2EAB ) 2 ( EAA + EBB 2EAB )
4
Nz 2
=
J + E0
4

E=

(6.13)

where we have defined the exchange energy J and the constant energy E0 as
;

E0 =

Nz
( EAA + EBB + 2EAB )
4

0.5

(6.14)

0.4

J is the energy required to replace two AB bonds with an AA and a BB bond, and we will see that it
plays a pivotal role in theories of ordering.

Free energy

0.3

Free energy

J = EAA + EBB 2EAB

0.2

T > Tc

0.1
0.0

The free energy is obtained by combining the equations for internal energy and entropy:
F = E TS = F0

Nz 2
J + NkB T [(1 + ) ln(1 + ) + (1 ) ln(1 )]
4

T < Tc

0.1

(6.15)

where F0 contains the terms that do not depend on . This function is shown for high and low
temperatures in Figure 6.4. At high temperature the function has a minimum at = 0, implying that
the equilibrium state is completely disordered. At low temperatures there are two equivalent minima at
non zero values of , so that the equilibrium state has a certain degree of order. Although not shown
here, the free energy curves evolve continuously between the two types of function as the temperature
is changed.

0.2
0

0.2

0.4
0.6
Order parameter

0.8

Figure 6.4: Free energy as a function of order parameter from the BraggWilliams model for temperatures
above and below the phase transition. The minimum
in the free energy gives the equilibrium value of the
order parameter at each temperature.

Chapter 6. Orderdisorder phase transitions

134

Temperature-dependence of thermodynamic variables


The equilibrium value of at any temperature is given by minimisation of the function for F:


1+
F
Nz
=
J + NkB T ln
=0

2
1


1+
2k T
= B ln
zJ
1

(6.16)

The equation for has to be solved self-consistently. The transition temperature Tc is given by the point
at which the free energy function changes from having a minimum to a maximum at = 0, which is
defined as the point at which the second derivative vanishes:


2 F
Nz
1
1
=
J + NkB T
+
2
1+
1
2

2
Nz
F
=
J + 2NkB T

2
2
=0
Nz
J + 2NkB Tc = 0
2
zJ
Tc =
4kB

(6.17)

This allows us to rewrite the equation for as






T
1+
Tc
=
ln
= tanh
2Tc
1
T

(6.18)

(6.19)

At temperatures close to Tc , will be small, so we can use the expansion of the logarithms:
ln(1 + x ) x

x2
x3
+

2
3

(6.20)

We can therefore expand the equation for as

T
Tc


+

3
3

3
( Tc T )1/2
Tc

(6.21)

This functional form for at temperatures close to Tc is the now-familiar mean-field theory result.

6.2. Mean-field theory of orderdisorder phase transitions: the BraggWilliams model

135

The susceptibility for T > Tc is obtained from the second-differential of F as we have previously met:



2 F
Nz
Nz 4kB Tc
=
J + 2NkB T =
+ 2NkB T = 2NkB ( T Tc )
(6.22)
2
2
z
Q2 T >Tc
 2  1
F
1
T >Tc =
=
(6.23)
( T Tc )1
2
2NkB
Q T >Tc
Again, we have an exponent of 1 as is characteristic of mean-field theories, diverging on cooling
towards Tc . The susceptibility is not easy to measure for site ordering phase transitions. In principle it
could be obtained from diffraction measurements on equilibrated samples. The susceptibility is much
easier to measure in magnetic ordering phase transitions, but it is a quantity that is easy to calculate in
computer simulations.

1.0

Critique of the use of the mean-field approximation


Order parameter

The mean-field theory of atomic ordering developed above, which is commonly known as the Bragg
Williams model, is very popular as a first step in the study of orderdisorder phase transitions,
particularly because it captures many aspects of the phenomena and requires only one adjustable
parameter, J or Tc . We can demonstrate how well the model works by comparing the results of
the theory with computer-generated data using the Monte Carlo method (described below). Our
comparison is carried out using the kalsilite structure, where we have equal numbers of Al and Si
cations ordering over sites with four neighbouring sites, and we have only nearest-neighbour interactions.
The comparison is given in Figure 6.5.
The important point to note from Figure 6.5 is that the BraggWilliams model predicts a value for
the transition temperature that is too high. This is a common problem with the BraggWilliams model.
The reason for this is that the BraggWilliams model assumes that short-range order, as given by the
numbers of each type of bond, is completely determined by the long-range order. Therefore the only
way to reduce the energy is to generate long-range order. However, if a certain degree of short-range
order can develop independent of the existence of long-range order, there is a reduced need for the
system to undergo the phase transition. As a result, the phase transition occurs at a lower temperature,
at the point at which the system can no longer develop further short-range order without driving
long-range order. This temperature is lower the smaller the number of neighbours, and for chains of
atoms (i.e. for a one-dimensional system) where each atom has only two neighbours ordering only sets

0.8
0.6

BW

0.4

CVM

0.2
0.0
0

0.2

0.4
0.6
0.8
Temperature / J

1.2

Figure 6.5: Temperature-dependence of the order parameter for ordering of equal numbers of two types of
atoms, each with four neighbours, and bond-ordering
energy J as defined in the text. The curves labelled BW
and CVM are obtained from the BraggWilliams and
bond-CVM models respectively. The points are from
Monte Carlo simulations (the tail of the simulation
data above the transition point is due to the finite size
of the simulation).

136

Chapter 6. Orderdisorder phase transitions

in at zero temperature. The Cluster Variation Method (discussed later) is an attempt to incorporate
short-range order into thermodynamic models.
Another important problem with the BraggWilliams model, which is common with many meanfield theories, is that the functional form of the temperature-dependence of the order parameter close to
the transition point is not found in experiments. Instead, the exponent is usually lower than 1/2. This
arises from the neglect of special sort of fluctuations, namely the growth of finite-sized clusters of the
ordered state. It should be appreciated that these so-called critical point fluctuations are likely to be
important in understanding these orderdisorder phase transitions, but are less important in displacive
phase transitions where the necessary conditions for the application of mean-field theories often hold.
Despite the problems of the BraggWilliams model outlined here, the basic approach is still often
used in the study of phase transitions. Part of the appeal is that the basic shape of the plot of
order parameter vs temperature looks about right, and the transition temperature can be treated as a
single adjustable parameter in fitting against experimental data. In using the equation to describe the
experimental data this approach may be as good as any other, but it is clearly dangerous to attempt to
use it for quantitative analysis. And actually it is a little churlish to be too critical when Bragg and
Williams proposed their model it was a significant breakthrough in the understanding of the mechanism
of ordering in alloys, and subsequently stimulated the whole field.

6.3

Computational methods
Formalism of interaction energies: case of equal numbers of two types of atoms
We consider the case of two types of atoms, A and B, and first take the case where there are equal
numbers of both types. We also consider the case where there is only one type of site for these atoms,
which has z neighbouring sites of the same type. We now have N atoms in total.
For any atom of type A, the probability that one specific neighbouring atom is also of type A is PAA ,
and the probability that the same neighbouring atom is of type B is 1 PAA . For any atom of type B,
the probability that one specific neighbouring atom is also of type B is PBB . Noting that for N atoms in
a periodically repeating sample the number of bonds will be Nz/2, the number of bonds of each type is

6.3. Computational methods

137

given as
1
NzPAA
4
1
= Nz(1 PAA )
2
1
= NzPBB
4

NAA =

(6.24)

NAB

(6.25)

NBB

(6.26)

We obtain an expression for PBB by adding up all the different types of bonds:
1
1
1
1
Nz = NzPAA + Nz(1 PAA ) + NzPBB
2
4
2
4

(6.27)

PBB = PAA

(6.28)

from which we have


We now denote the energies of the different bond types as EAA , EAB and EBB . The energy obtained
from the number of bonds of each type follows as
1
1
1
NzPAA EAA + Nz(1 PAA ) EAB + NzPBB EBB
4
2
4
1
1
= NzPAA ( EAA + EBB 2EAB ) + NzEAB
4
2
= NAA ( EAA + EBB 2EAB ) + E0

E=

(6.29)

where we establish the practice of subsuming all constant terms into one overall constant E0 . We see
that the energy can be expressed in terms of the number of bonds of one specific type, in this case the
AA bonds, and the energy required to exchange the atoms from two AB bonds to form an AA bond
and a BB bond.

Representation as a spin model


We can define a variable for each site, SA , such that SA = 1 if the site is occupied by an A atom, and
SA = 0 if it is occupied by a B atom. We can similarly define SB = 1 SA . The total number of A atoms
is given as a sum of this variable over all sites,
NA =

SAj
j

(6.30)

Chapter 6. Orderdisorder phase transitions

138
and the number of AA bonds can be written as
NAA =

SAj SkA

(6.31)

h j,ki

where j and k denote neighbouring sites, and the sum does not involve counting the same bond twice.
Accordingly we can rewrite the equation for the energy as
E = ( EAA + EBB 2EAB )

SAj SkA + E0

(6.32)

h j,ki

It is more customary to define a spin variable for a site, such that = +1 if the site is occupied by an
A atom, and = 1 if the site is occupied by a B atom. It follows that
SA
j =


1
1 + j
2

SBj =


1
1 j
2

(6.33)

and the energy can be written as


E=

1
( E + EBB 2EAB ) j k + E00
4 AA
h j,ki

(6.34)

where the constant term now includes additional constant terms involving sums of j values (these
sums must remain constant for orderdisorder phase transitions because the numbers of each type of
atom must remain constant).
Using the previous definition of the exchange energy,
J = EAA + EBB 2EAB

(6.35)

we can write the energy equation in terms of spins as


E=

1
J
j k + E00
4 h
j,k i

(6.36)

It has almost been assumed that we have been considering nearest-neighbour interactions, but nothing in
the theory has actually depended on this assumption. Therefore our main result can easily be generalised
to consider different types of neighbouring interactions, for example second-nearest neighbours, by the
simple generalisation
1
(6.37)
E = Jjk j k + E00
4 h j,ki
where the energy of a bond now depends on the bond type.

6.3. Computational methods

139

Calculation of ordering energies

H=

1
Jjk j k + j j
4 h
j
j,k i

(6.38)

Again, the spin variables j are defined as having value +1 if the site is occupied by one type of atom
(e.g. Al) and value 1 if it is occupied by the other type of atom (e.g. Si). The double summation in the
first term involves summing over all pairs of atoms, but ensuring that each pair is only counted once.
The exchange constant Jjk depends on the type of sites, whether nearest-neighbour or more distant
neighbours. The first term in the spin Hamiltonian is developed from the formalism of counting bonds
as given above. The second term accounts for the existence of different types of sites, and j acts as a
chemical potential. For example, in cordierite there are two distinct tetrahedral sites in the disordered
phase, one (T2) within 6-membered rings of tetrahedra, and one (T1) within 4-membered rings that link
together the 6-membered rings. One can assume that the energy of an Al or Si atom will be different if
it occupies one or other of these sites, and it is this energy difference that is represented by the chemical
potential term. The two tetrahedral sites in cordierite are shown in Figure 5, where we also show some

-25

Lattice energy (eV)

Central to any computational model of ordering is to be able to calculate the exchange energy J. The
usual approach is to set up a periodically-repeating group of unit cells of the crystal and to calculate
the energies of a number of configurations formed with different distributions of the ordering atoms.
Since the ordering atoms will have different sizes it is necessary to allow the structure to relax to the
lowest energy state in each case using lattice energy minimisation methods. Typically around 100
configurations with different amounts of order are required for this procedure to give reliable estimates
for J. This is a demanding task and so it is often only practical to use empirical model interatomic
potentials, although we have some experience of using electronic structure calculations for this task.
For a case where we consider only nearest-neighbour interactions to be important, the exchange
energy can be extracted from the slope of a plot of energy against the number of bonds of one type. For
example, for an aluminosilicate the energies can be plotted against the number of AlAl linkages. The
example of such a calculation for the mineral leucite, KAlSi2 O6 , is shown in Figure 6.6.
The basic ideas can be extended to include interactions beyond nearest-neighbour. In this case a
simple plot of the energy against numbers of bonds will not suffice, and the exchange energies will
need to be obtained by fitting against the database of configuration energies. It is common to re-express
the ordering energies in terms of the spin model given above:

-30
-35
-40
-45

15
5
10
Number of AlOAl linkages

20

Figure 6.6: Plot of lattice energy as a function of the


number of AlOAl linkages in leucite, obtained using
many different atomic configurations.

140
T1
T2

Figure 6.7: Part of the crystal structure of cordierite,


showing the two distinct tetrahedral sites T1 and
T2 . The solid lines with arrowheads show some of
the distinct second-neighbour interactions (J2 ), and
the dashed line with arrow heads shows one thirdneighbour interaction (J3 ).

Initial configuration
of atoms

of the possible neighbouring interactions.


We have calculated exchange energies for Al/Si ordering in a number of aluminosilicates using a
tested empirical model for the interatomic energies. Typically the nearest-neighbour exchange constant
has values of between 0.41.3 eV per bond, and the exchange energies for more distant neighbours
are around five time smaller. These exchange energies appear to be reasonable in that they predict
the correct ordered structure, and in Monte Carlo simulations (see below) they give reasonable values
for the transition temperatures. Although we cannot develop the discussion here, the real value in
calculating the exchange energies for relatively complex aluminosilicate structures is that they enable
an understanding of the ordering process. For example, in cordierite it was clear from the exchange
energies that the ordering process involves first the ordering of Al and Si cations along the chains of
4-membered rings of tetrahedra (Figure 6.7) whilst preserving Al-avoidance within the 6-membered
rings without initially generating long-range order.
The models discussed here have assumed that the important interactions involve pairs of atoms.
However, the models can be extended to include the energies involved in forming larger groups of
atoms. The energy functions will then include terms of the form Kijk i j k , for example, representing
the energy involved in forming ordered triplets.

Monte Carlo method

Swap positions of
two atoms

Energy lowered
by swap

Energy raised
by swap

Keep swap

Keep swap subject


to probability test

New configuration
of atoms

Chapter 6. Orderdisorder phase transitions

Repeat process
many times

Figure 6.8: Flow diagram showing the operations in


the Monte Carlo method.

Given a parameterised equation for the energy function, as developed above, the Monte Carlo method
can be used to simulate the ordering process. The principle of this method is very simple, and the
basic procedure is illustrated as a flow diagram in Figure 6.8. Starting with an initial configuration of
atoms, two atoms are chosen at random and their positions are exchanged. This exchange, which can
be represented in in the spin representation as the swapping of the signs of two of the spins, may give
rise to a change in the energy, E, which can be calculated from the energy function. The rules of the
Monte Carlo method stipulate that if the energy of the new configuration is lower, i.e. if E < 0, the
exchange is kept. If there is no change in energy, the exchange is also kept. But if the energy of the new
configuration is higher, it is kept only subject to a probability test, namely with probability
p = exp (E/kB T )

(6.39)

where T is a pre-selected temperature. Whether or not the exchange of the atoms is kept or rejected, the
final configuration is counted as a new independent configuration, and the exchange step is repeated

6.3. Computational methods

141

using the new configuration as the initial configuration. This process is then repeated many times (of
order of millions of times or more).
The Monte Carlo gives many configurations of different energies, and after a suitable period of
equilibration any configuration of energy E will occur with its correct thermodynamic probability
exp( E/kB T ). For each configuration quantities such as energy and order parameter can be calculated
and stored, and from the many configurations average quantities can be formed. Moreover, fluctuations
can be calculated, and these give additional thermodynamic quantities. For example, fluctuations in
energy give the heat capacity (per atom),
C=



1
E
2
2
h
E
i

h
E
i
=
T
NkB T 2

(6.40)

and fluctuations in the order parameter give the susceptibility,


2 F
2

 1


1  2
h i h i2
NkB T

(6.41)

The Monte Carlo simulation can be performed at many temperatures in order to give the temperaturedependence of these quantities. For example, the temperature dependence of the order parameter for
Al/Si ordering in cordierite and sillimanite has been calculated using Monte Carlo simulations with
exchange energies computed from empirical potentials, and the dependence of the order parameter on
chemical composition and temperature has been calculated for the feldspar system using model energy
functions. The temperature dependence of the order parameter in the kalsilite structure, KAlSiO4 ,
computer by Monte Carlo is shown in Figure 6.5. The reciprocal of the ordering susceptibility is shown
in Figure 6.9. Both figures show the quality of the data that is possible to obtain reasonably easily with
the method.

1 / Susceptibility (arbitrary units)


=

100
80
60
40
20
0
0.5

Thermodynamic integration
The free energy F for a given model at a particular temperature T can be calculated in a Monte Carlo
simulation using the method of thermodynamic integration. The starting point is to separate the
Hamiltonian H for the model into two parts
H = H0 + H

(6.42)

0.6

0.7
0.8
Temperature / J

0.9

Figure 6.9: Temperature dependence of the reciprocal ordering susceptibility in the kalsilite structure
computed by the Monte Carlo method.

Chapter 6. Orderdisorder phase transitions

142

where H0 is the Hamiltonian of a model that is similar to the model but which can be solved exactly.
The free energy can be calculated from this separated Hamiltonian from
F = F0 +

Z1

hH i d

(6.43)

where F0 is the free energy associated with the Hamiltonian H0 , and hH i is the average of ( H H0 )
calculated from a distribution function for the system determined by the Hamiltonian
H = H + (1 ) H0 = H0 + H

(6.44)

In order to calculate hH i , the distribution function can be obtained using a Monte Carlo simulation
of the Hamiltonian H . That is, a simulation is performed using the Hamiltonian H , and the
configuration of atoms generated by the simulation is used to calculate the energy (H H0 ).
To illustrate this point further, for the simulation of orderdisorder behaviour, the Hamiltonian H
may represent a small perturbation of a system with random disorder, i.e. of a system with H0 = 0.
Therefore, if the Hamiltonian is given by
H=
we can then write
H =

1
J
j k
4 h
j,k i

(6.45)

1
J
j k
4 h
j,k i

(6.46)

The free energy is obtained by performing a set of simulations with the Hamiltonian H , and to
calculate the average hH i = H / from the atomic configurations generated by the simulation. The
integral is then performed by numerical integration of the results from the simulations performed over
a range of values of .
The free energy function F0 ( T ) corresponds to the free energy of a random distribution of two types
of cations of proportion x and (1 x ) and is given by the usual result:
F0 ( T ) = E0 TS0 = kB T ( x ln x + (1 x ) ln(1 x ))

(6.47)

This method has been used to calculate the free energy and entropy in simulations of the ordering of
Al and Si atoms over the tetrahedral sites in the feldspar structure, where the focus was on the effects of

6.4. Beyond BraggWilliams theory: the Cluster Variation Method

143

having non-equal numbers of Al and Si atoms. It has also been used to calculate the entropy associated
with disordering of Mg and Ca cations in garnet solid solutions of various compositions, showing the
effect of non-random short-range order on the entropy.

6.4 Beyond BraggWilliams theory: the Cluster Variation Method

Table 6.1: Bond types and corresponding energies


and probabilities in the CVM solid solution with no
long-range order.

Basic idea for solid solutions


The main problem with the BraggWilliams model is that no account is taken of short-range order.
Short-range order involves the formation of specific non-random arrangements of small groups or
clusters of atoms, so a better theory will attempt to account for the effect of these clusters on the
thermodynamic properties. The Cluster Variation Method (CVM) is the most successful approach of
this sort.
We can illustrate the CVM by considering the simple example of a solid solution containing a fraction
x of A atoms and (1 x ) B atoms, where the important clusters are nearest-neighbour bonds. The
energies and probabilities of the different bonds are given in Table 6.4, denoting the A and B atoms as
black and white respectively.
There are constraints on the probabilities:
y1 + 2y2 + y3 = 1

(6.48)

y1 + y2 = x

(6.49)

y2 + y3 = 1 x

(6.50)

Bond

Energy

Probability

E1
E2
E2
E3

y1
y2
y2
y3

which combine to give


y3 = 1 + y1 2x

(6.51)

The number of ways to distribute all zN/2 bonds is given as


(2)

z/2 =

(zN/2)!
(y1 zN/2)!(y2 zN/2)!2 (y3 zN/2)!

(6.52)

where the superscript (2) denotes that the probability refers to clusters of 2 atoms (i.e. bonds) and the
subscript z/2 denotes the number of such clusters (equal to zN/2).
This number is a substantial overestimate of the true number because no account has been taken to
ensure that the distribution of bonds is consistent with the distribution of atoms. This point is illustrated

Figure 6.10: Two distributions of bonds. The top distribution has atoms all of the same type at each atom site
and is therefore realistic. The bottom one has atoms
of different type at the atom sites and is therefore
unrealistic.

Chapter 6. Orderdisorder phase transitions

144

in Figure 6.10, which shows how it is possible to have configurations in which the bonds place atoms of
different type at the same atomic site. In order to filter out the unrealistic configurations we take the
configurations generated by placing an atom at the end of each bond, thereby giving z atoms at each of
the initial atomic sites, and then we count the number of ways of distributing these zN new atoms:
(1)

(zN )!
(zxN )! (z(1 x ) N )!

(6.53)

where the superscript (1) indicates that the clusters are of size 1 atom, and the subscript z indicates that
we have zN clusters (albeit clusters of 1 atom each). Of this number of configurations, the number that
will be correct will be the number of ways of forming arrangements with one atom per site:
(1)

1 =

N!
( xN )! ((1 x ) N )!

(6.54)

The critical approximation is that the number of ways of correctly distributing the bonds is determined by the statistics of distributing the atoms, and is given as
(2)

(1)

(1)

(2) = z/2 1 /z

(6.55)

That is, we have selected the same fraction out of all possible configurations of bonds as the fraction of
correct atom distributions out of all possible configurations of z atoms on each site.
We can now calculate the entropy of the system, using the methods discussed earlier:
(2)

(1)

(1)

S/kB = ln (2) = ln z/2 + ln 1 ln z

= zN/2 [y1 ln y1 + 2y2 ln y2 + y3 ln y3 ] + (z 1) N [ x ln x + (1 x ) ln(1 x )]

(6.56)

The energy of the system is given as


E=

zN
(y1 1 + 2y2 2 + y3 3 )
2

(6.57)

From the constraints on the probabilities we see that the final free energy will be a function only
of y1 , whose value can then be obtained by minimisation of the free energy if the bond energies are
available. And from the equilibrium value of y1 the entropy can be computed. Alternatively, it may be
possible to obtain y1 from experimental measurements. For example, for aluminosilicates it is possible
to determine the numbers of AlAl, SiAl and SiSi bonds using NMR, and from these results it may be

6.4. Beyond BraggWilliams theory: the Cluster Variation Method

145

possible to use the CVM method to calculate the entropy. Before getting too excited though, we need to
realise that this will be a calculation of the entropy obtained within certain approximations as outlined
above and discussed below. However, since some account is taken of the existence of short-range order,
the result will be better than assuming complete random disorder.

Orderdisorder phase transitions


The model can be extended to calculate the free energy associated with an orderdisorder phase
transition. For simplicity, we now consider the specific case of ordering of equal numbers of different
atoms over sites with four neighbours each. We now have N A atoms and N B atoms (2N atoms in
total). We can properly account for all bonds by considering the bonds coming from A sites, as shown
in Table 6.4.
There are constraints on the probabilities:
y1 + y2 + y3 + y4 = 1

(6.58)

y3 + y4 = 1

(6.59)

y1 = y4

(6.60)

y1 + y2 =

where is the probability that site of type A is occupied by an atom of type A (i.e. the probability of
having a right atom in the site). From these relationships we obtain the relevant constraints
y2 = y1

y3 = 1 y1

y4 = y1

(6.61)

We follow the procedure from above and write down the numbers of configurations (number of
way of arranging the 4N bonds, the number of arrangements with 4 atoms on each atomic site, and the
number of arrangements with one atom on each atomic site), noting that now we are now counting N
atoms of each type rather than N atoms in total, that there is a fraction of atoms of type A on sites of
type A and of atoms of type B on sites of type B, and that z = 4:
(2)

(4N )!
(4y1 N )!(4y2 N )!(4y3 N )!(4y4 N )!

2
(4N )!
=
(4 N )! (4(1 ) N )!

2
N!
=
( N )! ((1 ) N )!

2 =

(6.62)

(1)

(6.63)

(1)

(6.64)

Table 6.2: Bond types and corresponding energies


and probabilities in the CVM system with an order
disorder phase transition and equal numbers of both
types of atom.

Bond

Energy

Probability

J
J
J
J

y1
y2
y3
y4

Chapter 6. Orderdisorder phase transitions

146
The entropy follows as
(2)

(1)

(1)

S/kB = ln 2 + ln 1 ln 4
4

= 4N y j ln y j + 6N [ ln + (1 ) ln(1 )]

(6.65)

j =1

Using the constraints on the bond probabilities, the total bond energy is given as
E = N J (y1 y2 y3 + y4 ) = 4Ny1 J + E0

where E0 is the part of the energy that does not factor into bond probabilities. The free energy is now
a function of two quantities, and y1 . The equilibrium values of these quantities can be obtained by
minimisation of the free energy, which will then give the order parameter. The result for is shown
in Figure 6.5, where it is compared with the Monte Carlo and BraggWilliams results. It is clear that
the CVM method using bond clusters gives a significantly better result than the BraggWilliams result,
in particular in giving a transition temperature that is lower than that given by the BraggWilliams
model and which is much closer to the Monte Carlo result. In the BraggWilliams model the bond
probabilities are constrained by the value of , rather than being determined as a separate minimisation
of the free energy. Thus for all temperatures above the transition temperature the relative numbers of
the different types of bond are always those of a purely random distribution of atoms, whereas one
would expect there to be more AlSi bonds than this because of the energy gain. In Figure 6.11 we
compare the fraction of AlSi bonds in the kalsilite model calculated by the CVM method with that
given by the Monte Carlo simulations. The CVM method has given a reasonable estimate of the number
of these bonds. The BraggWilliams model gives exactly 1/2 for all temperatures above the transition
temperature, and it is clear that short-range order always gives more AlSi bonds than this.

Fraction of AlSi bonds

1.1
1.0
0.9
0.8
0.7
0.6
0.5
0.0

(6.66)

0.2

0.4

0.6 0.8 1.0


Temperature / J

1.2

1.4

1.6

Figure 6.11: Calculations of the fraction of AlSi bonds


in the model of Al/Si ordering in kalsilite. The points
show the Monte Carlo results, and the curve shows
the results from the bond-CVM method.

Improved CVM calculations


The CVM calculations described here have used two-atom bonds as the basic clusters. This is attractive
since we usually express the energies in terms of bond energies, and experiments often give data
about bond probabilities. However, these are very small clusters, and for improved accuracy in the
calculations it is necessary to use rather larger clusters. An improvement in accuracy may not appear to
be particularly necessary from the results given here for kalsilite, Figures 6.5 and 6.11, but our example

6.5. Summary

147

has been a little misleading in that it is one for which the CVM method is likely to work reasonably
well. Where the bond CVM method will work less well is in the case where there is not equal numbers
of ordering atoms.
It is perhaps not entirely clear from the analysis above that as the size of the clusters increases,
the number of variables increases exponentially, and the problem then becomes extremely difficult.
Moreover, it is not at all a trivial matter to be sure that a calculation has used the relevant clusters. With
these qualifications, we are now in sight of modern research efforts.

6.5 Summary
1. Ordering of atoms across sites that are equivalent in the high-temperature phase can be found in
metal alloys, and in ceramics or minerals.
2. The mean-field approach is to represent each ordering atom as seeing an average environment as
far as the calculation of energy is concerned, and to approximate the entropy as that of random
arrangements within the constraint of a given degree of long-range order. That is, the mean-field
approximation ignores all short-range correlations other than those generated by the long-range
order, and assumes a maximum degree of entropy.
3. Because of its neglect of short-range order, the mean-field approximation typically over-estimates
the value of Tc by 50% in the best case; in the worst case the discrepancy between the mean-field
value of Tc and that for the real system can be huge.
4. The problem of atomic site ordering involves short-range interactions, and can be approximated
by simple spin-like models. These are amenable to study using the Monte Carlo method.
5. The cluster variation method (CVM) is a way to incorporate short-range order. It works on the
basis of calculating the ordering of bonds as independent objects, and corrects for cases where
the ends of bonds are inconsistent with the atoms by scaling the number of configurations based
on the numbers of arrangement of bond ends. The CVM method can be expanded to work with
clusters larger than bonds.

148

6.6

Chapter 6. Orderdisorder phase transitions

Short questions to test your basic knowledge


1. Define the order parameter for an atomic site-ordering phase transition using the example of the
ordering of 2 atom types on a bcc lattice.
2. Describe the central approximations of the BraggWilliams model of atomic site ordering.
3. In what ways is the BraggWilliams model a mean-field theory?
4. Give the values of some of the critical exponents given by the BraggWilliams model.
5. What are the main failures of the BraggWilliams model?
6. Why does the BraggWilliams model overestimate the value of Tc ?
7. Describe the main aspects of the Monte Carlo method.
8. In what way can the problem of atomic site ordering be represented as a spin model? In which
one way is there an additional constraint not found in a spin model of magnetism?
9. In what two ways does the CVM do a much better job of describing atomic site ordering phase
transitions than the BraggWilliams model?
10. Describe the central approximation of the CVM method.

6.7

Longer questions
Consider the ordering of equal numbers of two types of cations in a network structure. Each site has
four neighbours, and in the ordered structure all neighbours are of different types. Use a value for the
exchange energy, J, of 0.5 eV or 50 kJ mol1 , a typical value for Al/Si exchange.
The task is to use Excel to calculate the free energy of both the BraggWilliams and pair CVM
models, both being a function of the order parameter, and the pair CVM model also having a parameter
that describes the short-range order. (If you prefer, you are of course free to program your own code to
solve this problem using Matlab, Python, Fortran, C++, or whatever your favourite computer language
is.)

6.7. Longer questions

149

At any given temperature, you can determine the long-range order parameter (and short-range order
in the case of the CVM model) by using the solver/minimise function within Excel (please ask if you
do not know how to do this) or some other tool that will minimise a function with respect to one or
two variables. If you are using Excel, the target cell to minimise will be that containing the free energy
function, and the cells to minimise will be those containing information about the long-range order
(which may be written as a site occupancy) and the short-range order (which may be written as the
probability of forming a bond of a certain type).
1. Use Excel to determine the temperature-dependence of the order parameter in the BraggWilliams
model. Note that you have an exact equation for Tc from the lectures, which you can use to set the
temperature scale to be investigated.
2. Use Excel to determine the temperature-dependence of the long-range and short-range order
parameters in the pair CVM model.
In both cases, you should aim to use a range of temperatures from near zero to just above the
BraggWilliams Tc . Note that at very low temperatures the minimisation process may be unstable, so
do not worry about being very close to a temperature of zero.
Beware of producing too compact or complicated spreadsheets. The best spreadsheets are those
where intermediate steps are handled in separate cells, and then brought together in subsequent cells.
Treat each temperature on a separate row so that you keep a record of each temperature.
Produce and hand in the following graphs, with accompanying notes and comments, for both
models:
1. Plots of the equilibrium free energies of both models as functions of temperature.
2. Plots of the long-range order parameter for both models as functions of temperature.
3. Plot of the number of bonds of bonds of one type of atom for both models as a function of
temperature. In the case of the CVM model, this function is the short-range order fed directly into
the free energy. For the BraggWilliams model, this function is related to the potential energy.

Chapter 7

Critical phenomena

7.1 Recapitulation
We have now discussed several versions of mean-field theory that aim to describe various sorts of phase
transition. As we have seen, the major problem with these theories is that near the critical temperature,
as locally ordered regions begin to grow, the local environment seen by a particular site might be quite
different from the global mean; yet mean-field theories have no means to account for this. We might
simplistically view phase transitions as arising from a competition between energy and entropy, with
energetic considerations dominating at low temperatures and causing an ordered phase while entropy
dominates at high temperatures, favouring disorder. From this perspective, locally ordered regions
achieve a favourable energetic state at less entropic cost than full global order. For this reason, as we
have seen, phase transitions tend to occur in real life at lower temperatures than mean-field theory
predicts.
In order to move forward, then, we need to consider theories that take into account fluctuations in
the order parameter. In this chapter we will first consider how to quantify these fluctuations using
correlation functions. We then return to the idea of critical exponents, which we have investigated over
the course of this module; with the introduction of correlation functions we will have defined all six
of the most commonly considered such exponents. It is intriguing that these exponents appear to be
the same across large and apparently unrelated categories of phase transition; this is the concept of
universality classes. Finally we will introduce the theory of renormalisation groups, which is able to unify
these strands of the module and explain why critical exponents describe phase transitions so well
independently of the specific physical details of the transition.
151

152

7.2

Chapter 7. Critical phenomena

Correlation functions
Definition and behaviour
For the sake of concreteness, consider a familiar example: the two-dimensional Ising model. We will
frequently return to this model as an example throughout this chapter, although none of the arguments
we make will be particular to this model. An important difference between the two phases is that
there is long-range order in one but not in the other: that is, the values of neighbouring spins in the
low-temperature phase are highly correlated, but the correlation decreases with increasing temperature.
Since the correlations between neighbouring spins have proven so important, it is natural also to
consider the extent to which the fluctuations of nearby sites are correlated to one another. We can look
specifically at the fluctuations by subtracting the time average of each sites spin from the spins observed
in a particular snapshot. Thus we can directly compare low and high temperatures without worrying
about the fact that most spins have the same direction at low T (so the fluctuations are more or less just
the spins that point the other way) while this is no longer true at high T.
This motivates the definition of the spin-spin correlation function between the sites with position
vectors ri and r j as the thermal average of the product of each sites deviation from its individual
average:


g(ri , r j ) = (si hsi i)(s j hs j i)
(7.1)
just as we saw in section 3.17. We can simplify this by invoking the translational symmetry of our
model: since each site is identical, hsi i = hs j i hsi, and moreover g must be a function only of ri r j .
Thus we have
g (r) = h s i s j i h s i2 .
(7.2)
The next question is what this function actually looks like in a real system. We might expect that
g(r ) should decay exponentially with r, and indeed we find that it takes the form


1
r
g(r ) d2+ exp
(7.3)

1 Note

that this is not the order parameter.

Here d is the dimensionality of the model, is another critical exponent,1 and is a characteristic
length known as the correlation length. This will be small both at high temperatures, where entropy
dominates and there is little correlation between the spin of any site and its neighbours, and also at low
temperatures, because as we argued above the fluctuations, spins that point the wrong way, will

7.2. Correlation functions

153

also be randomly dispersed through the system, being rare enough that they are not significantly more
likely to occur on adjacent sites. On the other hand, at the critical temperature Tc , where the average
spin is still zero but the fluctuations are on the point of encompassing the entire system, the correlation
length is effectively infinite. In fact, we find that
| T Tc |

(7.4)

where is yet another critical exponent.

Correlation length and scaling


As a further method for visualising the divergence of the correlation length at the transition temperature,
consider the following operation. We are going to reduce our simulation volume by a factor of 9, as
follows: we divide our simulation into 3 3 cells, and replace each of these by a single site with the
majority spin of the parent nine. For instance, if five of the sites in a cell have = 1 and the other
four have = 1, then the majority rules and our replacement site will also have = 1, as shown in
figure 7.1. We call this operation renormalisation. The important part about it is not the majority rule
other rules for assigning the spin of the replacement site are possible but the fact that we are reducing
the size of our system in a way that more or less preserves the interactions between neighbouring sites.
If our system has N sites2 before renormalisation and N 0 afterwards, we can define a scale factor s of the
renormalisation by
N
= sd
(7.5)
N0
where d is the dimensionality. So in our two-dimensional example, N/N 0 = 9 and the scale factor is
s = 3, as one might expect from the description above.
Later we will come back and define renormalisation more precisely; for the time being lets consider
the effect of performing this strange transformation. To be specific, lets consider what effect this
renormalisation has on the correlation length. In the left-hand column of figure 7.2 are shown three
81 81 sections of a Monte Carlo simulation of the 2D Ising model, with temperatures just above Tc , at
Tc , and just below Tc . In each row, moving to the right in the figure corresponds to renormalisation
according to our majority rule (and if you look carefully, you can see each figure in the bottom left corner
of the one to its right). We can make two key observations about this process. First, the correlation
length decreases on renormalisation both above and below Tc , but appears not to change at Tc itself:

Figure 7.1: One way of renormalising the 2D Ising


model. 3 3 blocks are replaced by a single site with
the most prevalent spin among the 9 original variables.

2 Or,

more formally, degrees of freedom.

Chapter 7. Critical phenomena

154

T = 1.1Tc

T = Tc

T = 0.99Tc

Figure 7.2: Renormalisation of the 2D Ising model. In each row, each image is related to the one on its left by
renormalisation as shown in Figure 7.1.

7.3. The Widom scaling hypothesis: relationships between critical exponents

155

at this temperature, after renormalisation we can still see clusters of around the same length as those
in the original configuration. It is not entirely surprising that the correlation length should decrease
on renormalisation: after all, our renormalisation procedure reduces every length in the system by a
factor of s, including the correlation length. The fact that does not appear to decrease at the critical
temperature therefore suggests that it is in fact infinite at this temperature, and indeed detailed analysis
bears this out, as we shall see.
Our second observation is that the renormalised configurations are perfectly plausible states of the
same 2D Ising model as we started with, but the temperature appears to be changing. Renormalising the
configuration above Tc looks rather as if we have raised the temperature, and conversely renormalising
the configuration below Tc looks like we have lowered it. Again, at Tc itself no change is obvious. We
discuss the significance of these observations in turn in the following two sections.

Table 7.1: Definitions of critical exponents. For convenience we define = ( T Tc )/Tc . Note that the order
parameter and the critical exponent are unfortunately not the same thing!

Exponent

Definition

Reference

C (, = 0) | |
(, = 0) | |
(, = 0) | |
( = 0, ) ||1/
| |
g ( r ) r d +2

3.11
1.2
3.17
2.29
7.4
7.3

7.3 The Widom scaling hypothesis: relationships between critical exponents


Scaling relations
Our first observation was that the critical temperature is associated with some sort of scale invariance:
renormalising, or zooming out, a configuration at this temperature, doesnt change its characteristic
length scale . The simplest mathematical functions that are scale-invariant are power laws, which
partially explains why the critical exponents of a phase transition in other words, the power laws
that predict the limiting behaviour of physical properties near the transition are so important in
characterising that transition.
For this reason, lets reconsider the six critical exponents, whose definitions are summarised in
Table 7.1. We have elided over one detail in this table, which is that as previously noted, the critical
exponents that are defined both above and below the transition might in theory have different values
in these regions. For instance, the susceptibility might diverge with a critical exponent as we cool
towards the transition from the high-temperature phase, but with a different exponent 0 as we heat the
low-temperature phase.3 In general, however, we find that these pairs of exponents have the same value,
= 0 , so we will not make any further distinction between these pairs.
At this point is it useful to look at some specific results for values of the critical exponents. These can
be measured experimentally, obtained by appropriate computer simulation, or in some cases calculated
analytically through various schemes. The results for various models are given in Table 7.2.4

3 The

other exponents that this affects are and ; if it is


important to make a distinction in these cases, an unprimed
symbol refers to cooling while a primed symbol refers to
heating. Make sure you understand why this potential complication does not affect the other three critical exponents ,
, and .

4 It

is worth remarking that values of critical exponents from


experimental or simulation data are typically obtained from
logarithmic plots. If some quantity x | | , we have ln x =
ln | | + constant, and the value of the critical exponent is
obtained as the slope of the plot in the limit 0.

Chapter 7. Critical phenomena

156

Table 7.2: Values of the critical exponents, as defined in Table 7.1 for various classes of phase transition. The xy
model has the spins confined to a plane but with any orientation. The Potts model is a generalisation of the Ising
model, with the 3-state Potts model having 3 allowed orientations of the spin. The 3d 3-state Potts model gives the
same results as the 3d xy model.

Phase transition

2nd order mean field

Tricritical

1
2
1
2

1
2

1
2
1
4
1
8
1
9
1
16
5
16

7
4
13
9
7
8
54

15

14

5
6
1
4

1
4
4
15
1
4

2d Ising

2d 3-state Potts

1
3

2d xy

1
1
8

15

5
8

3d Ising

3d xy

0.34

1.31

4.92

0.66

0.11

0.36

1.39

4.86

0.70

0.03

3d Heisenberg

As an exercise, we can look for relationships between these critical exponent values within one
system. Although the values of the critical exponents differ considerable from one system to another,
check the results and show that
+ 2 + = 2

(Rushbrooke)

(7.6)

( 1) =

(Widom)

(7.7)

(2 ) =

(Fisher)

(7.8)

(Josephson, hyperscaling)

(7.9)

2 = d

where the names in brackets are the people who first identified these relationships. These relationships
between the values of critical exponents are called scaling relations.

Scaling and the free energy


Lets consider the effect of rescaling on the free energy F. For simplicity we write F as a function of
= ( T Tc )/Tc and the applied field . Furthermore, we assume that there is a power-law scaling

7.3. The Widom scaling hypothesis: relationships between critical exponents

157

relationship between these parameters and the free energy:


F ( p , q ) = F (, )

(7.10)

where is a rescaling factor and p and q are characteristic parameters of the particular Hamiltonian
of our model. This is known as the Widom scaling hypothesis. Historically this was proposed as a way
of justifying the empirically-observed scaling relationships between the critical exponents; it was later
proven using the theory of the renormalisation group. Here we will use it to derive these relationships
between the critical exponents. By definition the equilibrium order parameter is given by
=

(7.11)

and thus differentiating equation 7.10 with respect to gives


q ( p , q ) = (, ).

(7.12)

Now we can start considering particular values of the parameters. For instance, consider the lowtemperature phase ( < 0) in zero field ( = 0), and set = ( )1/p ; since is just the factor by
which we are scaling, we can choose any convenient value for it. Substituting into equation 7.12 gives

( )q/p ( = 1, = 0) = ( )1/p (, = 0)

(7.13)

(, = 0) = ( )(1q)/p ( = 1, = 0).

(7.14)

and hence

Comparing this to the definition of immediately leads to the conclusion that under the Widom
hypothesis = (1 q)/p.
Similarly, by definition the susceptibility is
=

(7.15)

so differentiating equation 7.12 with respect to gives


2q ( p , q ) = (, ).

(7.16)

Chapter 7. Critical phenomena

158
This time we set = 0 and = | |1/p to give

| |2q/p ( = 1, = 0) = | |1/p (, = 0)
(, = 0) = | |(12q)/p ( = 1, = 0)

(7.17)
(7.18)

and hence = (2q 1)/p under Widom scaling.5

5 In

We can proceed along the same lines to calculate the other critical exponents. For we will need to
consider the heat capacity at constant field, which by a standard thermodynamic argument is given by
C = T

2 F
2 F
=

T
.
T 2
2

(7.19)

Thus differentiating equation 7.10 twice with respect to gives


2p C ( p , q ) = C (, )

(7.20)

and again setting = 0 and = | |1/p gives

| |2 C ( = 1, = 0) = | |1/p C (, = 0)
C (, = 0) = | |

2+1/p

C ( = 1, = 0)

(7.21)
(7.22)

and hence = 2 1/p.


Finally, for we need to consider the order parameter again, but this time as a function of field at
the critical temperature. So we return to equation 7.12, and this time set = 0 and = 1/q . Then
1 ( = 0, = 1) = 1/q ( = 0, )
( = 0, ) =

1+1/q

( = 0, = 1)

(7.23)
(7.24)

and 1/ = 1 1/q or = q/(1 q).


All of these relationships are interesting, but expressing things in terms of p and q is in the present
analysis less relevant than the critical exponents themselves (although we will later see that there is
a significance to these apparently arbitrary parameters). We can combine the relationships we have
developed to give equalities that dont depend on p or q. These are named after the researchers who

fact, we ha

7.3. The Widom scaling hypothesis: relationships between critical exponents

159

first proposed them:


2p 1
1 q 2q 1
+2
+
=2
(Rushbrooke)
p
p
p
1q
1
1
( + 1) =

= = 2
(Griffiths)
p
1q
p
2q 1
1 q 2q 1

=
=
(Widom)
( 1) =
p
1q
p

+ 2 + =

(7.25)
(7.26)
(7.27)

We can deduce a relationship involving in a similar way. We begin by considering a renormalisation


process with scale factor s. If we consider the free energy per unit volume, which we label f , then by
analogy with equation 7.5 we have
f0
= sd
(7.28)
f
note that the ratio is the opposite way around to equation 7.5, since keeping the same free energy F
but reducing the volume will increase the free energy per unit volume f . Now by our scaling hypothesis
we also know that
f ( p ) = f ( ).
(7.29)
So suppose that our renormalisation process results in changing correlation length from to 0 = /s,
and that this corresponds to a change in reduced temperature6 from to 0 = p . Then combining
equations 7.28 and 7.29 gives
f0
f ( p )
sd =
=
= .
(7.30)
f
f ( )
Furthermore, from the definition of the critical exponent , | | , we can write
0 =

| |

= (s1/ | |)
s
s

(7.31)

but since also


0 | 0 | = ( p | |) ,

(7.32)

s1/ = p .

(7.33)

it follows that
Putting this all together, we combine equations 7.30 and 7.33 to give
sdp = p = s1/

(7.34)

6 Recall

that the fact that we can assign a reduced temperature


to the renormalised configuration at all was our second observation from Figure 7.2. We will discuss this further in the
following section.

Chapter 7. Critical phenomena

160
and hence
d = 1/p = 2

(Josephson).

(7.35)

7.4 Introduction to the renormalisation group


Universality classes
It is an interesting observation that many apparently unrelated phase transitions for instance, the
liquid-gas transition of xenon, the transition between miscibility and separation in binary mixtures of
organic liquids, the transition between He-I and He-II, and the order-disorder transition in -brass
show similar values of the critical exponents (which are almost all equal within experimental error).
This striking result suggests that there is an underlying mechanism for phase transitions that does not
depend on the specific interactions involved, but only on very general properties such as the nature of
the order parameter and the dimensionality of the system involved. The theory of the renormalisation
group provides this sort of general result. We will only have time to introduce the basic ideas involved,
but we will demonstrate how they can be used to calculate critical exponents very generally.

Real-space renormalisation
As promised, lets return to consider in more detail what we mean by renormalisation. In particular, in
this module we will consider real-space renormalisation, a theory originally due to Kadanoff.7 As we have
seen, renormalisation involves reducing the number of degrees of freedom of our model by blocking
groups of them together. There are a number of different ways to do this: we have already seen the
majority rules system, and could instead average the spins on the sites that are blocked together
(rounding to an allowable value for our model), or simply choose one site at random and assign the
blocked site the same value, a rule called decimation.

7 The alternative is k-space renormalisation, in which we


perform mathematically similar scaling operations, but this
time the aim is to eliminate parameters in models of the
Fourier transform of the order parameter. The name comes
from the fact that this transform lives in reciprocal space,
or k-space.

We now make the crucial assumption that the Hamiltonian after renormalisation takes the same
functional form as beforehand. For instance, if our original model, as usual, is a 2d Ising model, we
assume that we can also describe the renormalised system as a 2d Ising model, i.e., that we still only
need to consider nearest-neighbour interactions. It is not entirely obvious that this will be valid: for
instance, even if our original model contained only second-nearest neighbour interactions, we might
expect the renormalised model to depend on nearest-neighbour interactions too. In this case we would

7.4. Introduction to the renormalisation group

161

need to reframe our original model to include this extra parameter (which would initially be zero) in
order to proceed.
If the Hamiltonian has the same functional form as previously, the only thing that changes on
renormalisation is its parameters. For example, the state of the Ising model (in any number of
dimensions) can be described by two parameters, the external field B and J/kB T.8 By assumption,
renormalisation will take us to a new state completely characterised by new values of these parameters.
This can be described as motion (by discrete jumps) in a two-dimensional parameter space, which
is sometimes referred to as renormalisation flow. In mathematical notation, we represent our initial
Hamiltonian by a vector h in this space, and write that under renormalisation it transforms to a new
vector R(h).
In general our transformation R will have fixed points for which R(h0 ) = h0 . In the mathematical
theory of nonlinear dynamics, fixed points can be classified according to the behaviour of nearby points
as the transformation is iterated. If, on repeated renormalisation, nearby points get closer and closer
to the fixed point, it is called an attractive fixed point; if on the other hand they get further and further
away, it is called a repulsive fixed point. The third and most interesting possibility is a mixed fixed point,
which will attract points that are close to it in some directions, but repel those that are close in others.
To perform this classification on a fixed point h0 , we expand the effect of R on a nearby point as a Taylor
series:
h0 + h0 = R(h0 + h)

8 Note

that the coupling constant J and the temperature T


are not independent: only their ratio matters. Thus, while
we previously described Figure 7.2 in terms of changing
temperature, we could equally well have said that on renormalisation the coupling constant J increases below the critical
temperature, and decreases above it.

(7.36)
2

= R(h0 ) + Mh + O(h )

(7.37)

high T

where

R1 (h)
h1
R2 (h)
h1

R1 (h)
h2
R2 (h)
h2

Rn (h)
h1

R N (h)
h2

M=

..
.

..
.

..
.

R1 (h)
h N
R2 (h)
h N

..
.

R N (h)
h N

critical

(7.38)

here the subscripts represent components of the various vectors, the parameter space is N-dimensional,
and the derivatives are all taken at h0 . Since h0 is a fixed point, we conclude that for small h
low T

h0 = Mh.

critical
surface

(7.39)
Figure 7.3: Sketch of Hamiltonian phase space, showing the lines of flow, the low and high-temperature
attractive fixed points, the critical mixed fixed point,
and the critical surface.

162

9 We

will not consider awkward possibilities such as eigenvalues exactly equal to 1 here, but it should be clear that the
behaviour in this case can often be resolved by considering
higher-order terms in equation 7.36.

high T

critical

low T

critical
surface

Figure 7.4: A trajectory in parameter space that begins very close to the critical surface, and therefore
approaches the critical fixed point but eventually turns
away from it.

Chapter 7. Critical phenomena

From this simple relationship it is easy to see what will happen: in the directions of eigenvectors of
M corresponding to eigenvalues less than 1, the fixed point will be attractive, and in the directions of
eigenvectors with eigenvalues greater than 1, it will be repulsive.9
Returning to phase transitions, lets consider what fixed points we might expect in the renormalisation flow. If all of our spins take the same value, as at T = 0, then any sensible renormalisation
procedure will clearly retain the same configuration, and we must be at a low-temperature fixed point. On
the other hand, if the spins are entirely uncorrelated, as at T = (or J = 0), then again any sensible
renormalisation will retain this property and we have a high-temperature fixed point. Returning once
again to figure 7.2 suggests strongly that both of these fixed points are attractive, which can indeed be
proven by considering the effect in general of decreasing the correlation length by the scaling factor s.
In fact we discovered only one point in Hamiltonian space that is attracted to neither the high- nor
the low-temperature fixed points: this of course corresponds to the critical temperature, and in the
Ising model it is itself a fixed point. More generally we can imagine that the points attracted to the
high-temperature fixed point and those attracted to the low-temperature fixed point will be separated in
Hamiltonian space by a critical surface. Iteration starting from a point on this surface will lead towards a
critical fixed point, which is an example of a mixed fixed point: it will attract nearby points on the critical
surface, but repel nearby points in other directions. This situation is sketched in figure 7.3.

An example: calculating
We now have a formalism within which we can calculate the critical exponents. Since this is a little
mathematically involved, in this module we will only demonstrate this calculation for one, . Recall
that this describes the behaviour of the correlation length as the critical temperature is approached.
Our starting point will be to consider a trajectory that begins close to, but not actually on, the critical
surface. The flow will initially be roughly parallel to the critical surface, but at some point it will turn
out towards an attractive fixed point (Figure 7.4). We will attempt to estimate the rough location of the
point at which the trajectory turns.
First, we note that the correlation length will initially be very large, but decrease gradually so that
after n iterations it is

(7.40)
(n) = n .
s
The high-temperature fixed point, on the other hand, is characterised by the fact that there is no
correlation at all between adjacent sites. So a rough characterisation of the turning point is the point

7.4. Introduction to the renormalisation group

163

at which the correlation length is in some sense small enough. We will denote this threshold by ` (it
will turn out not to matter exactly what value we take):

= `.
sn

(7.41)

We can also consider this situation, however, in terms of the analysis that gave us equation 7.39.
Denote the critical fixed point by h0 , and evaluate M in at this point according to equation 7.38. We
know that h0 is a mixed fixed point, and therefore that M has at least one attractive eigenvalue a < 1
and at least one repulsive eigenvalue r > 1. For simplicity, well assume that the parameter space is
two-dimensional, and therefore that M is 2 2 and these are the only two eigenvalues. We denote the
corresponding eigenvectors by ha and hr . Then we can express any point h in terms of the eigenvectors:
h = h0 + xa ha + xr hr

(7.42)

and, provided that both h and its transformed value after renormalisation are close enough to the
critical fixed point, after m iterations of renormalisation this will become
h(m) = h0 + xa am ha + xr rm hr .

(7.43)

Now we havent assumed that the starting point of the trajectory is near the critical fixed point, just
near the critical surface: we need somehow to include this assumption in our working. We can easily
modify the previous analysis to allow us to start from a general position by allowing a number r of
iterations before we get close enough to h0 that equation 7.42 holds. If we want to evaluate the situation
after a total of n iterations, then, equation 7.43 becomes
h(n) = h0 + xa anr ha + xr rnr hr

(7.44)

(still assuming, of course, that h(n) is close to h0 ). To incorporate the fact that we start near the critical
surface, we consider the coefficient xr : obviously if our starting point is on the critical surface, it will be
zero. We can therefore expand this to first order in a Taylor series in | T Tc |:
(1)

xr = xr | T Tc | + O(| T Tc |2 ).

(7.45)

Putting all of this together shows that after n iterations we will have reached
(1)

h(n) = h0 + xa anr ha + xr | T Tc |rnr hr .

(7.46)

Chapter 7. Critical phenomena

164

The turning point occurs when the coefficient of the repulsive eigenvector is in some sense large
enough. Calling this threshold L and again not worrying too much about how large is large enough,
this gives the criterion
(1)
xr | T Tc |rnr = L.
(7.47)
We now compare our two characterisations of the turning point, equations 7.41 and 7.47. We are
not particularly interested in the number of iterations n it takes to get there, so we eliminate n between
these two equations:
 

(7.48)
n log s = log
`
!
Lrr
n log r = log
(7.49)
(1)
xr | T Tc |
 

log
log s
`
(7.50)
=
(1)
log r
r
log( L ) log( x | T Tc |)
r

We now define
=

log s
log r

(7.51)

(noting that this has formally nothing to do with our critical exponent at this stage, although they are
of course going to turn out to be the same thing). Then equation 7.50 rearranges neatly to
!

Lrr
(7.52)
=
(1)
`
x | T Tc |
r

10 More generally, in the case of a higher-dimensional


parameter space, it is not difficult to see that it is the greatest
eigenvalue that will be important here.

11 This

is incidentally why we have chosen to demonstrate


the calculation of ; the other critical exponents require the
details of the blocking procedure to be known as part of the
calculation.

and hence | T Tc | , showing that equation 7.51 is consistent with our previous definition of .
The point of all this mathematical manipulation is that none of the guesstimated quantities like `
and L turn out to matter: in equation 7.51 we have expressed the critical exponent in terms only of
characteristics of the renormalisation procedure we have chosen. In order to determine , once we have
chosen a renormalisation function with scale factor s, we need only to find its critical fixed point, expand
the transformation to first order about that point, giving a transformation matrix M, and calculate its
repulsive eigenvalue r .10 This may be somewhat mathematically involved, but conceptually it is not
difficult; moreover, within the constraints of our assumptions about the renormalisation procedure, we
are free to choose a transformation in such a way that this calculation is as easy as possible.11

7.5. Deriving the Widom scaling hypothesis

165

In the following section, we continue in a similar fashion to calculate values of all the critical
exponents. In the process we will justify the Widom scaling hypothesis, and hence show that the
relationships between the critical exponents that we previously derived are valid. For now, though, we
note in particular that all of the critical exponents depend on only two parameters (in the analysis of
section 7.3 these were p and q), plus the dimensionality d of the system. We conventionally choose these
parameters to be the critical exponents and themselves, and write , , , and in terms of them;
these relationships are given in Table 7.3.

Table 7.3: Critical exponents expressed in terms of


and (and the dimensionality d).

Exponent

7.5 Deriving the Widom scaling hypothesis


Having considered the theory of real space renormalisation, we finally backtrack to see how this can be
used to justify the Widom scaling hypothesis, equation 7.10. We begin with equation 7.28 but this time
we will explicitly write f as a function of the variables and . From the discussion of Figure 7.2, we
expect that corresponds to a repulsive eigenvalue of the Jacobian matrix M at the critical point, which
well call . Adjusting the field away from its critical value = 0 would result in a similar diagram12
and hence this corresponds to a different eigenvalue . So we have
f 0 = f ( , ) = sd f (, ).

12 You

Value
2 d
2)
(2 )
d+2
d2+

1
2 (d +

might like to think about exactly what this would look

like.

(7.53)

Actually, we can be more precise about those eigenvalues. By the definition of renormalisation, the
effect of first renormalising with scale factor s1 and then with scale factor s2 has to be a renormalisation
with scale factor s1 s2 . The only way this can be true in general is if the eigenvalues take the form
= sy

= sy

(7.54)

for some y and y .13 Combining equations 7.53 and 7.54 gives
f (sy , sy ) = sd f (, ).

13 Of

(7.55)

This looks very much like the Widom hypothesis, with the bonus that it now has a physical meaning:
it represents the fact that the free energy before and after a renormalisation step must be the same.
Indeed, if we set
= sd
p = y /d
q = y /d
(7.56)
we return to the Widom equation 7.10. Thus we arrive at another set of equations for the critical
exponents, summarised in Table 7.4.

course, there is nothing special about these two particular


eigenvalues; every eigenvalue of this matrix must have this
form.

Table 7.4: Critical exponents expressed in terms of y


and y (and the dimensionality d).

Exponent

Value

2 d/y
(d y )/y
(2y d)/y
y /(d y )
1/y
d 2y + 2

Chapter 7. Critical phenomena

166

7.6

A sketched example: the 1D Ising model


We conclude this section with a brief example showing that even the simplest of the models we have
considered, the one-dimensional Ising model, is far from trivial to analyse. To be specific, we will
normalise by decimation with scale factor s = 2; that is, the normalisation process will simply involve
erasing every second site in our model. Our Hamiltonion, as usual, is
nn

E = J Si S j B Si
i,j

(7.57)

where, again as usual, Im using the abbreviation nn to indicate that the sum is over nearest
neighbours.
Lets consider the first three sites, S1 , S2 , and S3 . Once S2 has been removed, the probability of the
resulting state has to be the same as it was before renormalisation. In other words,

P ( S1 , S3 ) =

P ( S1 , S2 , S3 )

exp JS2 (S1 + S3 ) + BS2

(7.58)

S2 =1

(7.59)

S2 =1

(note that we are for the moment ignoring the interaction of spins 1 and 3 with the external field)

= 2 cosh J (S1 + S3 ) + B)

= A exp J 0 S1 S3 + B (S1 + S3 )

(7.60)
(7.61)

where the last equality comes from the fact that, by the renormalisation hypothesis, the Hamiltonian
must have the same functional form before and after renormalisation, with potentially different values
of the parameters.
There are a couple of bookkeeping niggles to point out before we continue. The first is that, rather
than bother to deal with the partition function (which, as we know from statistical mechanics, would be
the denominator of the probability of any particular state), we simply include an arbitrary constant of
proportionality A in equation 7.61. We could calculate this explicitly but since it will cancel out in the
subsequent analysis theres not much point. The second point is that the new external field is in fact
given by
B0 = 2 B + B.
(7.62)

7.6. A sketched example: the 1D Ising model

167

This follows from the fact that each of the remaining spins will have a contribution like B from each
side (e.g., for site 3 the analysis above applies to its interactions with both sites 1 and 5), as well as
a contribution from its own interaction with the external field, which as noted above weve hitherto
ignored.
Returning to the main thread of the argument, we can substitute the four possible values S1 , S3 = 1
into equation 7.61 to give the simultaneous equations

2 cosh(2J + B) = A exp( J 0 + B )

(7.63)
2 cosh( B) = A exp( J 0 )

2 cosh(2J + B) = A exp( J 0 B ).
Solving for J 0 and B (as pointed out above we can ignore A), and using 7.62, we find the solutions
exp(4J 0 ) =

cosh(2J + B) cosh(2J + B)

cosh2 B
cosh(2J + B)
exp(2B0 ) = exp(2B)
cosh(2J + B)

(7.64)
(7.65)

or, letting x = exp(4J ) and y = exp(2B),


x (1 + y )2
( x + y)(1 + xy)
y( x + y)
.
y0 =
1 + xy

x0 =

(7.66)
(7.67)

We can now find the critical fixed point at xc = 0, yc = 1, corresponding to zero temperature and
field. (The high-temperature fixed point is at xht = 1, yht arbitrary, corresponding to J = 0; as we have
noted, in this model zero coupling is equivalent to infinite temperature. There is a final, attractive fixed
point at xlt = 0, ylt = 0; this corresponds to infinite external field. Note that in this model, unlike in
higher dimensions the phase transition occurs, or would occur, at zero temperature.) The Jacobian at
this point is simply
!
4 0
M=
(7.68)
0 2
so that we have (with s = 2) y = 2, y = 1. It immediately follows that = 32 , = 0, = 12 , ,
= 12 , and = 1.

168

Chapter 7. Critical phenomena

The algebra involved in this derivation is not difficult but it is a little tedious! It would be a useful
exercise to work through it in full either by hand or with the assistance of a computer algebra system.
For more complex models an analytical solution of this nature becomes impossible. Rather than
solve equations like 7.63 directly, it may be necessary to make approximations such as linearising about
the points of interest.

7.7

Summary
The key points of this chapter are as follows:
1. The critical temperature is associated with scale invariance and is therefore typically described by
power laws, in terms of critical exponents.
2. At the critical temperature, the correlation length becomes infinite.
3. The six critical exponents are not independent, but rather related by two underlying parameters.
One empirical way of explaining these relationships is by invoking the Widom scaling hypothesis
concerning how the free energy responds to changes of scale in temperature or external field.
4. Renormalisation group theory gives a formal explanation for this scaling behaviour. Renormalisation
involves systematically reducing the number of parameters of a model either in real or reciprocal
space.
5. In real-space renormalisation, the renormalisation is interpreted as a renormalisation flow through
the space defined by the parameters of the Hamiltonian. Typically there will be a high- and a
low-temperature attractive fixed point, separated by a critical surface containing the critical mixed
fixed point.
6. Phase transitions occurring in vastly different systems often have identical critical exponents. Thus
phase transitions can be divided into universality classes dependent only on the dimensionalities of
the system and the order parameter.

7.8

Short questions to test your basic knowledge


1. What does it mean to say that the behaviour at the critical temperature is scale invariant?

7.9. Longer questions

169

2. What is the correlation length? Explain why it is small at both very high and very low temperatures.
3. What is renormalisation? Name and describe some blocking algorithms that can be used for this
process.
4. Check that the relationships in table 7.3 are consistent with the data in table 7.2.
5. What is a universality class? What does the existence of universality classes tell us about the
underlying physics of phase transitions?

7.9 Longer questions


1. Consider a 1D Ising model with coupling constant J, and suppose that we renormalise by
decimation by simply removing every second spin. Show that this renormalisation corresponds to
replacing J by
(7.69)
J (1) = 21 log cosh 2J.
Hence show that this model undergoes no phase transition.
2. Consider a 2D Ising model on a square lattice where we incorporate both a nearest-neighbour
coupling K and a next-nearest neighbour coupling L. If we renormalise by decimation, by removing
every second spin from the lattice in checkerboard fashion, it can be shown that to first order we
move through KL-space according to
K (1) = 2K2 + L
L

(1)

=K .

(7.70)
(7.71)

(a) Find the fixed points of this transformation and label them according to the terminology of
this chapter. One of the fixed points we considered in section 7.4 appears to be missing: can
you explain why?
(b) Calculate the matrix M and its eigenvalues.
(c) Hence determine the exponent .

Chapter 8

Reconstructive phase transitions

8.1 Introduction and definition


Many phase transitions occur between two structures that are not closely related (unlike all the other
phase transitions we have studied). This type of phase transition is known as a reconstructive phase
transition.
The properties of a reconstructive phase transition can be summarised:
Because the structures of the two phases involved are dissimilar, the transition will involve
breaking and reforming of bonds. This means that these phase transitions are discontinuous.
Reconstructive transitions are slow, and it is possible for one phase to remain as a metastable
state even within the pressure/temperature range where the other phase should be stable. The
metastable state of diamond existing at ambient pressure and temperature, where graphite is the
true stable phase, is a good example.
Reconstructive transitions tend to require nucleation and growth of the new polymorph, which in
turn depends on the rate of atomic diffusion.

8.2 Examples
A number of examples are well-known. These include the phases of carbon (graphite and diamond),
H2 O ice, and iron (bcc and fcc). A surprisingly large number of materials including pure elements,
complex oxides, salts and molecular crystals undergo reconstructive phase transitions, particularly
when we consider pressure as well as temperature as an external variable.
171

172

Chapter 8. Reconstructive phase transitions

NaCl

1 Now

we need to be specific as regard which free energy we


use; for studies with pressure as the variable we need to
consider constant pressure, which means we need to use the
Gibbs free energy.

NaCl has, at ambient temperature and pressure, the crystal structure that is named after it; it has a
face-centred cubic lattice with Na at the origin and Cl half way along the cube edge. In the NaCl
structure, each ion has 6 neighbours of opposite type, in perfect octahedral coordination. At high
pressure (5 GPa) it transforms to the CsCl structure (primitive cubic with one atom at the origin of the
unit cell and the other in the centre), in which each atom now has 8 neighbours of the opposite type in
square coordination.
The NaCl structure has the lowest lattice energy (the potential energy of the crystal with all atoms at
rest). It also is slightly less dense (higher molar volume) than the CsCl structure (this statement depends
on the relative sizes of the cation and anion; in the case where they have identical size, we know from
the packing factions of the simple cubic close packed and body centred cubic structures of the elements
that the former has the higher packing fraction and hence lower molar volume).
We can write the free energy G1 at zero temperature called the enthalpy H as
G ( T = 0) = H = E + PV

(8.1)

The NaCl structure has a lower value of the internal energy E (that is, it is more negative) than the CsCl
structure. However, the volume V is lower, the PV term will mean that the difference in free energy
will be reduced at higher pressure. If we denote the two phases by the symbols and , the two free
energies will become equal at a transition pressure Pc given by
E + Pc V = E + Pc V Pc =
Pressure (GPa)

20
Melt
Diamond
10

(8.2)

Nothing clever about this really, and indeed, this provides the complete basis for understanding many
reconstructive phase transitions.

Carbon
Graphite

E E
V V

2000
4000
Temperature (C)

Figure 8.1: The phase diagram of carbon.

Carbon exists in many states when we consider the nanoscale structures (buckyballs, nanotubes), but
the two well-known bulk phases are graphite and diamond. These could not be more different; they
use their electronic structure in quite different ways (carbon atoms have three neighbours in a plane
in graphite, four neighbours in tetrahedral coordination in diamond), and diamond is held together

8.2. Examples

173

entirely by covalent bonds whereas the sheets of carbon atoms in graphite are held together by weak
dispersion (E 1/r6 ) interactions.

10
stishovite

Phases of silica

The phase called tridymite has a structure that is not dissimilar to that of cristobalite. The connection
is a close mirror of the difference between cubic close packed and hexagonal close packed structures;
based on relative displacements of planes of atoms with hexagonal symmetry, the ccl structure is
described as an ABCABC type of layering, and the hcp structure is described as an ABAB layering.
You can replace layers of atoms by layers of SiO4 tetrahedra in hexagonal 6-membered rings, and you
get a repeating layer, with cristobalite corresponding to the ccp analogue, and tridymite corresponding
to the hcp analogue. In reality it turns out to be hard to made the tridymite phase, and it is always the
cristobalite phase that grows from silica glass.
The coesite phase also consists of SiO4 tetrahedra, but forms a network that is as different from
the other phases as quartz is from cristobalite. On the other hand, the stishovite phase is completely
different. The Si atoms are now in octahedral coordination with oxygen, forming edge-sharing SiO6
octahedra. The structure of stishovite is isomorphic with rutile, TiO2 .

Solid H2 O
104.5 ,

The HOH bond angle in H2 O has a value of


which is close to the ideal tetrahedral angle of
109.47 . In the common crystal phase of ice, the water molecules are bound together in a tetrahedral
framework (reminiscent of silica) with hydrogen bonds forming the linkages. The link to silica is more
than coincidental; just as silica can form a large number of phases with different network topologies, so
can H2 O. The phase diagram is shown in Figure 8.3.
The common form of ice, known as Ih, has a hexagonal structure with the oxygen atoms on the
same places as the Si atoms in the silica phase tridymite, and as the Zn and S atoms in wurtzite. In this

coesite

high () quartz
2
low () quartz
cristobalite
tridymite

liquid

0
400

800

1200
Temperature (C)

1600

2000

Figure 8.2: The phase diagram of silica, SiO2 .

1000
IV
800
Presssure (MPa)

We have already met two phases of silica, SiO2 , namely quartz and cristobalite, in the context of their
displace phase transitions. The full phase diagram is shown in Figure 8.2.

Pressure (GPa)

The pressuretemperature phase diagram is shown in Figure 8.1. This shows the two crystalline
phases together with the melt. For the moment, observe that the phase boundary between the diamond
and graphite phases is virtually a straight line, unlike the melting curve.

VI

600

XII
V

II
400

Liquid
IX

III

200
Ih
0
120

80

40

Temperature (C)
Figure 8.3: The phase diagram of H2 O.

40

174

Chapter 8. Reconstructive phase transitions

Figure 8.4: Network structures of some of the ice phases, showing the positions of the oxygen atoms and the
hydrogen bonds that provide the tetrahedral coordination. Starting from top left in clockwise direction, the figures
show phases Ih, IV, VI and XII.

175

structure, the water molecules remain intact. However, there is some freedom as to the exact orientation
of the molecules. In short, each OH bond points towards one of the 4 neighbouring O atoms. There can
only be one H atom on the hydrogen bond (and there must be one), but this constraint is not enough
to give complete order. A sample configuration is shown in Figure 8.4 Thus the common form of ice
contains a lot of entropy. The ordered phase of ice at ambient pressure has not long been discovered.
The structures of a number of other phases of H2 O are shown in Figure 8.4.

Free energy

8.3. Thermodynamics of reconstructive phase transitions

Transition
temperature

8.3 Thermodynamics of reconstructive phase transitions


Free energy curves

(8.3)

From this equation it immediately follows that the slope of the free energy with temperature gives the
negative of the entropy:
G
= S
(8.4)
T
The fact that entropy is positive means that the slope is negative. and the fact that entropy increases
with temperature means that the curvature downwards will increase with higher temperature. Similarly,
the slope of the free energy with pressure gives the volume:
G
=V
P

(8.5)

The fact that volume is positive means that the slope is positive, and the fact that volume decreases with
pressure means that the slope will becomes smaller with higher pressure.

Temperature

Free energy

As ever, the key quantity is the free energy, but unlike the cases we have seen before, there is no
function for the free energy that includes both phases. Instead, we have a free energy function for each
phase. The equilibrium state is that for which the free energy has the lowest value. Thus the phase
transition occurs when the free energy of one phase becomes lower than that of another with changing
temperature or pressure. This point is demonstrated in Figure 8.5, which shows the effects on the free
energy of both temperature and pressure.
We can repeat this analysis algebraically starting with the differential of the Gibbs energy:
dG = VdP SdT

Transition
pressure
Pressure

Figure 8.5: The variation of the free energy with temperature (top) and pressure (bottom), showing curves
for two phases labelled and . The phase transition
occurs where the free energy curves cross, giving a
continuous variation of the free energy.

Chapter 8. Reconstructive phase transitions

176

Phase diagrams
Pressure

The way we have drawn the free energy curves implies that the phase is the stable phase at lower
pressures and temperatures. We can represent the transition for all temperatures and pressures by a
phase diagram of the form we have seen before, and which we illustrate in Figure 8.6. We have followed
the usual experimental observation that the phase boundary is approximately a straight line. We now
address the question of why this should be.

Temperature
Figure 8.6: Representation of a phase diagram; the
line represents the phase transition between the and
phases.

We expand part of the phase boundary and consider the effect on the free energy as we move from
one point on the phase boundary to another, Figure 8.7. At all points along the phase boundary, the
free energies of the two phases are the same, i.e. G = G . For a small step along the phase line given
by dT and dP, equation 8.3 gives
dG = V dP S dT
dG = V dP S dT

G =G

(dG ) = VdP SdT

(8.6)

Pressure

The fact that dG = dG along the phase line implies that (dG ) = 0, giving the final result

dP
dT

Temperature
Figure 8.7: Small section of the phase diagram, illustrating the construction leading to the Clapeyron
equation.

dP
S
=
dT
V

(8.7)

The equation is known as the ClausiusClapeyron equation (sometimes shortened to the Clapeyron
equation). It relates in a simple way the slope of the phase boundary to the differences in the volume
and entropy of the two phases. If S and V are both approximately independent of temperature
and pressure then the phase boundary will be a straight line. This is found to be the case for many
reconstructive phase transitions. In order to predict the equation of a phase boundary we therefore need
to know one point along the phase boundary and values for S and V.
We now consider whether it is reasonable to suppose that S and V are independent of T and
P. Clearly the volumes of both phases depend on temperature ands pressure generally we expect
volume to increase with temperature and decrease with pressure but it will be a small change. For
temperature, the change will be of the order of 1% over a temperature change of 1000 K.
The entropy is derived from the phonon free energy. In the high-temperature limit, kB T > h , and
assuming an Einstein model (all frequency values can be replaced by a mean frequency value), the free

8.4. Summary

177

energy and entropy for one mole of atoms is given as


F = 3RT ln (h /kB T )

(8.8)

S = 3R (1 ln (h /kB T ))

(8.9)

Subtracting the values for the two phases gives


S = S S = 3R ln( / )

(8.10)

Assuming that the phase has the higher volume, so that V > 0, we might reasonably expect that
phonon frequencies are lower in the phase. Thus we expect S > 0, and thus the slope of the phase
boundary in the PT phase diagram will be positive (although there are exceptions). Since we expect
that the frequencies will change with temperature and pressure the same way in both phases, the ratio
will be approximately a constant.
In fact, whilst we admit that V and S will have a second-order dependence on temperature and
pressure, we might further expect their ratio to be even less sensitive than either difference. The larger
effect will be from pressure. Given that we have said that the phase has larger volume, we might
reasonably expect that the volume of the phase will vary most with pressure, meaning that V will
decrease slightly with pressure. As the volume of the phase changes faster with pressure, we expect
the frequencies of the phonons in the phase to increase fastest with pressure. This will reduce the
size of the ratio / to shift in the direction towards 1. Thus we expect S to become smaller on
increasing pressure, balancing the change in V. In fact this argument does not depend exactly on the
fine details and will be generally true.

8.4 Summary
1. Reconstructive phase transitions involve changes in atomic structure and bonding, and although
there will often be some relationships between the structures of the two phases there is not a
simple groupsubgroup symmetry relationship of the form seen in displacive phase transitions.
2. It is frequently useful to explore reconstructive phase behaviour with the pressure variable as well
as the temperature variable.

178

Chapter 8. Reconstructive phase transitions

3. Often the phase boundary between two reconstructive phases is close to a straight line in pressure
temperature space. The slope is given as the ratio of the differences in entropy and volume of the
two phases via the ClausiusClapeyron equation.
4. The constancy of the difference in entropy along the phase boundary can be understood as arising
from the fact that the difference is given by the logarithm of the ration of the average phonon
frequency in both phases, which will not change much with pressure or temperature.

8.5

Short questions to test your basic knowledge


1. Define a reconstructive phase transition.
2. State and derive the ClausiusClapeyron equation for the phase boundary in pressuretemperature
space.
3. Explain why the phase boundary for a reconstructive phase transition is usually a straight line to
a good approximation.
4. Explain why the phase boundary for a reconstructive phase transition usually has a positive slope
in pressuretemperature space.
5. Give some examples of reconstructive phase transitions.

Further reading

The following books were useful when preparing these notes and may be of interest for your own
further reading:
Binney, J. J., N. J. Dowrick, A. J. Fisher, and M. E. J. Newman (1992) The theory of critical phenomena:
an introduction to the renormalisation group. Oxford: Clarendon Press
Blundell, S. (2001) Magnetism in condensed matter. Oxford: Oxford University Press
Burns, G. and A. M. Glazer (2013) Space groups for solid state scientists, third edition. Waltham:
Academic Press
Dove, M. T. (2003) Structure and dynamics. Oxford: Oxford University Press
Goldenfeld, N. (1992) Lectures on phase transitions and the renormalisation group. Reading, MA:
Addison-Wesley
Muller,
U. (2013) Symmetry relationships between crystal structures. Oxford: Oxford University Press

Nishimori, H. and G. Ortiz (2011) Elements of phase transitions and critical phenomena. Oxford:
Oxford University Press
Salje, E. K. H. (1993) Phase transitions in ferroelastic and co-elastic crystals, student edition. Cambridge:
Cambridge University Press
Toledano, J.-C. and P. Toledano (1987) The Landau theory of phase transitions. Singapore: World
Scientific
Yeomans, J. M. (1992) Statistical mechanics of phase transitions. Oxford: Clarendon Press

179

Das könnte Ihnen auch gefallen