Phase Transitions and Crystal Symmetry (PDFDrive)
Phase Transitions and Crystal Symmetry (PDFDrive)
Phase Transitions and Crystal Symmetry (PDFDrive)
Volume 38
Phase Transitions and
Crystal Symmetry
by
Yu. A. IZYUMOV
and
V. N. SYROMYATNIKOV
Institute of Metal Physics,
Ural Division of the U.S.S.R. Academy of Sciences,
Sverdlovsk, U.S.S.R .
..
KLUWER ACADEMIC PUBLISHERS
DORDRECHT I BOSTON I LONDON
Library of Congress Cataloging in Publication Data
Izlumov, f'U. A. (f'Uri) Aleksandrovich), 1933-
[Fazovye perekhody i simmetriJa kristallov. English]
Phase transitions and crystal symmetry I by Yu.A. Izyumov and V.N.
Syromyatnikov.
p. cm. -- (Fundamental theories of physics)
Rev. and enl. translation of: Fazovye perekhody i simmetriJa
kristallov.
Includes bibliographlcal references.
1. Crystallography. 2. Phase transformations (Statistical
physiCS) 3. Symmetry (Physics) I. SyromJatnikov, V. N. (Vladimir
N';i<.o1aavlch) II. Izlumov. ro. A. (IUri1 Aleksandrovich), 1933-
Fazovye perekhody i simmetriJa kristallov. III. Title. IV. Series.
QD921.19413 1990
548' .5--dc20 89-71632
CIP
ISBN-13: 978-94-010-7357-8 e-ISBN-13: 978-94-009-1920-4
DOl: 10.1007/978-94-009-1920-4
Preface . . . . . . . . . . . . Xlll
Notation . . . . . . . . . . . XIX
Appendix 439
Index . . 441
Preface
About half a century ago Landau formulated the central principles of the phe-
nomenological second-order phase transition theory which is based on the idea
of spontaneous symmetry breaking at phase transition. By means of this ap-
proach it has been possible to treat phase transitions of different nature in
altogether distinct systems from a unified viewpoint, to embrace the aforemen-
tioned transitions by a unified body of mathematics and to show that, in a
certain sense, physical systems in the vicinity of second-order phase transitions
exhibit universal behavior.
For several decades the Landau method has been extensively used to an-
alyze specific phase transitions in systems and has been providing a basis for
interpreting experimental data on the behavior of physical characteristics near
the phase transition, including the behavior of these characteristics in systems
subject to various external effects such as pressure, electric and magnetic fields,
deformation, etc.
The symmetry aspects of Landau's theory are perhaps most effective in
analyzing phase transitions in crystals because the relevant body of mathemat-
ics for this symmetry, namely, the crystal space group representation, has been
worked out in great detail. Since particular phase transitions in crystals often
call for a subtle symmetry analysis, the Landau method has been continually
refined and developed over the past ten or fifteen years. A general survey of
the various trends in the evolution of Landau's theory in the past two decades
is presented at the end of §l. The object of the present book is to provide an
exhaustive account of the current state of Landau's theory of phase transitions
as applied to crystals. The particular physical nature of a phase transition itself
is of no fundamental importance, since the various (structural, magnetic, ferro-
electric, and other) transitions are described in terms of a unified scheme based
on the mathematical foundations of the theory of space group representations.
XIII
XIV PREFACE
analysis at a phase transition. As for the reader who has just started studying
phase transition theory, he will be able to master methods of analysis and to
learn how to apply these in solving particular problems. The reader who is
unconcerned with the mathematical machinery of phase transition theory and
wishes only to become acquainted with the most fundamental results of the
theory is recommended, on reading the first introductory chapter of the book,
to start reading the last four chapters.
The authors wish to express their sincere appreciation to Prof. I.E. Dzyalo-
shinsky for his support of the idea to write this book, for the numerous bits
of advice as to the selection of the topics involved, and for the criticism and
discussions that have helped to improve the text.
Preface to the English Edition
Nearly five years have passed since the Russian edition of the book appeared.
In this time span some problems of the phenomenological theory of phase tran-
sitions in crystals have been the subject of intensive research and have evolved
into new independent provinces. Three major areas have arisen: the theory
of reconstructive phase transitions, aspects associated with the exploitation of
color symmetry, and the theory of incommensurate structures.
It is these new areas of research activity that have received considerable
attention in the present English edition (Chapters 8-10). Chapters 8 and 10,
dealing with reconstructive transitions and color symmetry, have been writ-
ten especially for the English edition. Chapter 9, devoted to incommensurate
structures, has been substantially revised and enlarged.
Chapter 8 treats martensitic phase transitions from the b.c.c. to the f.c.c.
and h.c.p. phases. These transitions are certainly first-order transitions and
go without the group-subgroup relation. However, a decisive role in these
transitions attaches to strains, from which one may construct the order param-
eter and use the expansion of the thermodynamic potential in powers of it, in
the spirit of Landau's second-order phase transition theory. In describing the
transition itself and the characteristic pretransient phenomena, an important
role pertains to inhomogeneous states, which are soliton solutions of non-linear
equations of elasticity theory. Chapter 8 provides a systematic review of recent
developments in the problem of martensite transformations.
Chapter 10 treats a set of phase transitions whose description calls for
color symmetry. Here belong transitions to incommensurate structures and
quasi crystals and also phase transitions in systems described by a quantum
mechanical order parameter. Traditionally, color symmetry is used to describe
and classify structures. This chapter lays emphasis on the more important
physical meaning of the color groups as a mathematical machinery that enables
one to ascertain the role of physical interactions responsible for the occurrence
xvii
xviii PREFACE TO THE ENGLISH EDITION
Yuri Izyumov
Vladimir Syromyatnikov
Notation
xix
xx NOTATION
The behavior of the various systems in the vicinity of second-order phase transi-
tions can be understood on the basis of the order parameter concept introduced
by Landau [1,2]. The order parameter (OP) TJ specifies a new physical property
that arises in a system as a result of a phase transition from the initial phase in
which this property was absent. Thus the OP TJ is defined so that TJ is equal to
zero on one side of the phase transition and is finite on its other side. Since the
phase transition at the transition point To itself is continuous, the OP TJ = 0
and increases monotonically in moving away from this point.
In the vicinity of the phase transition point the thermodynamic potential
~, which depends on temperature T and generalized forces X, should be char-
acterized also by the value of the OP and may, by virtue of its smallness in this
vicinity, be expanded in a power series of TJ. As is well known currently [2], this
expansion holds in the vicinity of To, except for the narrow interval near the
phase transition point itself, where OP fluctuations play an important role. On
the assumption that we are outside this interval, we represent the expansion as
(1.1)
where ~o is the value of the potential in the initial phase and r, v, u are certain
parameters of the system dependent on T and X.
1
2 CHAPTER 1
0<1> = 0, (1.2)
017
0 2<1>
07}2 > O. (1.3)
Equation (1.2) clarifies why the expansion (1.1) involves no linear term in 7}.
It will be assumed that in the vicinity of the phase transition the coefficient
u is positive for any T and that 1> reverses sign in accordance with the law (1.4).
Under these conditions the occurrence of a finite value of the OP 7} becomes
thermodynamically favorable. Indeed, equation (1.2) may be rearranged in the
form
(1.6)
for T < To. Thus the temperature T = To is the branch point of the solution of
the OP equilibrium-value equation. Implying the T and X dependence of the
quantities l' and u, we will call this equation an equation of state. The situation
of interest is depicted also in Fig. 1.1, where the occurrence of finite OP values
for T < To is described by the appearance of thermodynamic potential minima.
PHENOMENOLOGICAL PHASE TRANSITION THEORY 3
T>Ta T< To
T
o
Fig. 1.2. Temperature dependence of the order parameter m the
thermodynamic theory of second-order phase transitions.
VVe see how Landau's simple arguments lead to the fundamental result
(1.7) that the OP is temperature dependent (Fig. 1.2) near the second-order
phase transition. vVe emphasize that this result applies to the case where the
expansion of the thermodynamic potential in powers of the OP contains no
cubic term in TJ. Now we expressly consider the case where such a term is
present and verify that it gives rise to a first-order phase transition when the
continuity at the transition point is broken and an OP discontinuity arises.
We write down the equation of state (1.2) for the expansion (1.1) including
the cubic term
(1.8)
There are two non-zero solutions:
3v
18u 12 - ~
3v
TJ± = --± 2u'
(1.9)
8u
which will be real at a positive discriminant, that is, when
TA being the superheating temperature of the ordered phase and TB the su-
percooling temperature of the disordered phase (we assume that the ordered
phase exists at higher temperatures than the disordered phase).
A true phase transition occurs at temperature TB < Te < TA. The value
r = re that corresponds to Te and the equilibrium OP value 'f} = 'f}e are found
by simultaneously solving the equation of state (1.8) and from the equality of
the thermodynamic potentials in the ordered and disordered phases
(1.12)
PHENOMENOLOGICAL PHASE TRANSITION THEORY 5
This gives the value of the phase transition temperature and the value of the
OP jump at the transition point
v2
Te = To - - 4I '
uro
1]e = - -.
V
2u
(1.13)
It can be readily verified that the resulting 1]e value corresponds to the largest
(in magnitude) of the two possible OP values for r = re (Fig. 1.3). The state
with the smallest 1]e value has the largest energy and is therefore metastable.
We have arrived at the conclusion that the presence of a cubic term in
the expansion of <I> makes the phase transition a first-order phase transition.
A second-order phase transition occurs in systems where the coefficient v of
the cubic term is identically equal to zero. As will be seen in the section
that follows, this may be dictated by symmetry considerations. However, if
one allows for the fact that all thermodynamic potential expansion coefficients
depend not only on temperature, but also on generalized forces, it may turn
out that at some point (Te, Xc) the coefficients r and v go simultaneously to
zero. At that point, determinable from the system of two equations,
(1.14)
the phase transitions will be of second order. But that will be an isolated
point in the phase diagram on the (T, X) plane, whereas for a system with
v == 0 there is a phase transition line Te = Te(X) which is determined from one
equation,
(1.15)
If there are several generalized forces X, then one should certainly talk of some
surface in the space of temperatures and all generalized forces X rather than
a phase transition line. Instead of the isolated point of equations (1.14), it is
then possible to determine some line or surface of lower dimensionality.
We conclude this elementary thermodynamic analysis by determining the
regions of stability of the ordered and disordered phases in the space of the
potential parameters r, v, u, . ... We wish to specify concretely the stability
regions on the (r, u)-plane at fixed values of v. Thus at v = 0 we have, by
virtue of the foregoing, a simple phase diagram such as that sketched in Fig.
1.4. For <I> involving a cubic term, the situation described above is illustrated
by Fig. 1.5. The solid line refers to first-order phase transitions; it corresponds
to the solution
v2
r = re = - (1.16)
4u
6 CHAPTER 1
of the equation of state (1.8) and the energy equality (1.12) of the two phases.
The dot-and-dash line is the boundary of stability of the ordered phase, ob-
tained from the condition (1.3). This boundary is determined by the equation
9v 2
1'= - (1.17)
32u
and coincides with the boundary (1.10) of existence of real solutions to the
equation of state. Stable states are those with 'fJ f. 0, which correspond to the
region lying below the dot-and-dash curve. On the other hand, it follows also
from the condition (1.3) that the stability region of the phase with 'fJ = 0 is
defined by the inequalities l' > 0 and u > O. Consequently, the stability regions
of the phases with 'fJ = 0 and 'fJ f. 0 overlap on that part of the plane that
is bounded by the dot-and-dash curve and the abscissa axis. The region of
overlap corresponds to metastable states, t.he first.-order phase t.ransit.ion curve
passing within t.his region. When v -+ 0 Fig. 1.5 transforms into Fig. 1.4.
r
LI
Fig. 1.4. Phase diagram in the absence of a cubic term in the ex-
pansion of the thermodynamic potential. v = 0, phase with 'fJ = 0; 'fJ,
ordered phase with order 'fJ f. O.
r
\
\.71=0
'-.'-.....-
u
Fig. 1.5. Phase diagram in the presence of a cubic t.erm in the expan-
sion of the thermodynamic potential.
PHENOMENOLOGICAL PHASE TRANSITION THEORY 7
Landau was the first to pay attention to the fact that at any second-order phase
transition a symmetry change occurs in the system. Let the initial (more
symmetric) phase be specified by the symmetry group G, which leaves the
macroscopic density function p(1') of the system invariant. The density function
may be scalar, vector-valued, tensor-valued, etc., according to the physical
meaning of the quantity varying at the phase transition.
Since the state changes continuously at second-order phase transitions,
the symmetry of the new phase may lower only due to the loss of part of the
symmetry elements and will be described by a group G 1 that is a subgroup of
the initial group G. Mathematically, this may be represented as
(1.18)
From the condition that <1> be minimum, it follows that the linear terms in
this expansion cancel out. The structure of the quadratic terms is determined
by the fact that only one second-order invariant ~A (CA)2 exists for each IR
D V according to which the quantities CA transform under the action of the
group G elements. Higher-order (cubic, fourth-order, etc.) expansion terms
are invariant polynomials of the corresponding order; they may be found for a
given group G using familiar mathematical recipes.
With regard to the expansion (1.20) one may apply the same lines of argu-
ment as those that we have used in the elementary Landau analysis outlined in
the beginning of this section. The quantities rV depend on temperature and ex-
ternal forces through the generalized thermodynamic quantities X. At T > To
all the rV should be positive, so that the equilibrium value of the parameters
C A is equal to zero. If one of the quantities rV at T < To becomes negative,
the occurrence of a state with a finite value of CA, that is, the appearance of
a dissymmetric phase, is energet.ically favorable. As the temperature is varied
several quantities may, in principle, vanish one after the other; however, the
temperature T = To at which one of the quantities rV goes to zero for the first
time is the phase transition temperature. This temperature is determined from
the equation
(1.21)
where the CAfor the responsible representat.ion D V is labelled 1]A and the index
of this representation is omitted. Accordingly, the density function (1.19) of
the dissymmetric phase is expressed also only in terms of the quantities
It is the mixing coefficients 1JA of the basis functions of the responsible IR that
are called order parameter components; the number of these components is
equal to the dimensionality of the responsible IR. This definition is a general-
ization of the order parameter 1J concept originally introduced by Landau. The
equilibrium value of the order parameter components is determined from the
requirement that the expansion (1.22) be minimized with respect to 1JA and
specifies the dissymmetric phase density function (1.23).
Equations (1.22) and (1.23) express the hypothesis by Landau that a phase
transition should occur according to one IR. It may happen, however, that in
the expansion (1.20) it is not one, but two coefficients rV at once that may go
to zero at T = T c , for example for the IR's Dl and D2. The equations
(1.24)
define not the phase transition line Tc = Tc(X), as in the case (1.21), but
the point (Tc, X c). In the vicinity of this point it is necessary in the thermo-
dynamic potential expansion (1.20) to retain the coefficients CX for both IR's.
The situation described corresponds to the case of interacting order parameters.
Both IR's are responsible in this situation, so that at the point (Tc, X c) itself the
phase transition occurs according to a reducible representation of the symmetry
group of the system. The dimensionality of the total OP at the point (Tc, X c)
increases. This additional degeneracy may be accidental (as the crossing of two
lines (1.24) on the (T, X)-plane) or due to special symmetry considerations.
The general scheme of Landau which we have outlined above enables one to
find all admissible dissymmetric phases liable to arise from a given phase via
a second-order phase transition. The corresponding analysis reduces to the
construction of an expansion of the thermodynamic potential in powers of the
OP transforming according to an IR of some group G, followed by minimization
of the potential.
Landau has solved, in general form, the problem as to which of the initial-
group IR's cannot give rise to a second-order phase transition [2]. As was
clear from the first subsection, the presence in the thermodynamic potential
of cubic terms leads inevitably to a first-order phase transition. Therefore,
the condition [1] which restricts the list of IR's describing the second-order
phase transition consists in the requirement that the IR allow no third-order
invariants. Since the third-order quantities constituted by the coefficients CX
10 CHAPTER 1
(1.25)
The original Landau theory assumed the dissymmetric phase arising from the
phase transition to be homogeneous. Lifschitz [3,4] demonstrated that for spa-
tially inhomogeneous phases to occur, it is necessary that the thermodynamic
potential involve terms containing OP derivatives with respect to the coordi-
nates. Linear invariants in derivatives came subsequently to be called Lifschitz
invariants (their role in phase transition theory will be discussed in detail in
Chapter 8).
Following Landau and Lifschitz's fundamental writings and a rigorous
group-theoretic exposition of those papers in the book by Lubarskii [5], a fur-
ther basic step in the development of phase transition theory was made by
Dzyaloshinsky [6,7] who constructed a theory of incommensurate phases in
crystals with reference to the example of long-period magnetic structures. He
was the first to show that the various modulated phases in crystals may be
obtained as a solution of the non-linear differential equation that results from
minimization of a thermodynamic potential comprising OP gradients. The non-
linear differential equation coincided with the mathematical pendulum equa-
tion, and analysis of its solutions led to a soliton picture of the incommensurate
to commensurate phase transition. It was noted also by Dzyaloshinsky that
the wave-vectors of incommensurate phases actually possess different symme-
try (with the wave-vector length varying along a fixed direction), a fact that
leads to a sequence of modulated-phase transitions giving rise to intermediate
commensurate phases. This array is currently known as the "devil's staircase",
the name having already become rooted in the literature. The theory of incom-
mensurate phases was further evolved by Ishibashi [8], Dvorak [9], Bak [10],
Levanyuk [11], and Sannikov [12].
Over the past two decades the Landau theory has been developing in two
directions - evolving the physical picture of phase transitions and working out
PHEN01VIENOLOGICAL PHASE TRANSITION THEORY 11
fact has been made by Toledano [34], Michel [35], and Kopsky [36].
The above trends of the phenomenological phase transition theory were
based on the utilization of crystal space groups. Recently a new trend has
arisen which employs color groups (Wolf [37]; Janssen, Janner [38,39]; Litvin,
Kotzev, Birman [40]). The mathematical machinery of the color groups has
proved convenient in describing magnetic phase transitions [40], incommensu-
rate phases [37-39], and also quasicrystals [41].
The aforementioned trends in the development of the Landau theory dis-
regard the fluctuations, so the conclusions of these theories hold good outside a
narrow vicinity of the phase transition boundaries established by Ginzburg [42]
and Levanyuk [43]. In considering the role of fluctuations that exist within the
critical region, it is also very helpful to exploit the mathematical machinery of
the I-groups.
The present book expounds the major trends in the development of the
Landau theory and outlines the contemporary status of the phenomenological
treatment of phase transitions in crystals and the current state of the theory's
working methods.
As the present book deals with phase transitions in crystals, the theory's math-
ematical background rests on space group representations. Assuming a certain
mathematical sophistication on the part of the reader at a level of Landau and
Lifschitz's course [2], we restrict ourselves in this section to the most impor-
tant information from space group representation theory needed for the further
exposition. A more detailed treatment of the theory of the space groups them-
selves and of their representations can be found in many books, of which the
reader is recommended to refer to the monograph [44] where the notation em-
ployed is the same as that adopted here.
The space group IR's are specified by wave-vectors defined with the help
of reciprocal crystal lattice vectors. Labelling the primitive translation vectors
by t 1 , t 2 , t 3 , the primitive reciprocal lattice vectors b l , b 2 , b 3 are defined, as
is known, by the expressions
PHENOMENOLOGICAL PHASE TRANSITION THEORY 13
(2.6)
The number of star arms I" is evidently equal to the index of the subgroup G"
in the group G and does not exceed, for all the space groups, forty-eight - the
maximum number of elements in the crystal point groups.
Equation (2.6) allows one to construct matrices D{"}v (g) of the IR's of
the space group G from the group G" ofIR matrices d"V(g). The dimension of
these matrices will obviously be lvi,,, with Iv the dimension of the IR d"V. The
relation between these is given by the equation
Here A, J1. = 1,2 .. . Iv are the matrix indices of the representation d"v, Land M
denote the numbers of the representatives in the decomposition (2.6) and, at
the same time, are the numbers of the star arms. Thus, taking the wave-vector
group IR matrices from tables [45,46], formula (2.7) enables one to obtain the
matrices of the appropriate IR of the entire space group.
The basis functions of the group G" IR are Bloch functions of the form
(2.8)
where the u~V(r) are periodic functions with respect to crystal translations.
Under the action of a group G" element the set of 1jJ~v (A = 1,2, ... , Iv) trans-
form according to the equation
Iv
g1jJ~v =L d~>" (g )1jJ~v. (2.9)
p.=1
A basis of the IR of the entire group G is generated by the set Inlv of Bloch func-
tions {1jJ~lV}, {1jJ~2V}, ... , {1jJ~I"V} prescribed on all star arms "'1,"'2, ... ,"'L,
... , "'1". These functions transform under the action of a group G element g
by means of the equation
(2.10)
PHENOMENOLOGICAL PHASE TRANSITION THEORY 15
In the section that follows we will show in what fashion one should con-
struct basis functions of a crystal space group IR from physical quantities
specifying the state of a crystal after a phase transition. The method rests on
the familiar group theoretic formula for the projection operator [5]
(2.11)
whose action on an arbitrary function from some space yields the function
(2.12)
(2.14)
where XV(g) is the character of the IR D V and X(g) IS the character of t.he
reducible representat.ion D.
16 CHAPTER 1
(2.15)
The summation here is performed over the various partitions of the number
n = L:s r s qs (rsand qs being positive integers). For the initial value this
formula assumes the specific form
1 1 1
"3 XII (g3) + 2"XIl(g2)xIl(g) + 6Xll
3 3
[XII ](g) = (g),
References
The phase transition from the initial phase of the crystal gives rise to a state
which at the microscopic level may be specified by the occurrence of some
spontaneous property on each atom. This property is described by a scalar, a
vector or a tensor. Thus, for example, a static magnetic moment arises on the
atom in a magnetic phase transition and each atom may therefore be specified
by an appropriate pseudovector. In the case of a structural phase transition
by displacement (displacive transition) it is possible to assign to an atom in
a dissymmetric phase a polar displacement vector with respect to its position
in the symmetric phase; the assignment of the polar displacement vector on
each wholly characterizes this phase. If the magnetic ordering is accompanied
by some structural distortions, two vectors should be specified for each atom
in the dissymmetric phase: the magnetic moment pseudovector and the polar
atomic displacement vector. It is better in such a situation to ascribe to the
atom some tensor of rank two. For an ordering-type phase transition each atom
is characterized by a scalar quantity that represents the relative probability of
occupying determinate positions in the crystal.
In the general case one may talk of the occurrence of a certain tensorial
characteristic on each atom, implying that the scalar and the vector are tensors
of rank zero and rank one respectively. The dissymmetric phase resulting from
the phase transition is assigned by specification of the corresponding atomic
characteristic on each crystal atom, information on these characteristics be-
ing extracted from experiment. In keeping with Landau's idea, the state of
the dissymmetric phase may be characterized by a small number of quantities
19
20 CHAPTER 2
Consider a crystal that contains N primitive cells. Let there be in each cell U'R
identical atoms which occupy the same position in the group G and whose state
in the crystal is specified by an R-component tensor of arbitrary rank. Then
the state ofthe crystal as a whole is described by an RU'RN-dimensional column
that indicates the values of all tensor components on each atom. Let the state
of the crystal be characterized by a definite wave-vector K,. To explore the
transformation properties of such a multidimensional vector under the action
of group G K elements, it is convenient to introduce unit vectors of an RU'RN-
dimensional space:
L:$
¢),.f3 = ~f3 eiKtn , (3.1)
n
where ~f3 is an RU'wdimensional column with all the components equal to zero,
except one, which is equal to unity and corresponds to the jth atom of the zero
cell and to the ,Bth tensor component. The symbol L:~ denotes the direct sum
over all N crystal cells obtained from the zero cell by translations tn.
The <p~f3 are the eigenfunctions of the translation operator, for, by defini-
tion (3.1), the following equation holds:
(3.2)
Thus the functions <p~f3 are Bloch-type functions, and this is precisely what the
assignment of the index K, to them signifies.
MICROSCOPIC REALIZATION OF ORDER PARAMETERS 21
Let us see how the functions 1jJ{f3 transform under the operation of group
G" elements g = {hlt n }. An element of the group G" transfers an atom with
number j and coordinate rj from the zero cell (in the general case) into an
atom with number i of another cell:
(3.3)
(3.4)
Besides, the ,B-component of the atomic tensor transforms in terms of the other
a-components according to the tensor transformation law. Thus the result of
the action of the operator T(g) on the function ¢~f3 may be written in the form
D'f! (h) being the tensor transformation matrix under the action of the rota-
tional part h of the element {hlth}. The numbers of the atoms in this case
change according to the relation (3.4).
The vector thus obtained may be resolved in terms of the orbits of the
space introduced and the result may be represented as
(3.7)
where the 8-symbol takes into account the condition (3.4) for the permutation
of atoms under the action of the element g.
The tensor representation matrices have dimension R(J'R x R(J'R and it is
clear that the tensor representation should in the general case decompose into
group G" IR's d"V. Symbolically, this may be written as
(3.8)
22 CHAPTER 2
where X"V is the character of the IR d"V and X"R (g) is the character of the
tensor representation. According to equation (3.7), the tensor representation
character may be written as
(3.10)
with
(3.11)
(3.12)
In the expression (3.9), where the summation is carried out over all ele-
ments of the space group G", we may take the sum over the translations, so
that the equation contains only the sum over the group G" elements g per-
taining to the zero block. We label the set of these elements by G~, and the
number of these elements by IIG~II. Instead of the formula (3.9) we will then
have an expression that is convenient for practical calculations. Namely,
where d~~(g) is a matrix element of the vector representation d"v, and 'l/J is
some starting function. As the starting function we take the unit vector rP~f3
of the space of RURN-dimensional vectors and determine the action of the
MICROSCOPIC REALIZATION OF ORDER PARAMETERS 23
operator T(g) on the unit vector using equation (3.6). Then the expression
(3.14) becomes
L
1/;~V = 2::>h~(g) {dR(g)}ia,ji3tP~a. (3.15)
9 ia
The summation over integral translations may be performed using the explicit
form (3.7) of the matrices dR(g). This leads to an expression which involves a
summation over group G" zero-block elements g E G~ only:
.I.I<V
'1-'>. - L... >'/1-
""'" e-l< aij(g)8',g)
_ ""'" dl<v (g) L... . .Vai3(g)A..ia
R '1-'1<' (3.16)
9 ia
Taking into account the explicit structure (3.1) of the expression for the
vector column tP~a, we represent the basis function 1/;~v in the same form:
(3.18)
The quantity 1/;a(~vli) is the atomic component of the basis function. The
indices /-l, j, and /3, enclosed in square brackets in the expression (3.19), should
be fixed. These indices actually determine some start (initial function) for the
construction of a basis function. An alteration of the initial (starting) function
either gives rise to a new system of basis functions (for the cases where a given
IR dl<v is contained more than once in the tensor representation dR ) or leads to
a zero. If the atoms in the crystal occupy several positions of multiple points,
it is necessary in the formula (3.19) to substitute as the index j the number
of one of the atoms of a given position in order to obtain a set of 1/;a(,,/ Ii)
functions for the entire position and then, proceeding in the same fashion, to
obtain functions for the atoms of a different position. For practical calculations
of the expressions (3.19) and (3.13) one has preliminarily to compile a table of
the change in the numbers of the crystal primitive-cell atoms under the action
of the group G I< elements and to find the vectors of the returning translations
24 CHAPTER 2
aij for each atom. The necessary IR matrices are to be taken from relevant
tables (for example, [1]).
The equations (3.19) and (3.13) are fundamental working formulas by which
one calculates the basis functions describing a phase transition in a crystal. In
actual fact, for the prescribed wave-vector K, and group G" IR d'<l', the formula
(3.19) should be employed to calculate the atomic components 'IjIO'(AVli) for
zero-unit-cell atoms belonging to the same multiple-point position. In doing
so it should be borne in mind that the position of the multiple points of the
parent-crystal space group G with respect to the wave-vector group may split
up into individual assemblies of atoms that transform into each other (to within
integral lattice translations). We call each of these individual totalities an orbit
with respect to the group G". The calculation of the atomic components of the
basis function 1,&0' (A V Ii) for each orbit should apparently be done independently.
To this end, some atom j of a given orbit should be chosen as the start.
In the formulas (3.19) and (3.13) the summation is extended over the
elements of the group G". These formulas, however, may be modified into a
form where the summation is performed only over some of the G" elements,
more precisely those that leave the starting atom fixed. They form what we
call the stabilizer of the space group G". More rigorously and in more general
terms, it is the subgroup H of the space group G that is referred to as the
stabilizer of the space group G; the subgroup H maps a given atom onto itself
(to within integral translations). By transforming the basic formulas (3.13) and
(3.19) in such a way that they contain only the summation over the stabilizer
elements, we substantially facilitate computational work.
Thus we consider some orbit with respect to the group and ascribe to the
starting atom the number j = 1. We denote the stabilizer of the first atom by
H" and decompose the group G" into cosets relative to the subgroup H,,:
(3.20)
The representative element gi evidently transforms the first atom into a number
i atom (gi being the identity element of the group).
To begin with, we transform the phase factor entering into the expressions
(3.11) and (3.19). Recalling the relation (3.3), which serves as the definition of
the returning-translation vector aij(g), we rearrange the aforementioned factor
MICROSCOPIC REALIZATION OF ORDER PARAMETERS 25
in the form
(3.21)
ilVe now allow for the fact that the coordinates of any atom of a given orbit
may be expressed in terms of the coordinates of the first atom under the action
of the representative elements rj = gjr1 - aj1(gj), ri = girl - ai1(gi). Using
these expressions and multiplying each vector in the scalar product in equation
(3.21) by g;l, we obtain for the quantity iij(g) two important expressions:
(3.22)
where we have taken account of the fact that g;l", = "', since the element g;l
belongs to the wave-vector group G/<.
Consider now the expression (3.9) which defines the constitution of the
tensor representation. Substitute into it the expression (3.10) for the character
x'R(g) of this representation. Using the notation (3.21) for the phase factor, we
write (g E G,,)
(3.23)
9 j
The sum over j gives the number of atoms in an orbit. This number coincides
with the index of the subgroup H", which is equal to IIG"II/IIH"II. On extending
the summation in the last expression over the translations of the group H", we
write the final expression for nV in the form
where the summation is performed only over the zero block of the stabilizer
group.
26 CHAPTER 2
We now pass on to the expression (3.19) for the IR basis function. 'We re-
turn to the summation over all group G" elements g, rearrange this expression,
and take the first atom (j = 1) to be the starting function:
(we have introduced also the notation (3.21) for the phase factor). We represent
the o-symbol in the form: Oi,gl = 01,g-:-1 g1 , whence it is seen that the element
h = gi 1 g belongs to the stabilizer. 'Thus, just as in the expression for n",
the formula for the basis function contains actually the summation over the
stabilizer elements of the first atom. Removing the o-symbol in equation (3.25)
and using the second of the expressions (3.22), we then write (g E giH,,):
1/J"(~"li) = N- 1 L d~[1'1(ghi1(gihll(gi1g)D~[i31(g).
9
The resulting expression may be worked out (on extending the summation over
the translations of the group H,J to (h E H2):
In principle, this result already solves our problem of calculating the atomic
components of the basis function in terms of the stabilizer. However, by passing
on to the matrix representation, a still more convenient formula may be derived
from equation (3.26). The right-hand side of equation (3.26) involves a direct
•
matrix product d""(gih) x DR(gih). Representing each of the matrices as the
product of matrices of individual group G" elements
(3.27)
where (h E H2)
\Ifl''' = L(d""(h) x DR(h)hl1(h) (3.28)
h
element gi. Each non-zero column of the matrix thus calculated gives a basis
function atomic component that transforms by the IR d''''. A total of nV such
columns should be obtained, in keeping with the multiplicity factor of a given
representation included in the tensor representation. Thus the calculation of
the basis functions for all orbit atoms reduces to multiplying the matrices
together, and the calculation of the matrix wi v presupposes summation over
the stabilizer elements only.
The advantage of the formulas (3.27) and (3.28) for the basis function
as compared to the standard expression (3.19) is that it suffices to manipulate
solely with elements of the stabilizer H" and with representative elements in the
decomposition of the group C" relative to H". In these calculations one does
not have to perform the tedious procedure of compiling a table of atomic per-
mutations under the action of all group C" elements and finding the returning
translations aij (g). The stabilizer method has particularly great advantages in
those cases where the number of atoms in the orbit is large and the number of
elements in the stabilizer is, accordingly, small. In this case the matrix w~v is
easy to calculate by the formula (3.28) and calculating the atomic components
of the basis functions reduces to the matrices being simply multiplied together
according to the formula (3.27).
In the foregoing we have constructed basis functions for the wave-vector group
IR's. The IR's of the entire space group are realized on a basis that incorporates
basis functions for all arms "'L of the star {"'}. It is of course possible to
construct for each arm independently an appropriate individual basis, using
the formulas derived above. There is, however, a simple method of expressing
basis functions for any star arm in terms of the first arm "'1 alone.
To derive this relation of the basis functions, we recall the equation (2.5),
which expresses the star arm "'L in terms of the arm "'1 : "'L = gL"'L. Ac-
cordingly, the wave-vector group C"L is conjugated to the group C", by the
element gL:
(3.29)
A similar relationship will also exist between their representations and, hence,
between their basis functions:
Allowing then for the fact that under the action of the element gL the wave-
vector transforms according to equation (2.5) and the atom changes its number
and cell according to equation (3.4), we can obtain the following expression for
the atomic components of the IR d"LV basis functions:
Here the correspondence between the atom numbers i and i' as well as the
wave-vector ai'i is determined by
(3.33)
A SUMMARY OF FORMULAS
To begin with, we investigate in greater detail phase transitions that are char-
acterized by the occurrence of a phase with a given wave-vector K-, a phase
in which the state of each atom is described by a scalar function. The tensor
representation of the wave-vector group G" given in the general case by the
matrices (3.7) degenerates on a scalar basis into a scalar matrix representation:
( 4.1)
(4.2)
Xp"(g) --~
~ e- i "ajj(g)c5),g).
.. (4.4)
j
MICROSCOPIC REALIZATION OF ORDER PARAMETERS 29
The formula (3.19) for calculating the IR basis functions on the arm Ii, then
reduces to
7f>Wli) = Ld~r"l(g)e-i"aij(g)8i'9[jl,9 E G~. (4.5)
9
The formula (3.2) for a basis function on an arbitrary arm li,L (Ii,L = gLIi,) of
the star {Ii,} assumes the form
(4.6)
crystal). We also note that the atomic component of the basis function has no
other indices except the index A.
We quote formulas for calculating nil and 7f>(~lIli) according to the stabilizer
method. Instead of equations (3.24), (3.27) and (3.28) we have in the scalar
case
n~ = IIH~II-1 LX"II(h)-Yll(h), h E H~, (4.7)
h
.T,,,II
'J! i = /i1 ()d"l1(
gi
*
gi )'T,"II
'J!1 , (4.8)
T> Tc
(a) (b)
Fig. 2.1. Statistic lattice of AuCu type alloys III disordered and
ordered states.
It is a simple matter to see that this ordered structure is described by the
wave-vector
271"
Ii, = -(00 1). (4.10)
a
Indeed, if we locate the origin of coordinates at the position of atom 1, the
translation vectors to atoms 1', 1", I'" will be equal to a(O ~ ~), a( ~ 0 ~) and
a( ~ ~ 0) respectively; the factor e- iKt will be equal to unity for atom I"' and
be -1 for atoms I' and I", as is indicated in Fig. 2.1 by filled, half-filled, and
empty circles. The structure of a real ordered alloy may thus be completely
characterized by the equation
(4.11)
where tn is the translation vector from the initial to the nth atom, and 'f/ is a
parameter that specifies the state at a given (initial) lattice site:
!FA - PBI
'f/= (PA+PB)' (4.12)
with PA and PB the probabilities of the site being filled with species A and B
atoms. These probabilities depend on temperature so that for T ~ Tc 'f/ = O.
For the phase transition described, the quantity 'f/ is an order parameter.
This elementary phase transition in a crystal with one Bravais lattice re-
quires no special group-theoretic analysis.
Fig. 2.2. Relation between the unit cells of the initial phase and those
of the ordered phase for the Nb-H crystal. The dashed straight line
represents the centering vector in the rg cell.
1(~ 0 ~), 2(~ ~ 0), 3(0 ~ ~), 4(~ 0 ~), 5(~ ~ 0), 6(0 ~ ~). ( 4.13)
These six points generate one crystallographic position of the space group. Such
positions are pictured in Fig. 2.3, together with other equivalent positions.
The wave-vector of the superstructure belongs to the six-arm star {",9}
and the wave-vector group evidently has 48/6=8 elements in the zero block.
From tables [1], we find that they are constructed from the following point
symmetry elements:
(4.14)
32 CHAPTER 2
which form the space group elements. The latter turn out to have no accom-
panying translations.
We compile a table of permutations of the first atom (interstitial lattice
site) under the action of the group according to the general equation (3.3). For
example, we have for the element {h410}
~1 0)0
o 1
with a41 (h 4 ) = (I 0 0) the returning-translation vector (to save space, we
have represented it as a line). The operations of all group G" elements are
consolidated in Table 2.1. Along with the returning translations, the table also
furnishes the factors 'Yij(g) calculated according to the formula (3.21). Now we
pay attention to the circumstance that under the operation of the G" elements
the atoms 1, 2,4, and 5 transform into each other, whereas the remaining atoms
3 and 6 transform independently. Thus the six-fold position breaks up into two
orbits relative to the group G", viz., a four-fold and a two-fold position. Each
orbit now needs to be considered separately. For the first orbit the stabilizer
Hl of atom 1 consists of two elements, for the second orbit the stabilizer H3 of
atom 3 incorporates four elements:
( 4.15)
MICROSCOPIC REALIZATION OF ORDER PARAMETERS 33
Table 2.1
Atomic permutations under the operation of group 0", elements in
Nb-H
g3 3 3 3 3 6 6 6 6
C'I
..., II I 1I I -- 1 II II I
;.0 ai3(g) 000 010 222 222 011 001 222 222
....
0
li3(g) 1 -1 -1 1 -1 1 1 -1
Table 2.2
Irreducible representations of the group of the wave-vector Ii, = ~(11 0)
of the space group O~
T1 1 1 1 1 1 1 1 1
T2 1 1 1 1 -1 -1 -1 -1
T3 1 -1 1 -1 1 -1 1 -1
T4 1 -1 1 -1 -1 1 -1 1
T5 1 -1 -1 1 1 -1 -1 1
T6 1 -1 -1 1 -1 1 1 -1
T7 1 1 -1 -1 1 1 -1 -1
T8 1 1 -1 -1 -1 -1 1 1
34 CHAPTER 2
Table 2.3
Permutation representation basis functions for Nb-H and Ta-H crys-
tals
i 1 2 4 5 3 6
V
Tl 1 -1 -1 -1
T4 1 -1 1 1
T6 1 1 1 -1
T7 1 1 -1 1
T5 1 -1
T7 1 1
wi v = 1. (4.17)
The components of the basis functions on the other orbit atoms are to be found
from the formula (4.8). For example, according to Tables 2.1 and 2.2, the basis
function components for the representation Tl for the atom 2 pertaining to the
first orbit are
MICROSCOPIC REALIZATION OF ORDER PARAMETERS 35
Z= 0 - x
Z=I/4 - A
Z=I/2- 0
z=3/'t -0
(a) (b)
Fig. 2.4. Projection of the crystal structures of the disordered (a) and
ordered (b) phases of the interstitial alloy Ta-H on the (00 1) plane
(D~~ after [11]). Only interstitial sites filled with hydrogen atoms
are shown; the figures correspond to the following heights: +, z =
O·,Ll.,
A Z -- 4' 1. D , z -- 4'.
1. 0 , z -- 2' 3
The calculation results for all the IR's are presented in Table 2.3, which gives
values of the "atomic" components of the basis function at the interstices of
the parent-crystal primitive cell. In the adjacent cells the "atomic" compo-
nents differ in the factor exp( i/d n ), which is equal to -1 for the translations
a( ~ ~ ~), a(±1 00), etc. These data permit one to represent geometrically the
state of an ordered crystal described by basis functions of individual irreducible
representations. In such a portrayal one should bear in mind that all the in-
terstitiallattice sites in the disordered phase at T > Tc are equivalent, and the
probability of their being filled (the occupation probability of these interstices)
is determined by the hydrogen atom concentration C ofthe crystal (the ratio of
the total number of hydrogen atoms to the number of interstices). At T < Tc
these probabilities change and their distribution is characterized by the basis
function. Thus the value of the "atomic" component which is equal to unity
may be treated as an increase of the probability of a given interstitial lattice
site being occupied by hydrogen atoms, and the value equal to -1 denotes a
decrease of the occupation probability. The magnitude of the variation of the
occupation probability is characterized by a temperature-dependent parameter
TJ; it is this parameter which is the OP. The ordering described by the IR 75 of
the star {~g} is characterized by the occupation probabilities of sites 3 and 6;
these probabilities are equal to C + TJ or c - TJ. The corresponding arrangement
36 CHAPTER 2
of the hydrogen atoms in these interstices is depicted in Fig. 2.4. The occupa-
tion probabilities for the other interstitial lattice sites, 1, 2, 4, and 5, remain
equal to C. These structures will conventionally be called "gray" structures.
No "gray" interstices are shown in Fig. 2.4. The symmetry of the ordered
phase that corresponds to the representation T5 of the star {K.9} is described
by the group D~~.
In deciphering the ordered phase structure in the systems Nb-H and Ta-
H, the authors of [11] opted for a "black and white" version: The arrangement
of the atoms at interstices 3 and 6 is the same as that sketched in Fig. 2.4b and
the interstices 1, 2, 4, and 5 are empty. The symmetry of such a structure is
also described by the group D~~. In order to describe such a structure [10], it
is necessary to add to the basis functions T5 of the star {K.9} the basis functions
of the representation T5 of the point r (K. = 0) which are calculated according
to the same scheme.
A SUMMARY OF FORMULAS
Our objective now is to consider phase transitions in which the state of each
atom of the crystal in the dissymmetric phase is specified by a polar vector.
Here belong structural phase transitions of the displacive type, transitions to
the ferroelectric phase, etc.
If we assign to each atom in the dissymetric phase some polar vector, then
not only an atomic rearrangement but also a rotation of the atomic vector oc-
curs under parent-crystal space group transformations. The state of the crystal
as a whole should be described by a 30-N-component column vector (with 0-
the number of atoms in a primitive cell) whose individual components indicate
the projection of the vector for each atom. The 30-N -component unit vectors
generate the basis of the vector representation of the crystal space group. Since
a polar vector may be viewed as a tensor of rank one, we may exploit the re-
sults of §3, which are valid for a tensor of any rank. To this end, we need
just to identify the tensor transformation matrix D'f! (g) with the coordinates
transformation matrix R OI f3(g), for the polar vector transforms as the radius
vector (position vector). Thus the matrix of the vector representation d:;' of
the group G,. is determined by the equation
with Q' and f3 running over the three values: x, y, and z (equation (3.7)).
MICROSCOPIC REALIZATION OF ORDER PARAMETERS 37
(5.2)
According to the general formula (3.19), the atomic components of the basis
function on the arm", are given by the expression (g E G~)
and for an arbitrary arm "'L they are expressed in terms of 1jJ"(~Vli) by the
equation
1jJ"GLV li') = e-i"LBili(9L)LR"I'(hL)1jJI'(~Vli). (5.6)
I'
Now we supply the corresponding formulas employed in the stabilizer
method. Instead of the general expressions (3.24), (3.27) and (3.28), we have
for the vector representation
Fig. 2.5. Unit cell of the Nb 3Sn crystal and the displacement of Nb
atoms at a structural phase transition.
The unit cell atoms in the cubic phase of the crystal occupy the positions:
As can be seen from the figure, at the phase transition the unit cell is
preserved, so the transition is characterized by the wave-vector K, = o. We
demonstrate that the tetragonal phase that arises corresponds to the IR 75
of the group O~. To this end, we calculate the basis functions of this repre-
sentation by invoking the basis of the mechanical representation of the group
GI<=O~.
The group O~ contains forty-eight elements in the zero block:
(5.10)
where the asterisk marks the elements that contain a fractional translation
T = (~ ! !), so that h * = {h IT h}. Under the action of the group O~ elements
MICROSCOPIC REALIZATION OF ORDER PARAMETERS 39
Table 2.4
Permutations of the first atom of species A under the operation of the
zero-block elements of the group
1 1 2 2 5 6 6 5 3 4 3 4
hi3 hi4 hi5 hi6 hh hi8 hi9 hZD hZI hZ2 hh h24
3 4 3 4 1 1 2 2 5 6 6 5
h25 h26 h27 h28 h29 h3D h3I h32 h33 h34 h35 h36
2 2 1 1 6 5 5 6 4 3 4 3
(5.10) the atoms of sort A will transform into each other; the permutations of
the first atom are given in Table 2.4. An inspection of this table shows that the
non-zero block of the stabilizer of the first atom is generated by the elements
H:
(5.11)
which constitute the point group C4v . Using the same table the decomposition
of the group O~ with respect to the stabilizer may be represented, for example,
in the form
(5.12)
where the representative elements h3, h6, hs, h g , hID (not involved in the
stabilizer) transform the first atom into all the other species A atoms which
form one orbit:
We verify that the atomic displacements sketched in Fig. 2.5 are described
by the basis functions of the group O~ two-dimensional IR 75 with K, = O. The
matrices of this representation 75 are written out in Table 2.5. Using the
40 CHAPTER 2
structure of the stabilizer group (equation (5.11)), we can readily calculate the
basis functions of the representation 75. The general formulas (5.7)-(5.9) of
the stabilizer method for the case Ii, = 0 simplify:
For the representation 75 of interest, the formula (5.10) may be written out in
the form
wi = (~ ~) x
V
[R(hd + R(h2) + R(h27) + R(h2S ))
The matrices of the point transformations R(h) are written out in tables [1].
Using the tabulated matrices, we obtain the expression
(5.17)
which determines the basis function components on the first atom (we have
omitted the immaterial numerical factor 4). With the help of equations (5.13)
and (5.15) we now obtain the matrices W:,V for the other atoms of the orbit
and represent the result as
v =
'11 12
,
(1 -t) x (1'
-(
0
0
0
D, w;~,= CI ~n ~I X (
0
0
0
D,
W"V
5,6 --
( (2
_(2 ~,) (11 ooo 0)
X
0
0
. (5.18)
Each column of the matrix W:,V determines the two-component set of the basis
function of a given IR. As can be readily shown from the formula (5.14), nV = 1
for 75, so a unique set of basis functions may be constructed in the crystal under
MICROSCOPIC REALIZATION OF ORDER PARAMETERS 41
consideration. Taking the first column in the expressions (5.18), we write the
atomic components of the basis functions in the form
(for convenience, we have begun here to represent the non-zero column of the
3 x 3 matrices in the formula (5.18) as a three-component row).
The second set of basis functions
W~,2 = ( _f2)
1 x(±l 00), W;,4 = (-1)
f2 x(O ±1 0), W~,6 = ·(-f)
f x(O 0 ±1)
(5.20)
differs from the first in the common multiplier - f and is therefore linearly
dependent. Thus the expressions (5.19) are the basis functions of the two-
dimensional IR T5 on the vector basis.
The complex character of the basis functions (5.19) is a consequence of
the complex form of the matrices of the IR T5. However, this representation is
real-valued and all its matrices may be brought into real form using the unitary
tr ansformation
d(g) -+
d(g) = Ud(g)U+. (5.21)
As U we choose the matrix
(5.22)
Then all the matrices contained in Table 2.5 are modified into
Table 2.5
Matrices of the irreducible representation '15 with K, = 0 for the group
O~ [1] (f = exp 2;i)
(~ ~) (~ f~ )
hI, h2, h3, h4, h5, h6, h 7, h8,
h 25 , h 26 , h 27 , h28 h 29 , h 30 , h 31 , h32
h 9 , hIO,hll,hI2,
h 33 , h 34 , h 35 , h36 c; ~) h 13 , h 14 , h 15 , h 16 ,
h37,h38,h39,h40 (~ ~)
(~ f;)
n , h 23 , h 24 ,
C~ ~)
h 17 , h 18 , h 19 , h 2O , h 21 , h
h 41 , h 42 , h 43 , h44 h 45 , h 46 , h 47 , h48
In addition, the calculation results (5.24) for the basis functions may be
represented in tabular form (Table 2.6) by indicating the three-component
vector for each atom. It is now a simple matter to see that the atomic dis-
placements shown in Fig. 2.5 do correspond to the first basis function of the
representation '15.
Table 2.6
Basis functions of the two-dimensional representation '15 of the group
O~ for A-15 crystal
i 1 2 3 4 5 6
1/;
(5.25)
(5.26)
(5.27)
(5.28)
The group Ok has ten IR's for", = 0, of which four (71-74) are one-dimensional,
two (75 and 76) two-dimensional and four (77-710) three-dimensional; the rep-
resentations with an even number are odd with respect to an inversion and
conversely. 'Ve have the following constitution of the mechanical representa-
tion on the positions 16 (d) and 8 (a) respectively:
(5.29)
44 CHAPTER 2
We now calculate the basis functions for all the irreducible representations
that constitute the mechanical representation. Using the IR table we obtain,
by virtue of equations (5.15) and (5.16), the following matrix
for the one-dimensional representation T4. The column of this matrix deter-
mines the atomic component of the basis function on the atom 1. We label
this component by 'lj!1. On the other atoms of a given orbit, the basis function
components are determined by the action of the representatives (5.26). Thus
we obtain for the ith atom of the position 16 (d) the polar vector
(5.30)
In their entirety these vectors generate the basis function of the IR T4.
0 0 0 0 0 0 0 0 0
0 1 I I 0 1 1 I 0
0 I 1 1 0 I I 1 0
0 I 1 1 0 I I 1 0
\Ii"v
1 -- 0 0 0 0 0 0 0 0 0 (5.31 )
0 1 I I 0 1 1 I 0
0 1 I 1 0 I 1 I 0
0 I 1 I 0 1 I 1 0
0 0 0 0 0 0 0 0 0
which contains the sole line-independent column that determines the atomic
components of the three basis functions on the atom 1. For the other atoms
of the orbit the basis functions are determinable from the formula (5.15) with
the help of equation (5.26). The calculation results for all the IR's involved in
equation (5.29) are summarized in Table 2.7 [13].
As a result of a phage transition by one IR, atomic displacements arise
which may be described by a superposition of the basis functions shown in the
MICROSCOPIC REALIZATION OF ORDER PARAMETERS 45
Table 2.7
Basis functions of irreducible group representations with K, = 0 on the
vector basis (u = 2 - va,
v = 1- va)
i 1 2 3 4 5 6
7
dM = Ln'Md"v. (6.2)
v
The multiplicity factor n'M of the group G" IR may be calculated according
to the formulas (5.3) and (5.4) of the preceding section, subject only to the
correction that the matrix RcxfJ (g) should be replaced by 8g RcxfJ (g). Introducing
the same correction, the formulas (5.5)-(5.9) for calculating the basis functions
of the pseudovector representation become valid too.
A magnetic structure that is specified by a given wave-vector may be
represented as a superposition of pseudovector basis functions of some IR d"v
of the wave-vector group [14] of the initial (paramagnetic) phase of the crystal.
It would, however, be not out of place here to note that the symmetry group of
a paramagnetic crystal (what is known as the paramagnetic group Gl') is not
quite identical to the crystal space group but is a direct product of the group
G by the spin inversion group R:
Gl' = G x R, (6.3)
MICROSCOPIC REALIZATION OF ORDER PARAMETERS 47
(the latter consists of two elements, R = 1, I'). All the irreducible representa-
tions of the group G multiply into a doubled set of even and odd group G1'
IR's with respect to I'. It may be shown that the group G1' magnetic represen-
tation does not incorporate even representations at all and that the magnetic
structure is described only by odd representations of the paramagnetic group.
The formula for calculating the basis functions of odd group G1' representa-
tions coincides with that for the corresponding group G basis functions. For
this reason the magnetic phase transition from the paramagnetic state of the
crystal may be described by the pseudovector basis functions of the IR's of the
space group G [15].
(provided that 2V = U). Thus the magnetic moments form a triangular mag-
netic structure and lie in planes perpendicular to the body diagonals of a cube
(Fig. 2.6).
It is known that this magnetic structure is described by the 3-dimensional
IR 710 (n = 0) of the group O~o. We calculate the basis functions of this
representation on the pseudovector basis. The stabilizer of the atom 1 contains
48/12=4 elements H:
(6.5)
48 CHAPTER 2
beneath which we have written the numbers of the atoms that are obtained
from the atom 1 by the action of a corresponding representative.
vVe write out the matrices of the representation T10 of the requisite group
OJ.° zero-block elements:
h1 h2 h17 h 1S h3
no no n
0 0 0 0 0
G DG 1
0
I
0
0
I
0
1 DO 1
0
h5 h7 h9 hlO h25
n
1 I 0 0 0
G DO DG DO
0
0
0
0
0
1
0
I DO I
0
MICROSCOPIC REALIZATION OF ORDER PARAMETERS 49
The calculation of the basis functions is carried out according to the for-
mulas (5.7)-(5.9), with allowance for the fact that the matrix R(h) should be
replaced by 15 h R(h). Using the formula (5.7), we find for '110 the number nV = 3;
thus we should obtain three sets of basis functions that transform according
to this representation. The quantity W~v calculated according to the formula
(5.9) should be a 9 X 9 matrix. We represent this matrix in block form (the
numbering of the blocks corresponds to the numbers of the IR basis functions):
A= ( 02 00 0)
0 ,
0
B= ( 0
0 0)
1 0 ,
0 0 0)
D= ( 0 0 1 . (6.7)
o 0 0 o 0 1 010
This matrix determines the atomic components of the basis functions on the
atom 1. For the atoms 2 and 3, for example, the quantity W~v calculated
according to the formula (5.9) has the following structure:
W2v = hgW~V
0
= ( A'
D'
0 ~')
B"
D" B"
DII) ,
o B' D' o o
(6.8)
where A', B', D' are obtained by premultiplying the matrices A, B, D into
15 hg R( h g ), and A", B", D" into 15h. R( h s ). The same applies to the other atoms.
The non-zero columns of the matrices Wfv give the atomic components of the
basis functions. Inasmuch as the representation '110 is involved in the magnetic
representation three times, we have three independent columns. We take the
non-zero columns, viz., the first, the fifth, and the sixth. We label their total-
ity for the ith atom by "pi; then, employing the expressions (6.7) and (6.8), we
50 CHAPTER 2
2 0 0 0 1 0 0 1 0
0 0 0 0 0 0 0 0 1
0 0 0 0 0 1 0 0 0
0 0 0 0 0 0 0 0 1
\[11= 0 1 0 \[12 = 2 0 0 \[13 = 0 1 0 (6.9)
0 0 1 0 0 0 0 0 0
0 0 0 0 0 1 0 0 0
0 0 1 0 0 0 0 0 0
0 1 0 0 1 0 2 0 0
Proceeding in similar fashion, we calculate '!f!i for the nine atoms of the orbit
that remain. The final result may be represented in the form of Table 2.8.
Table 2.8
Pseudovector basis functions of the representation TlO for a garnet
crystal with magnetic atoms in position 24( c) ('" = 0)
The atomic components of the basis functions are written out in explicit form
only on the first trio of atoms. On the next following trios, the atomic com-
ponents are obtained by multiplying the above components into +1 or -1, as
shown at the bottom of the table
MICROSCOPIC REALIZATION OF ORDER PARAMETERS 51
References
52
SYMMETRY CHANGE AT PHASE TRANSITIONS 53
to ensure that the number of right angles is at a maximum and the cell vol-
ume is minimal. Along with the Bravais cell, one also exploits the concept of
the primitive cell, which is constructed on the three shortest crystal transla-
tions t l , t 2, t3' As an example, Fig. 3.1 illustrates the relation between the
primitive cell and the Bravais lattice for f.c.c. crystals.
4Z
a,O-----~:J
Fig. 3.1. The primitive cell and Bravais cell of the f.c.c. lattice.
Using equations (2.1), one can construct for each Bravais cell a correspond-
ing reciprocal-space cell, called a Wigner-Seitz cell. Its edges are generated by
the reciprocal-lattice primitive translation vectors b l , b 2 , b 3 • However, it is
the concept of the Brillouin zone that has proved to be the most convenient
for physical applications. The procedure of constructing a Brillouin zone boils
down to the following steps. Connect the chosen reciprocal lattice site with
the nearest sites. Through the middles of these segments, pass planes that
are perpendicular to them. If these planes generate a convex polyhedron, we
obtain the Brillouin zone. Otherwise we connect the chosen reciprocal lattice
site with the next nearest sites and, again, draw through the middles of the
resulting segments perpendicular planes with respect to the latter.
It is evident that the final result, that is, the shape of the Brillouin zone,
depends ultimately on the ratios of the Bravais lattice parameters. Thus, for
example, the Brillouin zone shape for a rhombohedral lattice depends on the
ratio of the Bravais cell height c to the base edge length a. For the case ~ > ~
the Brillouin zone is sketched in Fig. 3.2,a, and for ~ < ~ in Fig. 3.2,b. For
other lattices, a similar description of the Brillouin zones is available in books
[1,2].
Using the symmetry of the reciprocal lattice, it is possible to classify all the
points that belong to a Brillouin zone. A specific role in solid state physics is
54 CHAPTER 3
Fig. 3.2. Brillouin zones for the rhombohedral lattice for ~ > ~ (a)
and ~ <~ (b); r(O 0 0), F(O 0 ~), T(~ ~ 0), L(~ ~ ~).
played by what is called the symmetric or Lifschitz points. These are Brillouin
zone points that have a specific high symmetry, that is, some set of symmetry
elements that leave a given point fixed or transform it into an equivalent one.
A feature peculiar to the Lifschitz points is that the representation of the
appropriate vector", in terms of the vectors hI, h 2 , h3 contains numerically
fixed coordinates, whereas for the non-Lifschitz points it contains some running
parameters. In addition to the symmetric points, it is customary to single out in
the Brillouin zone symmetric lines and planes. Figures 3.2 through 3.5 present
the symmetric points for the lattices r rh, r c, r~, and r{.
A complete list of Lifschitz points for all Brillouin zones is given in Ta-
ble 3.1. The first column of the table indicates the symbol of these points
according to [1]. The second column gives the symbol of the wave-vector star
corresponding to a particular point after Kovalev [3J. The third column spec-
ifies the coordinates of a point in the reference frame of the reciprocal lattice
Bravais cell (in units of 2:). The fourth column gives the notation of the star
SYMMETRY CHANGE AT PHASE TRANSITIONS 55
Fig. 3.3. Brillouin zone for the simple cubic lattice r c; reo 0 0), X (0 ~ 0),
MO ~ O),R(~ ~ ~).
Fig. 3.4. Brillouin zone for the f.c.c. lattice r{; reo 0 0), W(l 0 ~),
L(~ ~ ~), X(O lO).
arms in terms of the primitive translation vectors of the reciprocal lattice. The
last column furnishes the representative elements whose operation on the first
star arms permits the remaining arms to be obtained.
Now we are in a position to perform a straightforward analysis of the
changes in the translational symmetry of a crystal at a phase transition over
56 CHAPTER 3
Fig. 3.5. Brillouin zone for the b.c.c. lattice r~; reo 0 0), H(O 1 0),
P(~ ~ ~), N(~ ~ 0).
Lifschitz stars. This problem was first solved by Lifschitz in his celebrated
paper [4]. In what follows we present an approach that enables one to obtain
the change of the lattices in the case of a magnetic phase transition as well [5].
where the summation over L is performed over all the arms Kl, K2, ... of the
star {K}. This equation relates the density in the nth crystal cell to the density
LL PL(O) in the zero cell from which the nth cell is separated by the translation
vector tn of the initial phase. The equation of interest may be derived from
the general density expansion (1.19), assuming the phase transition to go over
one star and allowing for the fact that the IR basis functions for the star arms
contain the factor exp(iKLt n ).
The quantity PL(O) should be called the arm contribution to the density
function. This quantity contains information about the density distribution in
the zero cell, but the details of this distribution as well as the symmetry of
SYMMETRY CHANGE AT PHASE TRANSITIONS 57
Table 3.l.
List of Lifschitz points for all Brillouin Zones
Designation: (1) symmetric point, (2) star number, (3) coordinates of
arm ~l at reference point Bi, (4) star arms, (5) representatives of k 4 .
1 2 3 4 5
Lattice f tr ; = B l , b 2 = B 2, b 3 = B3
bl
222 ki = !(bi + b 2 + b 3) hl
1 1 I
R 1
2 22 ki = !(b2 + b 3)
all hl
U 3 2 12 ki = !(bi + b 3)
10 hl
V 4 110
22 kl = !(bl + b 2) hI
X 5 100
2 kl = !b l hl
6 0102 kl = !b2 hl
Z 7 00 12 kl = !b3 hI
Lattice f m ; b l = B I , b 2 = B 2, b 3 = B3
D 8 2 2 kl = !(bi + b 3)
101 hI
C 9 22 ki = !(b2 + b 3 )
all hI
222 kl = !(b 1 + b 2 + b 3) hI
I I I
E 10
Z 11 00 12 kl = !b3 hI
B 12 100
2 kl = ~bl hl
Y 13 010
2 ki = ~b2 hl
A 14 110
22 ki = ~(bl + b 2) hI
58 CHAPTER 3
1 2 3 4 5
Lattice rb .
m' bl = HBl - B 3), b 2 = B 2, b 3 = HBl + B3)
V 4 2 2 kl = ~b2' k2 = -~b3
101 hlh4
A 7 0 120 kl = ~bl hl
Y 8 100 kl = Hb2 + b 3) hl
M 9 1102 kl = Hb 1 + b 2 + b 3) hl
Lattice ro; b l = B l , b 2 = B 2, b 3 = B3
X 20 100
2 kl = ~bl hl
Y 21 0 120 kl = ~b2 hl
Z 22 00 12 kl = ~b3 hl
T 23 22 kl = ~(b2 + b 3)
0 11 hl
U 24 2 2 kl = ~(bl + b 3)
101 hl
S 25 110
22 kl = ~(bl + b 2) hl
222 kl = ~(bl + b 2 + b 3)
1 1 1
R 26 hl
Lattice rb.
0, b l = ~(Bl + B2), b 2 = ~(Bl - B2), b 3 = B3
S 12 110
22 kl = ~bl' k2 = -~b2 hlh2
Y 15 010 kl = ~(bl + b 2) hl
z 16 00 12 kl = ~b3 hl
T 17 0112 kl = ~(bl + b 2 + b 3) hl
SYMMETRY CHANGE AT PHASE TRANSITIONS 59
1 2 3 4 5
Lattice rl; b 1 = ~(-B1 +B2 + B3), b 2 = ~(B1-B2+B3)' b 3 = HB1 +B 2-B3)
1 11
L 10 222 kl = Hbl +b2+b3),k2 = -~b3,k3 = -tb2,k4 = -~b1 h1h2h3 h4
-222 kl = ~b1,k2 = -~(b1+b2+b3),k3 = ~b3,k4 = ~b2
1 1 1
L 11 h1h2h3h4
2-22 kl = ~b2,k2 = ~b3,k3 = -Hbl +b2 + b 3),k4 = ~b1 hlh2 h3h4
1 1 1
L 12
1 1 1
L 13 22-2 kl = ~b3,k2 = ~b2,k3 = ~b1,k4 = -~(b1+b2+b3) h1h2h3h4
T 15 100
2 k1 = ~(b2+b3) h1
Y 16 0120 kl = ~(bl +b3) hl
Z 17 00 12 kl = ~(b1 +b2) h1
Lattice ra; b 1 = HB2+B3), b 2 = ~(B1 +B3), b 3 = ~(B1 +B 2)
10 22 k1 = Hb2+b3), k2 = ~(b1-b3)
111 hlh4
11 2 2 kl = ~(bl +b3), k2 = ~(b1-b2)
111 h1h4
12 H1
22 kl = ~(bl +b2), k2 = ~(b3-b1) hlh4
S 13 22 kl = ~bl' k2 = ~(b2-b3)
011 h1h4
R 14 2 2 k1 = ~b2' k2 = ~(b1-b3)
101 hIh4
T 15 110
22 kl = ~b3' k2 = ~(b2-bl) h1h2
Lattice r q; b l = B 1, b 2 = B 2, b 3 = B3
X 15 0120 kl = ~b2' k2 = -~b1 h1h13
M 16 011
22 k1 = ~(b2+b3), k2 = -~(b1 +b3) h l h l3
R 18 110
22 ki = ~(bl +b2) h1
S 19 00 12 kl = ~b3 h1
A 20 1 1 1
222 k1 = ~(b1 +b 2 +b3) h1
60 CHAPTER 3
1 2 3 4 5
22 kl = ~b3' k2 = Hb 1 -b2)
X 13 110 hlh14
Lattice rc; b 1 = B 1 , b 2 = B 2, b 3 = B3
L 9 2221 1 1
kl = ~(bdb2+b3),k2 = ~bl,k3 = ~b2,k4 = !b3 hlh26 h 27h28
X 10 001 kl = ~(bl +b2), k2 = ~(bl +b3), k3 = ~(b2+b3) hl h 5h 9
k5 = !(b3-bI), k6 = Hb3-b2) h6 h U
p 10 1 1 1 kl = :t(b 1 +b2+b3), k2 = -kl hlh25
222
H 12 001 kl = ~(bl +b 2-b3) hI
SYMMETRY CHANGE AT PHASE TRANSITIONS 61
1 2 3 4 5
Lattice r rh; b 1 = Bb b 2 = B 2, b 3 = B3
T 5 EO
22 kl = Hb 1 +b2), k2 = ~(bl +b3), k3 = Hb2+b3) hlh3 h5
1 1 1
L 8 222 kl = ~(bl +b2+b3) hl
Lattice rh; b 1 = B 1 , b 2 = B 2, b 3 = B3
M 12 loa
2 kl = ~bl' k2 = ~b2' k3 =-~(bl+b2) hlh3 h5
J( 13 EO
33 kl = Hb 1 +b 2), k2 = -kl h 1h 13
L 14 lOl
2 2 kl = Hb 1 +b3), k2 = Hb2+b3), k3 = -Hb 1 +b2-b3) hlh3 h5
1 1 1
H 15 332 kl = ~(bl +b2)+~b3, k2 = -kl h 1h 13
A 17 OOl2 kl = ~b3 hl
(7.2)
which should be written only for the arms "'L involved in the transition channel.
The solutions of these equations may, broadly speaking, be different for the
various versions of arm mixing, and this points to the necessity of introducing
the concept of the transition channel.
62 CHAPTER 3
The shortest vectors of the direct and reciprocal crystal lattices in the rectan-
gular coordinate system have the form
2~ 2~ 2~
b1 = -(0
a
1 1), b2 = -(1
a
0 1), b3 = -(1
a
1 0), (7.5)
the being the primitive translation vectors of the Bravais cell (cube edges).
ai
To start with, we consider the transition over the arm ~l (one-arm chan-
nel). The unique equation exp(i~ltD) = 1 is satisfied for
(7.6)
The first three vectors are the shortest translations of the lattice rg imbedded
into the parent rc lattice, as sketched in Fig. 3.6a. The transition over another
arm, for example ~2, gives the system of translations
(7.7)
SYMMETRY CHANGE AT PHASE TRANSITIONS 63
(b)
Fig. 3.6. r~ lattices arising from the r~ lattice in the one-arm channels
of the star {~g}.
Thus the same rg lattice arises which is inserted into the initial-phase lattice in
a different fashion (Fig. 3.6b). Each of the six arms of the star (7.3) evidently
describes the same lattice, which is oriented in one of the six ways relative to
the parent lattice.
Consider now the two-arm channels. In the channel (~1~4) the system of
equations
(7.8)
which generate the lattice r q. A lattice of the same type, but imbedded in
a different way, arises in the channels (~2~5) and (~3~6). However, in the
two-arm channel (~4~5) we obtain a lattice r rh which is determined by the
vectors
(7.10)
The same lattice obtains in all the remaining two-arm channels, etc.
The results of a similar investigation of all the possible Lifschitz stars for
all the fourteen initial Bravais lattices are summarized in Table 3.2 [6]. The first
column indicates the symbol of the lattice resulting from a phase transition over
a given channel. The second column gives three Bravais cell edges, expressed
64 CHAPTER 3
in terms of the parent-lattice Bravais cell edges al, a2 and a3' In the case of
centered lattices the semicolon is followed by a specification of the centering
translations. The third column gives the change in primitive cell volume at
the transition. In the next two columns the corresponding transition channel
is indicated: The fourth column gives the number of the star according to
Kovalev (Table 3.1) and the total number of star arms (in parentheses), while
the fifth column specifies the set of arms participating in the transition. The
following notation is adopted here: (i) denotes that any star arm participates
in the transition, (ij) stands for any pair of arms taking part in the transition,
etc.; the parenthesized numbers (12), (34), (123), etc., indicate exactly which
arms participate in the transition.
This table lists all the possible changes of the lattices at a transition over
Lifschitz stars. Such results were first obtained by Lifschitz [4] and later repro-
duced in a book [7]; however, the above writings did not specify the transition
channels.
Table 3.2.
Lattices liable to arise as a result of a translation over Lifschitz stars
for 14 Bravais lattices
Transition Channel
f fD Description of the Bravais cell
of the new lattice n star star arms
1 2 3 4 5 6
fm al,a2,a3 2 8
1 2 3 4 5 6
.
rbm 2al, 2az, 2"1 ( al + a3); al + az 4 10-13(4) (ij)
rf0 2al,2aZ, 2a3;aZ + a3,a3 + al,al + az 8 10-13(4) (ij k), (1234)
rg ro al,az,as 2 18
rbm as,al + aZ,al - az; Hal + az + a3) 2 10(2)-15(2) (i)
C1
rb0 2al, 2az, a3; al + az 4 10(2)-15(2) (12) ::I:i
;..-
'i:i
rf0 2al,2az,2a3;aZ + a3,a3 + al,al + az 4 16(2) (i), (12) ~
tr1
2 19 ::::0
rq rq ab a Z, 2a3 ~
rq al +az,al -az,as 2 18
rvq al + az, al - az, 2a3; al + a3 2 20
1 2 3 4 5 6
rvc 2al,2az,2as;al +az +as 4 11(3) (ij), (123)
rc 2al, 2az, 2a3 8 10(3) (123)
rvc rc al,aZ,a3 2 12
o
r0b al, az + as, az - a3; Hal + az + a3) 2 9(6) (i) :::r:1
>
"'0
rq al + aZ,al - aZ,a3 4 9(6) (14), (25), (36) I-j
trl
i=C
rrh al - ~(-al + az + a3),aZ - Hal - az + a3),a3 - Hal + az - a3) 4 9(6) (ij) except (14), (25), (36), C,..)
1 2 3 4 5 6
ff0 2(at +az), 2(at-az), 2a3; 2at,at +aZ+a3,at-a Z+a3 4 14(3) (ij)
=
>
en
M
~
fh at + 2aZ,al - az, 2as 6 15(2) (i), (12)
~
fh 2at, 2az, 2a3 8 14(3) (123) Z
--_._.-
en
~
(3
Z
en
-'I
......
72 CHAPTER 3
MAGNETIC LATTICES
(7.11)
for determining all the vectors tD and tD for a given transition channel.
Consider, as an example, the magnetic phase transition from the f.c.c.
lattice over the three-arm star {,dO} with the arms (Table 3.1):
(7.12)
(7.14)
(7.15)
SYMMETRY CHANGE AT PHASE TRANSITIONS 73
/!1
rmJm
jrJ(j-_____--<J
Pc 4/rnrnm PmJrn
Fig. 3.7. Magnetic lattices arising from the r~ lattice by one, two and
three arms of the star {,dO}. Dark and clear circlets represent sites
connected by antitranslations; the asterisks mark parent-lattice sites
that are not related to black-and-white sites by either translations or
antitranslations. These parent-lattice sites will sometimes be called
'lost' translations.
(7.16)
G1 C G. (8.1)
2) The group 01 is an admissible group of the new phase (at the transition
over the representation DII of the initial-phase group G) if the restriction of
this representation to the aforementioned group contains the identity represen-
tation.
This criterion, due to Birman [9], follows from the fact that the density
function p of the new phase should be invariant under the subgroup G 1 , and an
invariant is a quantity that transforms according to the identity representation.
Therefore, being reducible in the general case, the restriction of the group G
IR to the subgroup G 1 should contain the identity representation. Otherwise
the group G 1 cannot arise by the irreducible representation DII.
3) If the suspected groups of the new phase turn out to involve two sub-
groups, G1 and Gi, so that one of them is a subgroup of the other (for example,
Gi C Gd and, in addition, both are admissible according to the Birman crite-
rion and are obtained at a transition over the same representation, then only
the maximal group (G 1 in this case) should be left in the final list of subgroups.
This maximal-subgroup principle follows from the natural requirement that the
description of the crystal symmetry be complete: It is necessary to enumerate
in the group all the symmetry elements that are available in the crystal after
the phase transition.
4) The representation over which the phase transition goes should be part
of the corresponding tensor representation (permutational, mechanical, mag-
netic, etc.) realized on the parent crystal. This reflects the real structure and
physical content of the order parameter.
SYMMETRY CHANGE AT PHASE TRANSITIONS 75
p= I:CA<PA' (8.2)
A
T(g)p=p. (8.3)
76 CHAPTER 3
Using equations (8.2) and (2.9), we rearrange the left-hand side of equation
(8.3) to read
CX = L Di1,v(g)CX"
A'
(8.4)
Table 3.3
Selection of the group of the new phase according to the Birman
criterion
0 3h 0 h7 Oh Ag Au Bg Bu Eg Eu F 2g F 2u F lg Flu
D§h
D24
2h
DZ'x
2h + + +
DZ'x y
20
D 2h D28
2h 2h + + +
Dl
2
D7
2
DZ'x
2 ++ + + + +
D6
2
D9
2
DZ'XY
2 ++ + + + +
Ortho-
C 13
2v
C 20
2v
Cz,x y
2v + + + + + +
rhombic C1 6 Cxy,z
2v
22
C 2v 2v + + + + +
Civ 19
C2v cz,x
2v + + + + +
Ch C~h qh + + + + +
qh Cih
xy
C2h + + + + +
Ci C~ C2 ++ + + + + + + + +
Mono- C~ Ci C2xy ++ + + + + + +
clinic
Csl Cs4 C sZ + + + + + + +
CS4 C83 cxy
8 -I- + + + + + + +
Tri-
C~, C~
• Ci + + + + +
clinic
ct C1l Cl ++ + + + + + + + +
80 CHAPTER 3
The left-hand half of Table 3.3 gives all the cell-preserving subgroups for
the space groups O~ and oX of the structures A-15 and C-15. The number of
these subgroups is the same in both groups and is equal to thirty-two. This is so
because the point group Oh has thirty-two subgroups with crystallographic ally
non-equivalent orientations. Each of the subgroups will have a space subgroup
of its own. Thus the lists of the thirty-two space subgroups for the structures
A-15 and C-15 differ only in the upper numbers of these subgroups. As the
general notation of these subgroups straightforwardly for all the structures, we
may take, for the time being, the notations of the corresponding point class
from column 4. Similar point subgroups with different orientation in the class
Oh differ in Table 3.3 in the orientation indices of their generators: In the
International Notation
zx - z xy - zx
D 2d = 4z 2x m xy , D 2d = 4z2xymx, D 2" = mzmxmy ,
D;"x y = mzmxymxy, c~~xy = 2zmxymxy, ....
The subgroups are grouped into blocks according to crystal systems. The
right-hand half of Table 3.3 enumerates the admissible subgroups according
to Birman for all the IR's of class Oh; these subgroups are marked with a +
symbol.
Using this table, it would be easy to write a list of admissible (maximal)
subgroups obtained from each IR. To this end, one needs to take only those
IR's that are involved in the mechanical representation. For the structure A-15
(the structural formula of the compound is A3B) this representation has the
following constitution (calculated according to the formula (5.1)):
(8.7)
and for the structure C-15 (the structural formula of the compound is A2 B)
(8.8)
As a further step, we need to determine the type of mixing for each admis-
sible IR that leads to the symmetry of the phase described by a given group G1 .
We wish to illustrate this stage of work by a particular example. We choose
the three-dimensional representation TS for the structure C-15. In Table 3.3
there are in this case twelve subgroups G 1 that are admissible according to Bir-
man. We take the first of them, Dn. The density function of the new phase is
P = Po + Dp, where
(8.9)
SYMMETRY CHANGE AT PHASE TRANSITIONS 81
should be invariant under the group G 1 = Dn. We separate from the group
Oh IR 78 matrices the restriction of the representation to the subgroup D~dxy.
Requiring that 6p be invariant under D~dxy, it suffices to take only the gener-
ators of this group: h38(4zI) and h16(2xy). Acting with the representation 78
matrices of these elements on the vector column of the mixing coefficients, we
obtain a system of equations,
Table 3.4
Groups of the new phase and mode mixing types for all IR's in A-15
and C-15 compounds
C C
C, (2 ± v'3)C 00 C D Z,xv D Z,xv C
D4h D4 2h 2d 4h C4v
ThTd CCO - C~:'z C;K C~:'Z
C C
- - C 1C 20 - C•Z - C 8z
The wavy line indicates the groups arising from the passive representations in
which the phase transition is of first order. It is this situation that occurs in the
N aCl-structure compounds V 3Si and Nb 3Si where the tetragonal-phase space
group corresponding to the two-dimensional IR 75 of the group O~ turns out
to be D~h.
9. Domains
Most of the phase transitions come about when the temperature is varied; as a
rule, the crystal symmetry reduces as the temperature is lowered. In keeping
with the Curie principle (which states that the dissymmetry which has arisen in
the system should be present also in the causes that generate it), the symmetry
of the crystal should not change when the temperature is varied, for the effect
of temperature on the crystal has the symmetry of a scalar. The paradox noted
is resolved as follows. At a phase transition crystals break up into domains.
The symmetry of each domain is lower than the parent-crystal symmetry, but
SYMMETRY CHANGE AT PHASE TRANSITIONS 83
The set of vectors 11" obtained by the action of the representatives of the de-
compost ion (9.1) on 110/ forms an orbit of 110/ relative to the group G. Since
there corresponds to each equivalent vector 11" some domain of the dissymmet-
ric phase, the number of different domains n is equal to the number of cosets
in the decomposition (9.1):
_ II Gil (9.3)
n-IIGDII'
IIGII and IIGDII being the orders of the groups G and GD respectively.
The stabilizers of different vectors 11" are evidently conjugate groups:
It may happen that some of the representatives g/1 do not change the group GD.
Such a set of elements, including the elements of the group G D itself, forms a
group and is said to be the normalizer NG D of the group GD in the group G.
In this case the domains corresponding to the elements of the normalizer NG D
have a common symmetry group GD. The number of such domains d is equal
to the number of cosets in the decomposition of the normalizer N G D relative
to the subgroup GD:
n= md, (9.6)
subgroup of the abstract point I-group, that is, a subgroup of the "image" of
the responsible IR. A classification of domains may be constructed also within
the framework of the relation between the I and ID groups that act in the OP
space. The number of domains will then be equal to the number of cosets in
the decomposition of the I group relative to the subgroup ID or equal to the
ratio of the orders of these groups
(9.7)
Both approaches, based on interrelations of the space groups G and G D
and the abstract groups I and ID, are completely equivalent; so are the corre-
sponding methods of constructing invariant expansions of the thermodynamic
potential.
The introduction of the concept of the transition channel (§7) has made natural
the introduction of a more detailed physical classification of domains. The
expression (9.3) for the number of domains, equal to the index of the subgroup
G D in the group G, may be rearranged to read
(9.8)
GO, G~ being the corresponding point groups of the phases and T, TD their
translation groups. The numbers no and t respectively indicate the factor by
which the number of rotational and translational elements decreases.
We wish to see what constitutes the first cofactor 1]0 in the expression
(9.8). Let the phase transition be accompanied by an increase in unit cell
volume, that is, let the transition go over a non-zero wave-vector star {"'}. In
an overwhelming majority of the phase transitions observed, the transition goes
over a one-arm channel of the star. The symmetry group of the dissymmetric
phase at this transition either coincides with the wave-vector group, GD = G",
or a subgroup of it, GD c G". In the first case the number of domains differing
in orientation coincides with the number of arms (no = I,,). We call such
domains arm domains. In the second case an additional domain structure may
arise within each arm domain with a fixed arm "'L. Such domains are said to
be rotational. The number of rotational domains is easy to determine from the
equation
IIG~II
m = IIG~II" (9.9)
86 CHAPTER 3
no = mlk. (9.10)
Here gr stands for representatives of rot.ational domains in the zero block, and
9 s = {h1Its} for the 'lost' translations of the group G". If we now decompose
the group G with respect to the subgroup G",
(9.13)
(9.14)
(9.15)
Finally, the relation of the atomic vectors in two arm domains (the Lth and
the first) is given by
(9.16)
1. Find the number and type of domains that are liable to arise at a struc-
tural phase transition O~ -+ D~h in A-15 compounds (§5). This transition
occurs without changes in translational symmetry ('" = 0) and therefore no
arm domains arise.
We choose as the symmetry group of the first domain the group D~h(Z),
where Z denotes the orientation of a four-fold axis. Then the coset decom-
position of the initial-phase symmetry group O~ relative to D~h(z) has the
form
(9.17)
the operation of the elements g5 and g9 on the first domain. These domains
88 CHAPTER 3
are depicted in Figs. 3.8a and 3.8b (the calculations have been carried out
according to formula (9.14)).
2. Find the number and type of domains in C-15 compounds undergoing
=
a structural phase transition from Ok with K- O. In the compounds HN 2 and
ZrV 2 the symmetry of the low-symmetry phase has not yet been determined
unambiguously. A knowledge of the domain structure may prove to be essential
for a conclusive solution of this issue. According to the experimental data
available, the dissymetric phase has either the symmetry D5~ or D~d [10]. As a
result of the Ok -+ D~d transition four types of domains should arise, as there
are four representative elements of the decomposition of the group Ok relative
to D~d : gl, g6, g7, gs· As can be seen from Table 3.4, the group D~d may be
obtained as a result of a transition over the three-dimensional representation
7S with the mixing coefficients (C C C). To obtain the structure of the rth
domain, we act on the structure of the first domain with the elements g6, g7,
and gs. The results of calculations according to the formula (9.14) are cited in
Table 3.5 and sketched symbolically in Fig. 3.9 (to make the figure simpler, all
the cell atoms are reduced to a point).
For the structural phase transition Ok -+ D5~ we may choose gl, g2, g5,
g6, gg, gll as the representative elements. The atomic displacements in the first
domain are described, according to §5, by the representation 7S(0 0 C). The
other five domains are obtained by the action of the representative elements on
the first domain. The results of this operation are consolidated in Table 3.6.
3. Consider now the domain structure that arises at a transition with
K- =F O. As has been shown in §4, the ordering of hydrogen atoms in the
Ta-H(D) and Nb-H(D) systems is described by one star {K-9} arm of the parent
disordered phase group O~. The symmetry of the ordered phase (see Fig. 2.4)
SYMMETRY CHANGE AT PHASE TRANSITIONS 89
{ 5
is described by the group G D = D~g. In this example we thus deal with the
case G~ = G~ = D2h and G D C G,., that is, only arm and antiphase domains
should arise. As the elements of the decomposition of the group O~ relative
=
to G" D~~ we may take the representative elements of the arms li,L =
gLIi,l
which are given in Table 3.1, namely, gl, g2, g5, g6, g9, gu. The group
G" = D~~ may in turn be decomposed into two cosets relative to the subgroup
G D = D~~ with Ii, f:. 0:
(9.18)
where gs = {h1Its}, and one of the lost translations of the initial phase is
ts = a3, a1 + a3, a2 + a3 (Fig. 2.4). As a result, twelve domains should arise at
the transition (Fig. 3.10); each of the domains may be obtained from the first
90 CHAPTER 3
domain by the action of the elements gl, gIg., g2, g2g8, g5, g5g., g6, g6g.,
g9, g9g$, gll,gllg.·
Table 3.5
Domain structure in C-15 compounds at the phase transition
oX -+ D~d with", = O.
Domains Atoms A Atoms B
1 2 3 4 5 6
Table 3.6
Domain structure in C-15 compounds at the phase transition
oX -+ D5~ with", = o.
Domains Atoms A Atoms B
1 2 3 4 5 6
Btl BB
.:.
[J [J
4 ....
[J a
I [J
L[Jgggs
z ==
t [J
II It
Igg
-0
z f /2 -6
Z == J/4 -x
z 1 -0
z = J/2 - ....
Fig. 3.10. Domain structure in the interstitial phases ofTaH and NbH
at the phase transition O~ - D~~ over the star {K9}. Projections of
the unit cell of ordered phases on the z = a plane.
OZ . Az-1.xz-3·Oz-1· A Z _3
, -_ l4'~' - 2' , - 4' , - ,~, - 2
ceeds a low-symmetry phase at some value of Te. Hence there naturally follows
the generally accepted initial-phase selection principle according to which one
should choose as the initial phase the high-temperature phase, the most sym-
metric of the phases really observed. Thus it is normally the paramagnetic
phase of the crystal that serves as the initial phase for magnetic phase transi-
tions.
In a number of cases, however, such an approach to the selection of the
initial phase does not allow the entire set of observed properties to be described
correctly. Thus in some crystals a domain structure is liable to occur in the
high-symmetry phase G existing up to the melting temperature. Such a do-
main structure is characterized by the following properties. First, the various
domains may be brought into coincidence with each other by crystallographic
symmetry elements that do not belong to the group G observed. Second, by
applying external fields, a displacement of domain boundaries and a change-
over of domains can be observed. The presence of a domain structure with
these properties below the melting point may be viewed as the result of a
hypothetical phase transition Go --+ G, with Go a hypothetical initial phase.
For example, the occurrence in an orthorhombic crystal of domains that are
transformed into each other by a hexad axis C6 is indicative of a hexagonal
hypothetical initial phase.
The concept of a hypothetical initial phase may prove constructive also
m another situation. In some crystals undergoing a G --+ G 1 transition it
is not possible to correctly describe the entire set of observed characteristics
(susceptibility, the temperature dependence of the order parameter, the type
of the phase diagram, etc.) proceeding from the type of the thermodynamic
potential <1> constructed for the relevant representation DV of the group G.
Assuming that a hypothetical Go --+ G --+ G 1 transition may exist, another
potential <1>0, invariant for the group Go, may be constructed that provides an
adequate description of the phenomena observed.
Finally, crystals exist that exhibit G 1 --+ G 2 phase transitions where
the groups G 1 and G2 are not symmetrically related, that is, G 1 (j. G 2 and
G 2 (j. G 1 . In these crystals the atomic displacements u at such transitions
satisfy the condition u ~ a, where a is the lattice spacing. Such transitions
may, nevertheless, be described using Landau's phase transition theory if one
introduces a hypothetical initial phase Go such that Gl C G2 and G2 C Go
and if one considers the sequence of transitions Go --+ G 1 --+ G 2 •
In all the cases where specific characteristics of a crystal undergoing a phase
SYMMETRY CHANGE AT PHASE TRANSITIONS 93
(10.1)
where u is a small displacement (lui ~ a), then the element g should be part of
the paraphase Go. One should also take account of the fact that the Go -+ G
transition may be accompanied by an atomic ordering. When determining the
paraphase structure one may therefore try to combine, in the sense of equation
(10.1), the positions occupied by atoms of different sorts.
In the general case geometric criteria of the search for the paraphase pro-
vide no unambiguous answer. Further selection from among the remaining
candidates for the role of the paraphase is done on the basis of physical crite-
ria; that is, the domain structure observed in the phase G should correspond to
the symmetry change Go -+ G, Go being the symmetry group of the paraphase.
In the case of an array of transitions Go -+ G 1 -+ G 2 , where G 1 -+ G 2 is the
observed transition and Go the paraphase, the anomalies in the physical prop-
erties at the transition G 1 -+ G 2 should be correctly described by a potential
<1>0 invariant under the paraphase group Go. The phase diagram obtained from
an analysis of the potential <1>0 should coincide with that observed. A detailed
analysis of physical criteria is available in [24] and [25].
transition D~~ ---+ Civ occurs in this compound. One of the features peculiar
to this compound is the occurrence of a complicated domain structure in some
samples at room temperature, that is, in the phase D~~ [26]. It is remarkable
that the angle between the edges of unit cells belonging to different domains
is close to 60 0 • Another feature of the high-symmetry phase in ammonium
sulfate is that by applying small pressures to the sample it is possible to cause
the domain boundaries to move and to bring the entire crystal into a single-
domain state [27].
The above peculiarities of (NH4hS04 have been accounted for on the
assumption that a hypothetical high-symmetry paraphase is present in this
compound [28].
In searching for the paraphase, we base ourselves on the experimentally
determined structure of the phase D~~. According to structure analysis data
[29], the S atoms occupy the position 4(c) of the group D~~ with the parameters
x = 0.250, Y = 0.417, and z = 0.250. The oxygen atoms occupy the positions
4( c) and 8( d). The nitrogen atoms occupy the positions 4( c), the hydrogen
atoms the positions 4(c) and 8(d).
We choose the possible paraphase symmetry groups. These should be
super groups of the group D~~. The presence of sixty-degree domain walls
indicates that we may restrict ourselves to hexagonal supergroups. From space-
group subgroup tables, we find that we may choose as the paraphase symmetry
groups the groups D~~ and D~h which have as their subgroup the group D~~
in channel 2. This channel describes the possible imbedding of the cell fo of
the phase D~~ into the cell fh of the hypothetical paraphase D~h (Fig. 3.11).
In order to choose from the groups D~h the one that is the group of
the paraphase in this compound, we need to determine into which paraphase
multiple-point positions the positions occupied by the atoms in the phase D~~
should pass. Since the sulfur atoms in the phase D~~ occupy one position
4(c), it is with this position that we start our analysis of the transformation
of coordinates at the transition D~~ -> D:;.4. The simplest of the methods is
the geometric one. Using experimental data, we sketch the arrangement of S
atoms in the cell fo as centers of triangles in Fig. 3.12. On comparing the
images of the multiple-point positions of the groups D~h and D~h' we find that
only the position 2( c) of the group D~h' that is, the centers of triangles in Fig.
3.13, coincides with the location of the S atoms. It is the group D~h that was
chosen as the paraphase in [27].
Fig. 3.12. Arrangement of S04 in phase D~~ after [29]. The S atoms
are located within the triangular pyramids. The numerals denote
height . • , oxygen atoms 0(3) 8(d); 0, atoms 0(2) 4(c); the atoms
0(1) 4(c) lie at the vertices of the pyramids. The shaded circlets
represent pyramids with the vertices up.
triangles portrayed in Fig. 3.13. The dashed line in Fig. 3.12 singles out the
fo cell which coincides with that pictured in Fig. 3.13.
the 0(1) atoms being in the 4(f) position is ~. The paraphase thus constructed
enables one to describe the entire set of experimental data on the physical
properties of the phase D~~. Thus the assumption as to the hypothetical phase
transition D~h -> D~~ in (NH 4hS04 accounts for the observed peculiarities
and the occurrence of a domain structure in the phase D~~ that really exists
up to the melting point. Analysis of the paraphase structure has permitted the
important conclusion that the D~h -> D~~ transition should be accompanied by
displacement of the oxygen atoms relative to the sulfur atoms and by ordering
of the orientations of the tetrahedra, as is shown in Fig. 3.12. This means that
the transition D~h -> D~% should be described by at least two interacting order
parameters [28].
The idea of the paraphase has proved constructive in analyzing optical
spectra of complicated compounds. Thus the introduction of a hexagonal para-
phase has made it possible to decode low-frequency combination scattering
spectra of monoclinic sodium trihydroselenite [30].
References
gPn (. . . 1],). ••• ) == Pn(. .. g1],). ... ) = Pn(. . . 1],). ••• ), (11.2)
g1],). = I: DX,).,(g)1],).'·
,).,
(11.3)
100
ANALYSIS OF THE THERMODYNAMIC POTENTIAL 101
the generators of the group are used as the elements g. It is only the group
generators that one should take to derive the equations (11.2) which determine
the coefficients of a polynomial. Depending on the ratio between the numbers
n (degree of polynomial), Iv (dimensionality of the IR), and r (number of
generators in the group), part of the initial literal coefficients may turn out
to be indeterminate. The number of these coefficients gives the number of
independent invariant polynomials of given degree Pn.
When desired, this number may be calculated in advance since it coincides
with the multiplicity factor of the group G identity representation involved
in the symmetrized product of the representation [D]n(g) according to which
the product of the quantities constituted by the set {1].>.} should transform.
According to the general formula (2.14),
Let it be noted that there is only one second-order invariant for the IR DV
and that invariants of degree three exist only for passive representations.
The scheme described above may be illustrated in detail by particular
examples, in which we restrict ourselves to degrees not higher than four.
V(h,) = G 1
o D(h14) = (0I 1 0)
0 ~ , V(h,,) = G
o
1 0
0) .
o 001 o 1
According to equation (11.3), we may write the operation of these elements on
the OP components:
h 5 1]1 = 1]2, h 14 1]1 = 1]2, h 25 1]1 = 1]1,
h 5 1]2 = 1]3, h 14 1]2 = -1]1, h 25 1]2 = 1]2, ( 11.5)
h 5 1]3 = 1]1, h 141]3 = -1]3, h 25 1]3 = 1]3·
102 CHAPTER 4
is invariant.
Now we construct third-degree polynomials of the most general form
On acting additionally with the element h14' we obtain the new equations:
VI = V4 = V5 = O. Thus only one term remains invariant with respect to the
elements h5 and h 14 :
P3 = 1JI1J21J3· (11.8)
Consider the fourth-degree polynomial
(11.10)
Summing the results (11.6), (11.8) and (11.10), we represent the invariant
expansion of the thermodynamic potential as
We consider the group O~, the two-dimensional representation 7"5 for K, = 0 (this
representation describes the structural phase transition in the superconducting
compounds Nb 3 Sn and V3 Si from the class A-15).
ANALYSIS OF THE THERMODYNAMIC POTENTIAL 103
The matrices of this IR are given in Table 2.5. Choosing the elements
and h25 as the generators of the group, we write, by means of the
h 5 , h14
corresponding matrices of the representation, the result of their action on the
two OP components, 111 and 112:
(11.12)
(11.13)
Equating this result to P3 yields two equations, (V3 = V3 and (2V3 = V3, which
are simultaneous if V3 = O. Thus the invariant third-degree polynomial has the
form
P3 = 11~ + 11~. (11.14)
the action of the element h14 gives Ui = U2 and U4 = U5. Acting with the
element h5 on the polynomial obtained, we find additionally Ui = 0 and U4 = o.
We obtain the only invariant fourth-degree polynomial
(11.15)
104 CHAPTER 4
U= V2
1 (1 1 (11.17)
(11.19)
GENERAL REMARKS
(12.1)
(12.2)
where G n is a normal subgroup of G n +1 . The factor groups GdG 2 , G 2 /G 3 ,
G 3 /G 4 , ... are called the factors of a normal series.
Now we wish to see how the IRBI for the group 1 itself can be constructed
if one knows the IRBI relative to a normal subgroup H' of the group 1 (H' ~I).
Assume that we are given some representation D = { ... D( a) ... } of the group
1 of finite order N(I). The normal divisor of the group 1 is the subgroup
H' ~ 1 of order N(H'). In terms of h, 12 , 13 , ••• we denote the IRBI relative to
the group 1 for the given set of quantities C 1 , C 2 , C3 , ... Cc, which transform
according to the representation D = { ... D(a) .. .}, where a E 1, as follows:
Let :11, :12, :13, ... be an IRBI with respect to the variables 0 1, O2,03, .. . Oz
transforming according to the representation D(S) of the group H'.
Let us see how the quantities :11, :12, ... transform under the action of the
elements a E I. We partition the group I into cosets relative to the subgroup
H':
(12.4)
L L
(12.5)
If :lp is an invariant under the group H', the quantity 0 L:lp = :1//) will also
be invariant under H' and obey equation (12.5).
We take now any invariant Iq from the IRBI of the group I. The quantity
Iq is invariant under the action of any element of the group I, specifically under
the subgroup H'. Therefore Iq may be represented as a polynomial in the group
invariants:
(12.6)
Consequently, the polynomial P(lt ,12, ... ,IG) == P( {Iq}) may be represented
in the form
(12.7)
Consider the case when the order of the factor group 1/ H' is equal to two:
(12.9)
(12.10)
108 CHAPTER 4
The requirement that the polynomial P'( {..1p }, {..1;}) be invariant under the
group I, as follows from equation (12.9), resides in the equation
(12.11)
On denoting
..11 = U1, ..12 = V1,···, ..1H = Zl; (12.12)
..1{ = U2, ..1~ = V2,· .. , ..1iI = Z2,
the expression (12.11) may be rearranged to read
(12.13)
(12.15)
The multiplication table for the factor-group elements has the form
Using the expressions (12.15) and (12.16), we can write the operation of the
group I elements on the quantities {.1l, .12, ... ,.1H}:
(12.21)
Equations (12.20) imply that the polynomials are symmetric with respect to a
cyclic permutation of the arguments {Ul, Vl, ... }, {U2, V2, ... } and {U3, V3, ... },
so that Theorem 2 may be applied.
Theorem 2. The IRBI for polynomials in ths system of variables (Ul, Vl, Wl, ... ,
Zl, U2,V2,W2, ... ,Z2, U3,V3,W3, ... ,Z3) invariant under a cyclic permutation
of the indices 1,2,3 consists of the elements
and the expressions resulting from them by substituting Vi, Wi,"" zi for Ui:
the expressions resulting from them by substituting for Ui, Vi the various com-
binations of two different symbols chosen from (Uj, Vi, Wi, ... , Zi):
and the expressions resulting from them by substituting for Ui, Vi, Wi the various
combinations of three different symbols chosen from (Ui' Vi, Wi, ... , Zj).
Returning to the notation (12.21), we may write the IRBI for the group I
(12.15).
From the above examples, it becomes clear how the IRBI is constructed
from the quantities C).. transforming according to the initial-phase symmetry
group IR DV. The IRBI construction procedure boils down to the following
stages.
1. From the table of space group representations, we write out the matrices
of the representation DV.
2. From among the group G elements, we choose those whose matrices in
the representation D V coincide with the identity matrix. We thus obtain the
kernel of homomorphism H.
3. We construct the factor group PH = G / H. In practice, we should
simply choose from all the representation F V matrices all the distinct matrices
and label the latter by DV(a). Multiplying the matrices DV together, we then
make up a Cayley square (compile a multiplication table) which enables us to
determine the abstract group I isomorphic to the factor group GH.
4. The resulting group I is expanded in a normal series (12.2).
5. Starting from the last term of the series, we express the invariants
for the higher-order term through the lower-order terms until we arrive at the
group I of interest.
(12.22)
where G is the symmetry group of the initial phase and X(g) the character of
the IR explored.
From the total number of invariants of a given degree N p , we need to sub-
tract those that may be represented as polynomials in lower-degree invariants.
The number of these invariants is determined by simply making an exhaustive
search. Finally, it is necessary to exclude from the remaining invariants those
that are related to lower-degree invariants in a non-polynomial fashion, that
is, those due to a so-called syzygy [1). For example, if an invariant I of degree
112 CHAPTER 4
Finally, when searching for the MIRBI, we may immediately discard in-
variants whose degree is higher than the order of the given] group. We may
do so by virtue of Theorem 4.
Theorem 4 (E. Noether). The degree of IRBI elements for polynomials in-
variant under a finite transformation group does not exceed the order of that
group.
We wish to construct the IRBI for the group oXfrom which the phase transition
takes place according to the representation with K, = 0. We start by finding the
kernel of the homomorphism of the representation T7. The zero-block matrices
of the representation T7 are given in §11, where it is shown that the kernel of
the homomorphism H in our case coincides with the space group Gl. Indeed,
the matrices of the representation T7 for the elements {h110}, {h 25 Ir}, {h1It n },
where tn is any translation and T a fractional translation, coincide with the
identity matrix. The point group ], isomorphic to the factor group G H, is
specified by a set of twenty-four different representation T7 matrices and, as
can be readily seen, coincides with the point group T d • Let a; label a group
] = Td element assigned by the matrix D(g;) of the representation T7. For
example, the element corresponds to the matrix
At the outset we find the IRBI for the group G 3 = C2 • In doing so, we
shall take account of the fact that according to Noether's theorem the maximal
degree of the polynomial invariants involved in the IREI does not exceed the
order of the group N(G3 = C2 ) = 2.
We denote the variables transforming according to the representation 77
by "11, "12, "13· We let the operator a2 act on these variables: a2("I1,"I2,"I3) =
("11,-"12,-"13). We introduce the variables U1 = "12, U2 = -"11, V1 = "13, V2 =
-"13 and set "11 = h,3' According to Theorem 1, the IREI for the group G 3
may be written in the form
(13.3)
(13.4)
I _ 1-22
I Ll = 1112 + r122 + TJ3'2 2.1 - TJ1TJ2TJ3, 3.1 - TJ1TJ2 + TJ1TJ3
22
+ TJ2TJ3'
22
Finally, we construct the required IRBI for the group Td = T+a 13 T. Since
a13(TJ1,TJ2,TJ3) = (TJ2,TJ1,TJ3), we have, according to equation (13.5), a13h1 =
ILl, a13 I2.1 = 12.1> a13 I3.1 = 13.1, a13 I4.1 = -14.1, Denoting U1 = 14.1 and
U2 = a13I4.1 = -14.1, we obtain one more invariant of the group Td : 14 = IL
(along with the three invariants present: h = ILl, h = h1' and 13 = hI).
Being reducible, this invariant is not part of the IRBI. Indeed, assume that
the invariant 14 = IJ.1 is not independent, that is, is polynomially expressed in
terms of the invariants II, 12 and 13 , Allowing for the degree of all the four
invariants, we write this relation in general form with some coefficients to be
defined:
Using the explicit form of the invariants and equating the coefficients of like
degrees of the basis functions, we obtain A = 0, B = 0, C = -4, D = I, E =
-15,F = 2, G = 2, so that
(13.6)
Our goal is to construct the IRBI for the group O~ using the OP components
that transform according to the IR T5 with", = O.
ANALYSIS OF THE THERMODYNAMIC POTENTIAL 115
a1 (~ ~) , a2 (~ €20) ' a3 0 C
2 €~) ,
a4 C~ ~) , a5 (~ €2) (0
o ' a6 1 o1) .
(13.7)
a1 a2 a3 a4 a5 a6
a1 a1 a2 a3 a4 a5 a6
a2 a2 a3 a1 a5 a6 a4
as as a1 a2 a6 a4 as (13.8)
a4 a4 a6 a5 a1 a3 a2
as a5 a4 a6 a2 a1 as
a6 a6 a5 a4 as a2 al
(13.9)
The I group (13.7) is isomorphic to the point group Csv . Indeed, the group Csv
contains two third-order elements, H3 and Hs (rotations about the 3 axis), and
three second-order elements, H 19, H 21, H 23 (reflection planes passing through
the 3 axis). If we denote the elements of the group C3v by the symbols
we can readily see that the Cayley square for the group C3v coincides with that
for the group (13.8), which proves the isomorphism of both groups.
116 CHAPTER 4
(13.10)
(13.11)
(13.12)
We now act on these invariants with the representative element g6, allowing
for the fact that g6"11 = "12, g6"12 = "11:
(13.13)
ANALYSIS OF THE THERMODYNAMIC POTENTIAL 117
and it is these invariants that form the IRBI. On comparing this conclusion
with the results of §11 (11.14), we see that the fourth-degree invariant reduces
to If, that is, it is not involved in the MIRBI.
I = G 1 + a25G1, G 1 = G 2 + a13G2,
(13.15)
G2 = G3 + a5G3 + a9G3, G 3 = G4 + a3G4, G4 = {a1,a2}.
The matrices corresponding to the elements a1, a2, a3, a5, a9, a13, a25 have
C00) C 0 0) C' ~ ) , a, (~
the form
a1 o 1 0 ,a2 0 -1 o , a3 0 1
0
01,
1 0)
C 0) C ~J
o 0 1 0 0 -1 0 0 -1 1 o 0
a, G 1)
-1 0
(13.16)
o 0 ,a13 -1 0 o ,a25 0 -1
1 0 0 0 -1 0 0
Exploiting the theorems of §12, we find the form of the invariants that
generate the required IRBI:
I1 = Til2+ 2
'f/2 + 'f/3' 2 = 'f/1 'f/2 + 'f/1 'f/3 + 'f/2 'f/3' 3 - 'f/1 'f/2 'f/3·
212222221_222 (13.17)
118 CHAPTER 4
From the example considered above, we see that the procedure of con-
structing the IRBI in the case of n, # 0 (for the Brillouin-zone points) is virtu-
ally no different from the case n, = O.
The IRBI for the groups C3 and C3t1 were obtained in §13. For the group
C 3 the IRBI consists of three invariants:
2 -- 1/13 - 31/17'122, I 3
11 -- 1/12+ 1/22 , 1 = 1/23 - 2 .
31/11/2 (14.1)
(14.2)
It is not difficult to construct the IRBI for the groups C4 and C4t1 . The
subgroup C4t1 may be decomposed into cosets relative to the subgroup C4 :
C 4t1 = C4 + 91C4, where 91 is the reflection in the plane parallel to the tetrad
axis. The group C4 may be represented also as C4 = C2 + 92C2, where 92 is a
rotation by 90° about the tetrad axis. The group C2 consists of two elements
C 2 = {g3, 94}, where 93 is an identity element and 94 = 9~.
Straightforward verification readily shows this decomposition of the group
C4t1 to be an expansion in a normal series. Therefore, in keeping with the
recommendations of §12, we begin the construction of the IRBI with the group
C 2 • Regarding 1/1 and 1/2 as the components of a two-dimensional vector, we
find according to Theorem 1 of §12: :11 =
1/~, ':12 =
1/~, .:13 =
1/11/2. Allowing
for the action of the element 92 on the resulting invariants .:11,2,3, we find the
IRBI for the group C 4 : -
(14.3)
(14.4)
The IRBI for the other groups Cn and Cntl are found in a similar way.
= =
Passing over to trigonometric notation, 1/1 1/ cos <p and 1/2 1/ sin t/J, the IRBI
for the groups Cn and Cntl may respectively be written as [12]
(14.5)
(14.6)
11 -_ 1Jl2 +2 2 [_ 1_444
1J2 + 1J3, 2 - 1J11J21J3, 3 - 1Jl + 1J2 + 1J3,
(14.7)
I 4 = 1Jl4( 1J22 - 1J32) + 1J24( 1J32 - 1Jl2) + 1J34( 1Jl2 - 1J22)
In analysing phase transitions, it very often suffices to deal with the poten-
tial's broken polynomial series that allows for terms of degrees not higher
than four in the OP. Introducing the concepts of the OP space, of the vector
11 = (1Jl, 1J2, .. . 1JI.,) in this space, and of the I groups permits enumeration of all
122 CHAPTER 4
possible model potential types without making an exhaustive search for all the
space group IR's but proceeding from the symmetry of the OP space. Thus,
for example, all the types of potential in the 'fJ4 model for a four-component
OP were derived in [15].
As already stated in the beginning of this section, all the I groups that arise
in the analysis of phase transitions should be subgroups of the rotation group
O(lv) in lv-dimensional space and these subgroups should have irreducible vec-
tor representations. In keeping with [15], we call these subgroups irreducible
subgroups. For example, irreducible subgroups of the three-dimensional ro-
tation group are the groups of the cube (Th' 0, Oh), tetrahedron (T, Td ), and
icosahedron (Y, Y h ). A list of all irreducible subgroups of the group 0(4) is
given in [15]; however, no definite information is available in the literature for
rotation groups of dimension higher than four.
Consider a set of different combinations of OP components of the form
'fJIPl 'fJ2P2 Pl v ,w Ilere PI + P2 + ... + Pl = p. Th e numb er 0 f suc h comb'lila t'Ions
... 'fJlv v
is equal to C~+P_I' For example, for a two-component OP (Iv = 2) we may set
= =
up ct 5 different fourth-order (p 4) combinations: 'fJ[, 'fJi, 'fJr'fJ2, 'fJI'fJ~, 'fJr'fJ~·
Each of these combinations may be viewed as a unit vector of the Cfv+P-I-
dimensional vector space [V/~]. To any set of invariants specified by a group I
we may then assign its own subspace E£
in the space. We denote the dimension
of the subspace Et, that is, the number of invariants of degree p, by Cp(I).
It is evident that one and the same set of invariants of specified degree
may be involved in the IRBI for different I groups; this follows from the IRBI
construction procedure itself, which is based on a step-by-step derivation of
invariants for groups generating a normal series. We call the maximal I group
corresponding to a given set of invariants of degree p, that is, realized in the
subspace Et, a centralizer [15] and denote it by Ip.
The group Ip may be regarded as the stabilizer of the subspace Et. By
decomposing the group O(lv) into cosets relative to the stabilizer Ip , it is pos-
sible to find the decomposition representatives Up and, by operating with these
representatives on the subspace Et, to construct a reducible representation of
the centralizer Ip. The multiplicity factor of the identity representation in-
volved in this representation will then be equal to Cp(I), that is, the number
P
of invariants of degree that generate the subspace E£.
These arguments permit formulation of the major steps in the construction
of invariants that make up thermodynamic potentials. The first step is to find
the vector representation r v for the lv-dimensional rotation group O(lv). From
ANALYSIS OF THE THERMODYNAMIC POTENTIAL 123
among all the subgroups of the group 0(1,,), one needs to take as possible 1
groups only irreducible subgroups, that is, subgroups that have an irreducible
vector representation of the rotation group 0(1,,). At the second step it is
necessary to ascertain the constitution of the symmetrized powers of the vector
representation [r v]P' From the group 0(/,,) IR's involved in [r v]P, one has to
pick out those representations the restriction of which on a subgroup 1 contains
an identity representation. The multiplicity factor of the identity representation
gives the number of invariants Cp(I) of degree p for a given 1 group. From all
the 1 groups with the same Cp(I), it is required to choose the maximal group; it
is this group which will be the centralizer r p' Finally, in some way or other (for
example, by the straightforward method) the corresponding set of invariants
may be constructed for each 1p.
This program was implemented in [15] for a four-component OP since the
group 0(4) and its irreducible representations are related to the quaternion
group [18], of which everything is known. An analysis of the symmetrized
powers (p = 2,3,4) of the group 0(4) irreducible vector representation showed
that from all the subgroups it is possible to single out twenty-two centralizers
14 and five centralizers 13 . These sets of invariants exhaust all the cases that
may correspond to some space group four-dimensional IR. Unfortunately, one
cannot tell which of these IR's correspond to the model potentials obtained.
UNIVERSAL CLASSES
As has been established in [16], a total of 106 1 groups exist that are formed
by space group IR's with Lifschitz stars. This means that one and the same
thermodynamic potential should describe a large number of different phase
transitions which may differ from one another in initial symmetry groups and
IR's describing these transitions. Specifically, the transitions may differ in star
numbers and, therefore, in the way the unit cells increase in size.
Invariant characteristics of all such phase transitions are the group 1 and
the number of OP components. We shall say that all the phase transitions that
are described by one potential, that is, that have one 1 group, form a universal
class. The authors of [16] have enumerated all the groups describing phase
transitions of which the initial-phase symmetry may be described by one of the
188 space groups (for ferroelastics). From the tables [16], it is seen that most
of the aforementioned transitions belong to universal classes with the groups
1 = C 4 , C4v , Th, 0, Oh. Therefore it is of practical interest to know all the
IR's involved in each of the above universal classes. Such information enables
124 CHAPTER 4
one, from a knowledge of the relevant IR, to write for a given phase transition
the explicit form of the IRBI and the types of solutions to the equations of
state (mixing coefficient types) and then to calculate the low-symmetry-phase
symmetry groups corresponding to them.
Table 4.1 lists all the three-dimensional IR's leading to each of the five
three-dimensional I groups: T, Th, Td, 0, and Oh. The numbers of IR's in
Table 4.1 are given according to Kovalev and Zak.
Table 4.1
I groups of three-dimensional IR's for cubic and hexagonal space
groups
Lattice rc
X M R
Group
IR I IR I IR I
Tl Tl Zl Th Tl Zl T T4 Z4 Th
T2 Z2 Th T2 Z2 T
T3 Z3 Th T3 Z3 T
T4 Z4 Th T4 Z4 T
Tl Tl Zl Th Tl Zl T T7 Z4 Th
h
T2 Zs Th T2 Zs Th Tg Zg Th
T3 Z2 Th T3 Z2 T
T4 Z6 Th T4 Z6 Th
TS Z3 Th TS Z3 T
T6 Z7 Th T6 Z7 Th
T7 Z4 Th T7 Z4 T
Tg Zg Th Tg Zg Th
T2
h - - - - - - T7 Zg Th
Tg Z4 Th
ANALYSIS OF THE THERMODYNAMIC POTENTIAL 125
Lattice fc
X M R
Group
IR I IR I IR I
Tl
d 7"1 ZI Oh 7"1 ZI Td 7"4 Z4 Oh
7"2 Z2 Oh 7"2 Z2 0 7"s Zs Oh
7"3 Z3 Oh 7"3 Z4 Td
7"4 Z4 Oh 7"4 Z3 0
T4
d 7"1 Z4 0
7"2 Zs Td
- - - - - -
7"3 Z3 0
7"4 Z2 Td
7"4 Z2 Oh 7"4 Z4 0
0h
1 7"1 ZI Oh 7"1 ZI 0 1"7 Z4 Oh
7"2 Z6 Oh 7"2 Z6 Oh 7"s Z9 Oh
7"3 Z2 Oh 7"3 Z2 0 7"9 Zs Oh
7"4 Z7 Oh 7"4 Z7 Oh 7"10 Z10 Oh
7"s Z3 Oh 7"s Z3 0
7"6 Zs Oh 7"6 Zs Oh
7"7 Z4 Oh 7"7 Z4 0
7"s Z9 Oh 7"s Z9 Oh
126 CHAPTER 4
Lattice rc
X M R
Group
IR I IR I IR I
0 3h 7"1 Z4 0
7"2 Zg Oh
7"3 Z5 Td
7"4 ZlO Oh
- - - 7"5 Z2 Td - - -
7"6 Z7 Oh
7"7 Z3 0
7"8 Z8 Oh
Lattice rfc
X
Group
IR I
T2 7"1 Zl T
7"2 Z2 T
7"3 Z3 T
7"4 Z4 T
r 3
h 7"1 Zl T
7"2 Z5 Th
7"3 Z2 T
7"4 Z6 Th
7"5 Z3 T
7"6 Z7 Th
7"7 Z4 T
7"8 Z8 Th
ANALYSIS OF THE THERMODYNAMIC POTENTIAL 127
T2 Z2 0 T2 Z3 Td
T3 Z4 Td T3 Zg Oh
T4 Z3 0 T4 Z4 0
T d5 Tl Z3 0 T5 Z6 Oh
T2 Z5 Td T6 Zl 0
T3 Z4 0 T7 Z7 Oh
T4 Z2 Td TS Z2 Td
0 3 Tl Zl Td
T2 Z2 0
T3 Z3 0 rvc
Lattice
T4 Z4 Td
H
04 Tl Z5 Td Group
IR I
T2 Z4 0 T3 T4 Z4 Th
T3 Z3 0 T4 T4 Z4 Th
T4 Z2 Td
7',5 T7 Z4 Th
h
O~ Tl Zl 0
TS Zs Th
T2 Z6 Oh
T7 T7 Z4 Th
h
T3 Z2 0
TS Zs Th
T4 Z7 Oh
T3 T4 Z4 Oh
d
T5 Z3 0
T5 Z5 Oh
T6 Zs Oh
05 T4 Z5 Oh
T7 Z4 0
T5 Z4 Oh
TS Zg Oh
OS T4 Z5 Oh
T5 Z4 Oh
ogh T7 Z4 Td
TS Zg Oh
Tg Z5 0
TlO ZlO Oh
128 CHAPTER 4
Lattice rh Lattice rh
M L M L
Group Group
IR I IR I IR I IR I
Sg Tl Zl T Tl Zl Th Cg Tl Zl T - -
T2 Z2 Th T2 Z2 Th C~ T2 Z2 Th - -
D l- 5 Tl Zl Td Tl Zl Oh C66
S
T2 Z2 0 T2 Z2 Oh CJh Tl Zl T Tl Zl Th
CJv Tl Zl Td Tl Zl Oh T2 Z2 Th T2 Z2 Th
C§v T2 Z2 0 T2 Z2 Oh CJh Tl Zl T Tl Zl Th
C~v Tl Zl Td - - T2 Zs Th T2 Zs Th
Civ T2 Z2 0 - - TS Z2 Th TS Z2 Th
DAd Tl Zl Td Tl Zl Oh T4 Z4 Th T4 Z4 Th
D~d T2 Zs Oh T2 Zs Oh Cih Tl Zl T - -
TS Z2 0 TS Z2 Oh T2 Zs Th - -
T4 Z4 Oh T4 Z4 Oh TS Z2 Th - -
D~d Tl Zl Td - - T4 Z4 Th - -
Did T2 Zs Oh - - Dg Tl Zl Td Tl Zl Oh
TS Z2 0 - - T2 Z4 0 T2 Z4 Oh
T4 Z4 Oh - - D4 TS Zs Oh TS Zs Oh
6
Cg Tl Zl T Tl Zl Th D 65 T4 Z2 Oh T4 Z2 Oh
C46 T2 Z2 Th T2 Z2 Th
Ct
ANALYSIS OF THE THERMODYNAMIC POTENTIAL 129
Lattice rh Lattice rh
M L M L
Group Group
IR I IR I IR I IR I
D~ Tl Zl Td - - Dgh Tl Zl Td - - -
D3
6 T2 Z4 0 - - D~h T2 Z3 Oh - - -
T3 Z3 Oh - - T3 Z2 0 - - -
D~ T4 Z2 Oh - - T4 Z4 Oh - - -
CJv Tl Zl Td Tl Zl Oh Dh Tl Zl Td Tl Zl Oh
T2 Z3 0 T2 Z3 Oh T2 Z5 Oh T2 Z5 Oh
T3 Z2 Oh T3 Z2 Oh T3 Z4 0 T3 Z4 Oh
T4 Z4 Oh T4 Z4 Oh T4 Zs Oh T4 Zs Oh
Cgv Tl Zl Td - - - T5 Z3 Oh Ts Z3 Oh
C~v T2 Z3 0 - - - T6 Z7 Oh T6 Z7 Oh
ctv T3 Z2 Oh - - - T7 Z2 Oh T7 Z2 Oh
T4 Z4 Oh - - - TS Z6 Oh TS Z6 Oh
D~h Tl Zl Td Tl Zl Oh D~h Tl Zl Td - - -
D5h T2 Z3 Oh T2 Z3 Oh D~h T2 Z5 Oh - - -
T3 Z2 0 T3 Z2 Oh D: h T3 Z4 0 - - -
T4 Z4 Oh T3 Z4 Oh T4 Zs Oh - - -
Ts Z3 Oh - - -
T6 Z7 Oh - - -
T7 Z2 Oh - - -
TS Z6 Oh - - -
130 CHAPTER 4
References
1. E. Spenser: Theory of Invariants, Mir (1974).
2. G.F. Smith, and R.S. Rivlin: Arch. Rational Mech. Anal. 12,93 (1963).
3. A.E. Green, and J .E. Adkins: Large Elastic Deformations and Non-Linear
Continuum Mechanics, Clarendon Press (1960).
4. W. During: Ann. Physics 7, 104 (1958).
5. Yu.M. Gufan: Fiz. Tverd. Tela 13, 225 (1971).
6. L. Michel, and J. Morzymas: in P.Kramer and A. Rieckers (eds.), Lecture
Notes in Physics, Springer, 79,447 (1978).
7. V.D. Kinev, V.E. Naish, and V.N. Syromyatnikov: Fiz. Met. Metalloved.
49, 241 (1980).
8. C.W. Curtis, and A. Reiner: Representation Theory of Finite Groups and
Associated Algebras, Wiley (1982).
9. V.E. Naish, Yu.N. Skryabin, and V.N. Syromyatnikov: Fiz. Met. Met-
alloved. 52, 1147 (1981).
10. Yu.M. Gufan: Zh. Eksp. Teor. Fiz. 60, 1537 (1971).
11. Yu.M. Gufan: Termodinamicheskaya Teoriya Fazovykh Perekhodov, Izdat.
Rostov. Univ. (1982).
12. Yu.M. Gufan, and V.P. Sakhnenko: Zh. Eksp. Teor. Fiz. 63, 1909 (1972).
13. V. Janovec, V. Dvorak, and J. Petzelt: Czechoslovak J. Phys. B 25, 1362
(1975).
14. E.V. Gorbunov, Yu.M. Gufan, N.A. Petrenko, and G.M. Chechin: Kristal-
lografiya 26, 8 (1981).
15. L. Michel, J .C. Toledano, and P. Toledano: in Symmetries and Broken
Symmetries in Condensed Matter Physics, IDSET (1981).
16. J .C. Toledano, and P. Toledano: Phys. Rev. B 21, 1139 (1980).
17. Yu.M. Gufan, and G.M. Chechin: Kristal/ografiya 25, 453 (1980).
18. P. Du Val: Homographies, Quaternions, Rotations, Oxford University Press
(1964).
19. Yu.I. Sirotin: Kristallografiya 12, 208 (1967).
CHAPTER 5
131
132 CHAPTER 5
(15.1)
Here ~n(1]).) is a polynomial of degree n in the variables 1]).. As a rule, the 1]).
are chosen as the op components.
In keeping with the Landau theory, one begins the construction of the
phase diagram by finding all the extrema of the potential ~, that is, by solving
a system of non-linear equations
a~ = 0, (\,,= 1, 2, ... ,I v ) ,
JOl (15.2)
v1]).
where Iv is the dimension of the relevant IR. The subsequent analysis is broken
up into several stages.
At the first stage one finds the types of solutions to equations (15.2). Each
solution type is specified by the number of non-zero 1]). components and by the
ratios between these. As shown in Chap. 3, all the solution types possible,
that is, all mode mixing types, are found from group-theoretic considerations,
each solution type corresponding to a low-symmetry phase of its own. The
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 133
one can find the ratios between the coefficients of the potential <I> at which a
given solution corresponds to a minimum. The equality sign in equation (15.3)
gives the expression for the stability boundary of the phase investigated.
The final stage in the construction of the phase diagram is to explore the
positive definiteness of the potential. Analysis of the first- and second-order
derivatives of the potential with respect to the parameters 1]>. provides informa-
tion about the behavior of <I> in the vicinity of local extrema as a function of the
values of the phenomenological coefficients involved in the expression (15.1);
in this case the behavior of the potential <I> in the other regions of the space
of the variables remains undefined. Two possibilities can evidently be singled
out. One of them is that the value of the potential <I> increases continuously
when the parameters 1]>. enhance along any direction in the multidimensional
space of the OP 11 = {1]>.}. In this case the local minima found from the condi-
tions (15.2) and (15.3) describe the stable states of the system into which the
latter may pass as a result of a phase transition. The other possibility is that
at certain values of the coefficients of the potential the <I> becomes negative at
least in one of the directions in the space of the parameters 1]>. and increases in
magnitude with 1]>.. In this case the local minima describe metastable states; in
constructing the phase diagram in the space of phenomenological coefficients, it
is therefore necessary, in terms of the potential model (15.1) chosen, to isolate
the region of positive definiteness of the potential <I> in which the local minima
describe stable states. When it is desired to explore a potential with values of
coefficients at which the condition of positive definiteness is violated, one has
to allow for the expansion terms that follow next in the series (15.1).
We now formulate the scheme of testing the thermodynamic potential
(15.1) for positive definiteness. The behavior of <I> at large values of the para-
134 CHAPTER 5
"'"'
~7] 12 >. -- 1. (15.4)
>.
The quantity <I>n (7]>.) then goes over into Rn<I>n (7]D and the condition of positive
definiteness for <I>n(7]>.) reduces to the requirement that the quantity <I>n(7]~) be
positive definite on the multidimensional sphere of unit radius prescribed by
equation (15.4). The quantity <I>n (7]D will be positive definite if it is positive at
the points of the minimum on the sphere (15.4). These points are found from
the equations
~<I>~ = 0, <I>n = <I>n(7]~) + Jj (2: 7]~ - 1), (15.5)
7]>. >.
°
trema 7]~o found from equations (15.5) are substituted into <I>n, and it is re-
quired to fulfill the condition <pn(7]~o) > which results in inequalities for the
coefficients involved in <I>n(7]~o).
The stages enumerated above are initial steps for the construction of the phase
diagram itself. The biggest task is to determine the stability boundaries of the
phases. To this end, it is necessary to simultaneously solve equations (15.3)
(in the case of the equality sign) with the equations of state (15.2). Practical
methods for performing such phase diagram analysis have been developed by
Gufan et al. [1-3]. These methods are based on the use of well-known concepts
in polynomial algebra.
From a mathematical viewpoint, the problem reduces to one of finding the
common roots of the polynomials
ao al ... an 0 0 ... 0
R(J,g) = 0
0 ao
0 ... 0
al ., .an
ao
0 ... 0
al ... an
} s lines
(15.7)
bo bi .. , b. 0 0 ... 0
0 bo bi ... b. 0 ... 0 } n lin"
0 0 ... 0 bo bi ... b.
The equation R(J, g) = 0 defines the stability boundary of a given phase in the
space of potential parameters. A further step is to find the phase transition
'lines' (surfaces). In those cases where the regions of stability of two neigh-
boring phases do not overlap, the stability boundary delimiting these phases
is the boundary of second-order phase transitions. But if there is a region
in which two or more phases coexist, a first-order phase transition boundary,
determined from the condition that the energies of those phases be equal in
magnitude, should pass in that region. The problem of finding first-order phase
transition 'lines' is solved in similar fashion to that of finding stability lines.
The polynomial f( x) in this case is determined from the equality of the poten-
tials of the neighboring phases.
Let it be noted that in analyzing phases described by more than one com-
ponent of the multicomponent OP, the number of equations of state is equal
to the number of non-zero OP components. In this case the stability lines and
the phase transition lines are found by successively calculating the necessary
number of resultants.
When finding stability lines and phase transition lines, one should allow for
the fact that the requirements that the principal minors be equal to zero and
the energies of the phases be equal in magnitude are fulfilled for all solutions
to the equations of state, including those corresponding to a maximum of the
potential as well as those corresponding to imaginary OP values. To single out
non-physical solutions, it is therefore reasonable to analyze the distribution
of the roots to equations of state in the space of thermodynamic-potential
coefficients, which permits one to isolate the regions of existence of real-valued
solutions.
The problem of the distribution of solutions to equations of state is solved
with the help of the Descartes rule [4].
1. The number of positive roots of a polynomial f( x) with real coefficients
136 CHAPTER 5
either is equal to the number of sign reversals in the series of coefficients I(x)
or is an even number less.
2. The number of negative roots is equal to the number of sign reversals
in the series of polynomial coefficients I( -x) or is an even number less.
Additional information about the distribution of roots to equations of state
may be derived from an analysis of the discriminant of the polynomial in ques-
tion [4]. The discriminant D(J) of the polynomial I(x) = aoxn + alx n - 1 +
... + an-1X + an is related to the resultant of the polynomial I(x) and to its
derivative I' (x) by the formula
R(J,J') = (_1)n(n-l)/2 ao D(J). (15.8)
If I(x) is a polynomial with real-valued coefficients and has no multiple
roots, then D(J) > 0 if the number of pairs of conjugate complex roots I( x) is
even. D(J) < 0 in the case of an odd number of conjugate complex roots.
In the subsequent sections of this chapter it will be shown in what fashion
these concepts can be exploited in a practical analysis of phase diagrams [3].
Despite the progress made using the theorems of polynomial algebra cited
above, the problem of constructing a phase diagram in the case of a tran-
sition over a multicomponent OP and with allowance for sixth-degree and
higher-order invariants in the thermodynamic potential expansion (15.1) in-
volves considerable technical difficulties. One of these difficulties is the neces-
sity to calculate literal determinants (resultants) of high order. Substantial
difficulties are encountered also in analyzing expressions for the phase transi-
tion lines and stability lines that are obtained by calculating such resultants.
The large number of phenomenological variables (coefficients in the thermody-
namic potential), which augments dramatically with increase in the degree of
the invariants allowed for in the potential and in the case of a multi compo-
nent OP, leads to a substantial increase of the various phase diagram versions.
Each version corresponds to a particular cross-section of the multidimensional
space of thermodynamic-potential coefficients in which the phase diagram is
constructed. That is to say, each version is characterized by a certain ratio of
these coefficients.
It might be well to point out that the above difficulties may be negotiated
by means of a computer. Problems of computer processing of high-degree
polynomials have been considered in the monograph [3], where all the problems
of polynomial algebra are treated on the basis of the concept of the innor
introduced by the author. Computer processing of resultants is employed also
in [5,6].
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 137
We begin the construction of phase diagrams with the simplest case of a one-
component OP, assuming the thermodynamic potential to contain only even
degrees in the expansion:
(16.1)
(16.2)
The following model contains a sixth degree in the expansion of <I> in 7]:
(16.3)
For the potential (16.3) to be positive at large values of 1171, the following
inequality should be fulfilled: v > 0; the coefficient u may in this case be of
arbitrary sign. The equation of state
(16.4)
138 CHAPTER 5
r r
7/-0 71=0
... lJ. LL
71, 711
(a)
II III (b)
Fig. 5.1. Stability regions of solutions: ( a) rJi and (b) 1}~ for the 1}6
model.
has the solution
2
1}1 = - -u+u
-3v g1 r v
- -2, (16.5)
3v u
1}~ = -~
3v
- ~J1-
3v
3rv.
u2
(16.6)
Allowing for the equation of state (16.4), the stability condition 8 2q,j81}2 > °
for these solutions may be written in the form
r + U1}2 < 0. (16.7)
Account must be taken also of the conditions under which the solutions (16.5)
and (16.6) are real. One of these conditions is
1 - 3rv >
u2 '
° (16.8)
while the others are formulated separately for each of the solutions, 1}I and 1}~:
u > 0, r < °
or u < 0, r> ° and u < 0. (16.9)
Simultaneous use of the stability conditions (16.7) and the reality conditions
(16.8) and (16.9) leads to the inequalities
1} = °
is defined by the inequality r > 0.
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 139
Finally, to construct the phase diagram, we find the phase transition lines.
From the condition <1>(1]) = 0 (equality of the energies in the phases with 1] ::P 0
and 1] = 0) and allowing for the equation of state, we obtain
(16.12)
The solution of this equation, 1]2 = 0, defines the line of second-order phase
transitions, r = O. As follows from the form of the solutions (16.5) and (16.6),
the condition 1]2 = 0 is satisfied only by one solution, namely 1]f. It is this
solution that describes the OP change at the second-order phase transition.
A similar conclusion may be drawn directly from an analysis of the stability
conditions for the solution 1]r. As can be seen from Fig. 5.1( a), the stability
region of the initial phase 1] = 0 and that of the low-symmetry phase 1]f do not
overlap, but they have a boundary in common; it is this boundary which is the
second-order phase transition line.
One more equation results from equation (16.12):
The expression (16.13) specifies the line of the first-order phase transition to
the phase described by the solution 1]~ (Fig. 5.1(b)). This line will be obtained
by substituting into equation (16.13) the expression (16.6) for 1]~:
U2
r=-. (16.14)
4v
The OP jump on the phase transition line (16.14) is
2r 2r
1] = - -; = - 4v . (16.15)
In final form, the phase diagram for the potential (16.3) is sketched in
Fig. 5.2. Thus the 1]6 model turns out to lead to an extension of the domain
of existence of dissymmetric phases on the (r, u) plane. With negative u a
new phase arises, the transition to this phase being of the first kind. With all
positive u the phase transition remains a second-order transition, just as in the
1]4 model.
Let us consider the consequences of allowing for the next following term of
degree eight in the potential <I> expansion. This problem is an example of
140 CHAPTER 5
"
'1"'0
I
Fig. 5.2. Phase diagram for the 7]6 model with a one-component order
parameter.
(16.16)
are cubic for the quantity 7]2. Following the standard procedure of constructing
phase diagrams, we could write out the solutions of this equation in general
form and then use these to analyze the conditions under which the phases are
stable and their energies are equal. Technically, however, this procedure turns
out to be very complicated. Therefore, as proposed in §15 [3], we begin the
construction of the phase diagram by analyzing the number of roots of the
equation of state in the space of the coefficients r, u, v, and w.
Figure 5.3 presents the distribution of the roots to equation (16.16) in the
plane of the variables rand u for the case v < 0 in keeping with the Descartes
rule. The notation (2,0)+ 1- implies that in the corresponding region of the
parameters rand u the equation of state has either two positive solutions for
7]2 or none and one negative root.
More complete information on the distribution of the roots of equation
(16.16) may be obtained by examining its discriminant. To this end, we rear-
range equation (16.16) in the form
y3 + 3py + 2q = 0, (16.17)
2 U
Y= 1] + -,
V
3p = - ,
4w
2w
r 2vu _ 3v 2
2q = 4w - (4w)2' u = u - 8w' (16.18)
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 141
"
becomes
f = ~u ±
2w
(_8_)
27w
1/2 (_U)3/2 (16.19)
The discriminant curve is shown in Fig. 5.4 (dot-and-dash line). The coordi-
nates of the points through which the discriminant curve passes are as follows:
(16.20)
The coordinate U c can be found by solving the cubic equation that results from
equation (16.19) by substituting into the latter the expression rc = O.
Availing ourselves of the familiar fact that in the case of D > 0 the cubic
equation has one real root and two conjugate roots, and for D < 0 three
different real roots, we are in a position to refine the picture as to how the
roots to the equation of state are distributed on the (r, u) plane (Fig. 5.4).
The 0 [{ line may be disregarded as this is the line of concurrence of two
negative roots for TJ2 and corresponds to imaginary values of the OP TJ.
142 CHAPTER 5
r
\. let
(2,+r) B
(n
\ I
])<0 ]»0
\
\c IJ LI
0 ---Q .
/-/---- -, .Y
/
/if (1+) /) "A
~
(f+)
]) >0 (3
(16.21)
Using the equation of state (16.16), the expression (16.21) may be rearranged
to read
(16.22)
The stability boundary is given by the system of two equations (16.22) and
(16.16). The condition for this system to be solvable is, as shown in §15, that
the resultant be equal to zero:
r 2u 3v 4w 0
0 r 2u 3v 4w
R= u 3v 6w 0 0 =0 (16.23)
0 u 3v 6w 0
0 0 u 3v 6w
Let us see what changes the states of the system undergo along the a{3,6(J'
line (Fig. 5.4). On the a - , segment the equation of state has one real root
and the thermodynamic potential has the shape characteristic of the TJ4 model
(Fig.5.5a). At the point, two more positive solutions arise on TJ2; the TJ depen-
dence of <I> is depicted in Fig.5.5b. In the region OAGO (Fig. 5.4) a first-order
phase transition is possible between the phases which is described by different
solutions of the equation of state (TJ1 and TJ3)' This is an isostructural phase
transition with the temperature dependence of the OP, shown (schematically)
in Fig. 5.6 [7]. Finally, in the region 8 - (J' the system again has one solution.
T
o
We will consider in addition those spatial regions (r, u) where two positive
solutions of the equation of state for ",2 exist. The corresponding potential is
portrayed in Fig. 5.5c. These regions, as is seen from Fig. 5.4, overlap with
the stability region of the initial phase", = 0, so the phase transition from the
initial phase to the low-symmetry phase in these regions will be a first-order
transition. As can be seen from Fig. 5.5a, a second-order phase transition
in the ",8 model is possible; the second-order phase transition line originates
at the point C and is described by the equation r = 0 (the boundary where
the stability regions of the initial and low-symmetry phases touch). The final
form of the phase diagram is shown in Fig. 5.7, where the dot-and-dash lines
indicate the stability boundaries of the phases (Fig. 5.4), the solid line is the
line of first-order phase transitions, and the dashed line is the line of second-
order phase transitions. On the ED line and its continuation to the left a phase
transition from the initial phase occurs, while on the EA line an isostructural
phase transition takes place.
Fig. 5.7. Phase diagram for the ",8 model with a one-component order
parameter [7].
Next we find the explicit form of the equations for the line of the phase
transitions depicted in Fig. 5.7. The AE line, corresponding to an isostructural
phase transition, is found from the condition that the minima of the potential
cI> sketched in Fig. 5.5b be equal in magnitude:
(16.24)
and is determined by the equation
V v3
r= - u - - - . (16.25)
2w 8w 2
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 145
The ED Line (and its continuation to the left) is found from the condition that
the energy of the initial phase be equal to that of the dissymmetric phase
cI>(7]) =0 (16.26)
allowing for the equation of state (16.16). The ED line is determined by the
equation
r* = C:)u* + (2:W) (-u*) 1/2 3/2 (16.27)
r
v
Fig. 5.8. Phase diagram in the space of the parameters (1', u, v) for
the 7]8 model in the case of a one-component order parameter.
cl>, occur only for negative coefficients in the highest degree with respect to 17
in the initial model. Addition of higher degrees with respect to 17 does not lead
to qualitatively novel phase transition types, but only increases the number of
isostructural transitions [7].
The succession of solutions to equations of state in the case when a new degree
17 4 is added to the thermodynamic potential has been investigated in general
form by Aizu [8]. For convenience in reasoning, we transcribe the potential
(16.1) as
(16.28)
(16.29)
and consider one of the solutions Ii (Cl, C2, ... ). Following Aizu, we will call the
function J;( Cl, C2, ... ) an n - s type function if the following two conditions are
fulfilled:
1. The function J;(Cl, ... ,Cn,Cn+l,O, ... ,O) is finite (different from zero)
at finite values of the coefficients Cl, ... , Cn+l and c" = 0, where K > n + 1.
2. °
lim Cn+di (Cl, ... , Cn , Cn+l, 0, ... , 0) "¥= at arbitrary values of the
C n +l -+0
coefficients Cl, ... , cn .
Assume also that all coefficients c,,' except the coefficient Cl, are temper-
ature independent and Cl ,...., (T - Tc).
To elucidate the meaning of the concept introduced, we determine to what
type the two solutions of the equation of state which we had in the 17 6 model
belong. In the new notation these solutions (16.5) and (16.6) have the form
(16.30)
(16.31 )
To start with, consider the solution !1(Cl, C2, C3). Assuming that C3 -+ 0,
expand equation (16.30) in a series of C3:
(16.32)
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 147
(16.33)
(16.34)
(16.35)
Thus the value of the OP at the phase transition point is equal to zero, so the
1 - s function (16.30) describes a second-order phase transition.
Consider the second solution (16.31). Evidently,
(16.36)
Allowing for the fact that the function h(cl, C2, C3) is finite at finite values of
the coefficients Cl,C2 and C3, we infer that h(cl,c2,c3) pertains to the 2 - s
type. We find the kind of phase transition described by the 2 - s function
f2(cl,c2,c3). Since
(16.37)
we conclude that the second solution to the equation of state, (16.31), which
is a 2 - s function, describes a first-order phase transition.
These results may be generalized and formulated in the form of the state-
ment that all equation of state solutions of n - s type with n > 1 describe
first-order phase transitions, while a solution of 1 - s type describes a second-
order phase transition. Thus, if there is an explicit form of the solutions to
the equation of state in the TJ2n model, it is easy to draw a conclusion as to
the kind of the phase transition to the state described by each solution. It is
noteworthy that while only a 1- s-type solution corresponds to a second-order
phase transition, the solution types corresponding to first-order transitions are
2 - s, 3 - s, etc. The superscript n may serve as a classificational index among
the various first-order phase transitions in a given system [8]. By adding new
terms to the thermodynamic potential, we shall naturally obtain new types
of solution to the equation of state, which correspond to higher values of n.
148 CHAPTER 5
Conversely, the breaking of a series for CP leads to the loss of some solutions of
the corresponding n - s type which describe first-order transitions.
On returning to the initial notations, C1 = r, C2 = U, C3 = v, etc., of the
thermodynamic potential (16.1) and comparing the expression (16.33) with the
solution (16.2) in the "14 model, we see that the solution (16.5) to the equation
of state in the "1 6 model derives from the unique solution (16.2) to the equation
of state for the preceding "14 model. As for the solution (16.6), it is qualitatively
new and is due entirely to allowance for the higher degree in the expansion of
the potential in the OP "1.
We will study the features peculiar to phase transitions due to the 2-component
order parameter by the example of systems characterized by a potential of the
form
(17.1)
This is the simplest and at the same time a sufficiently general potential that
may arise in the description of a phase transition going over the two-dimensional
representation of a large number of space groups specifying the initial phase.
By contrast with the one-component OP, it is not one but several dis-
symmetric phases differing in symmetry that may arise in the situation under
consideration. These phases correspond to different solutions of the equations
of state arising in the potential minimization process. For the potential (17.1)
there will be two such equations:
(The symbol ('T)l 'T)2) denotes that 'T)l :j; 'T)2 :j; 0.) The initial phase will be denoted
by 0. It is, however, a simple matter to see that the system of equations (17.2)
has no solution of the type ('T)1'T)2). Indeed, on setting up the difference of the
equations (17.2) (having preliminarily reduced the equations by the factors 'T)l
and 'T)2), we obtain the equation (2Ul - U2) (17f - 1]D = 0, which is satisfied
only if 'T)f = 'T)~. Thus only solutions of the first and second types enumerated
in equation (17.3) are possible at arbitrary Ul and U2.
Consider phase 1. The equations of state reduce to one:
(17.4)
whence we find 'T)2 -1' /2Ul, so that the OP is real subject to one of the
conditions:
l' < 0, Ul < ° or l' < 0, Ul > 0. (17.5)
We find the stability region of phase 1. To this end, we calculate the
second-order derivatives of the potential (17.1):
(17.7)
If we substitute here the expression for 1]2, determined from the equation of
state (17.4), then these inequalities may be reduced to the following two in-
equalities:
1':::; 0, (17.8)
The set of stability (17.8) and reality (17.5) conditions yields the ultimate
result: Phase 1 has a region of stability of real solutions which is given by the
inequalities (Fig. 5.9)
(17.9)
150 CHAPTER 5
Phase I
P hQ.se 2
Fig. 5.9. Phase diagram for the.,,4 model with a two-component order
parameter. 1, local-stability lines of phases; 2, positive-definiteness
lines of the potential <II.
(17.10)
(17.11)
(17.12)
The condition for the solution of the equation of state (17.10) to be real has
the form
r > 0, 2Ul + U2 < 0 or r < 0, 2Ul + U2 > o. (17.13)
Finally, the stability region of real solutions for phase 2 is restricted by the
inequalities (Fig. 5.9)
(17.14)
Completing the construction of the phase diagram, let us see what ad-
ditional constraints may be placed on the phase stability region by the re-
quirement that the potential (17.1) be positively definite. In keeping with
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 151
"'4
iF.
= U1 R4(-4+-4)+
"11 rl2 U2 R4(-2-2)
"71 "72 .
In turn, the requirement that <1>4 be positive definite reduces to exploring the
sign of <1>4 at the points of the minimum on the circle of unit radius r;r +
r;~ = 1 (§15). These minima are found from the condition fJ<I>4/fJ"7i = 0,
where <l>4(/-l) = <1>4 + /-l(r;r + r;~ - 1). The equations have three solutions:
(r;lo 0), (0 r;20) , (r;lo, ±r;2o). We substitute each of the solutions into <1>4. In
=
the case of (r;lo 0) or (0 r;2o) type solutions, /-l -2U1R2, r;~o =1, <l>4(r;>.o) =
R4u1. From the condition <l>4(/-l) > 0, we find that the coefficient U1 should be
positive, U1 > O. The solution (r;10, ±r;20) yields <1>4 (r;10 , ±r;20) = ~R4(2u1 +U2),
whence we find the second inequality: 2U1 + U2 > O. Thus the set of inequalities
U1 > 0,2U1 + U2 > 0 determines the region of positive definiteness of the
potential (17.1). This region is singled out in Fig. 5.9 by a double dot-and-
dash line. Thus the region of positive definiteness of <I> in this example does
not change the regions of existence of phases obtained from the local-stability
conditions. On the (ru2) plane the regions of stability of both dissymmetric
phases are shown in Fig. 5.10.
-2uI
1
l' hQ.se 2
2uI
Phose I
... z
U
Fig. 5.10. Stability regions of the phases in the "7 4 model in the (r, U2)
plane.
The "7 4 model, as is seen from the results obtained, does not describe the
transition between phases 1 and 2. Indeed, phases 1 and 2 have a common plane
2U1 - U2 = 0 on which the stability conditions for these phases are violated.
The phase transition line 1 f-+ 2 may be expected to lie in this plane. This line
should be found from the condition for the potentials of the two phases to be
equal: <1>("71 = "7, "72 = 0) = <1>("71 = "7, "72 = "7). On substituting the solutions of
the equations of state (17.4) and (17.10), the condition for the potentials to be
152 CHAPTER 5
equal becomes
r( 2U l - U2) _ 0
4U1(2u1 + U2) - .
This equation does not give us an expression for the transition line, that is, the
relation of the temperature-dependent coefficient r to the other parameters, U1
and U2. The problem whether a phase transition is possible between phases 1
and 2 can be solved only by going beyond the framework of the 1J4 model, that
is, by adding higher-order terms to the potential.
THE 17 6 MODEL
In the potential (17.1) we now take into account the sixth-order terms, which
we represent as (l7f + 1J~? and (l7f + l7n1Jfl7~' This form is equivalent to the
assumption that the IRBI for the potential contains only two invariants: l7f+l7~
and l7fl7~' Note also that the fourth-order term l7t + l7i in equation (17.1) may
be expressed in terms of the same two invariants. We can largely simplify the
intermediate calculations and bring the phase diagram into a more symmetric
form if we rewrite the resulting expression for <I> in equivalent form:
To analyze similar expressions that are symmetric with respect to the substi-
tu tion 171 f-+ 172, it is convenient to introduce the trigonometric variables
(17.17)
The requirement that the potentials (17.15) and (17.17) be positive definite
leads to constraints on the values of the coefficients V1 and V2: V1 > 0, -V1 <
V2 < V1'
We begin the construction of the phase diagram by analyzing the equations
of state
(17.18)
(17.19)
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 153
From the second equation of state, (17.19), it follows that either sin 4¢ = 0
or U2 + V2"1 2 = O. The case sin 4¢ = 0 refers to two phases:
cos4¢ = +1, "1 =1= 0 (phase 1+), cos4¢ = -1,"1 =1= 0 (phase 1-), (17.20)
(17.22)
(17.23)
reduces to the inequality "1 6 sin 2 4¢ < 0, which does not hold with real "1. Thus
phase 2 has no stability region in the "1 6 model. Phase 2 will occur for the first
time in the phase diagram only in the successive "1 8 model.
Consider now phases 1+ and 1-. The equations of state reduce to one
equation,
(17.24)
Fig. 5.11. Stability regions of the phases for the 1]6 model in the
(r,u2) plane.
by a dot-and-dash line in Fig. 5.11 are the stability lines of phases 1+ and 1-,
obtained from the first stability condition (for the case Ul > 0).
The stability region of phase 1+ is determined by the inequalities
(17.26)
For phase 1- ,
(17.28)
In Fig. 5.11 these lines are indicated by a double dot-and-dash line for the case
V2 < O. Inspection of this figure shows that phases 1+ and 1- have a region
of overlap through which a first-order phase transition line 1+ +-+ 1- should
pass. The final form of the phase diagram is given in Fig. 5.12. As can be
readily seen, the expressions for the first-order phase transition lines 0 +-+ 1+
and 0 +-+ 1- coincide, up to the replacement of the variables, with the case of
the one-component OP:
Ul ± U2
(17.29)
r= 3(Vl±V2)'
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 155
Fig. 5.12. Phase diagram for the 17 6 model in the (r, U2) plane.
The first-order phase transition line 1+ ...... 1- may be found from the
condition <1>(17+) = <1>(17-) which is easy to reduce to the form
(17.30)
(17.31)
°
for phase 2, (17,17) for phase 3, (17, -17) for phase 4, and (171,172) for phase 5. We
check whether these types of solutions satisfy the equations of state. Phase
is a trivial solution of the equation of state. The solution (±17, 0) leads to the
quadratic equation for 17:
(17.33)
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 157
The next solution (0, TJ) substituted into equation (17.32) yields two equations:
whence it follows that phase 3 (as well as phase 4) exists only on the line which
in parametric form is described by equations (17.34).
In the case of a general-type solution (TJ1, TJ2) we find that the equation
3v(3TJr - TJ~) = 0 holds, from which it follows that 3TJr = TJ~. The quantity is
determinable from the equation
(17.35)
TJ1 2 -
_ - -
3v ± [(3V)2 r ] 1/2
-8u - -2u . (17.36)
, 8u
From the condition for the square root to be positive, we find the region in
which the solution is real:
9v 2
r<-.
- 32u
(17.37)
The stability region of the solution (17.36) is found by the standard technique.
The region is determined by two inequalities:
Fig. 5.13. Phase diagram for the TJ4 model with a cubic term in the
(r,v) plane.
The phase transition lines 0 +-!- 1± are found from the equality of the
thermodynamic potentials of the phases, <I>(TJ) = o. This leads to the equation
which should be solved simultaneously with the equation of state (17.33). Thus
we obtain the relation between r and v:
v2
r=-, (17.40)
4u
which is shown in Fig. 5.13 by a solid line. The OP jump on this line is
(17.41)
invariant even in the 1J4 model leads to a first-order phase transition. A con-
sequence of truncating the series for the potential on fourth-degree invariants
was, as shown in the foregoing, the absence of a whole number of stable solu-
tions describing the low-symmetry phases. In the absence of cubic invariants
(potential (17.1)) the phase transitions from the initial phase to dissymmetric
ones are second-order transitions. To ascertain the character of the phase tran-
sition between two dissymmetric phases, higher-order terms should be added
to the potential.
We begin the analysis of the phase diagram with the phases 1±. The OP in
this phase satisfies the equation of state
(17.43)
We find the inequalities determining the stability region of this phase. The
matrix D, formed from the second-order derivatives of the potential <I> with
respect to the variables 1Jl and 1J2, assumes diagonal form on substituting the
= =
solution 1Jl ±1J, 1J2 O. The stability conditions reduce to two inequalities:
A major difficulty in analyzing the equations of state (17.43) and the stability
conditions (17.44) is that the solution of the quartic equation cannot be written
out in explicit form. Therefore, in keeping with the recommendations of §15,
we proceed in the following fashion.
We find the stability line of the phases 1±. To this end, we leave only
the equality sign in the inequalities (17.44). After a little manipulation the
equation of state and the first stability condition are then reducible to
An explicit expression for the stability line can be found from the require-
ment that the resultant of the system (17.45) be equal to zero. Assuming the
parameter v to be small, it is however possible not to calculate the entire resul-
tant completely but to exploit perturbation theory [3]. As a result we obtain
the coupling between the parameters r, v, U, W, ..• as an expansion in powers
of v:
9v 2 5w 3
r= 32u + 2(8u)3(3v) . (17.46)
(17.47)
r = - ~ (_v)3 / 2
2uv ± __ (17.48)
W w 5/ 2
with ~ = 4UZ2 - w 2 . The lines (17.46) and (17.48) are shown in Fig. 5.14 by
a dot-and-dash line. From equation (17.48), it is seen that when ~ > 0 the
regions of existence of the phases 1+ and 1- do not overlap, and when ~ < 0
a region of coexistence of the phases 1+ and 1- occurs.
\~ .
r
l
-' ,,/
~/ II u
.<1/
--~~~-- "-",. ~~--------
f+
//
2 .
I
Fig. 5.14. Phase diagram for the T/6 model with a cubic term in the
(r, v) plane.
As has been shown in [3], the region of existence of phase 2 (solution type
(T/IT/2)) is bounded by the lines (17.48), so the phase transitions 1++-+ 2 and
1- +-+ 2 at ~ > 0 will be second-order phase transitions. For ~ < 0 phase 2
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 161
Fig. 5.15. Phase diagram for the rl model with a cubic term in the
space of the variables (r, v, A).
has no stability region. The phase transitions 0 +-+ 1+ and 0 +-+ 1-, just as in
the "14 model, will be of first order. The lines of first-order phase transitions
can also be found with the help of perturbation theory. In final form, the phase
diagram is presented in Fig. 5.15.
(18.1)
whence it is seen that the general-type solution (1J11J2173) in the 1J4 model exists
only at the isolated point 2Ul = U2, while in the rest of the region of the
parameters .,., Ul, and U2 the system of equations (18.4) has the solution (1J 1J 1J),
which corresponds to phase 1. Similarly, it may be shown that in the 1J4 model
phase 5 coincides with phase 1, and phase 4 with phase 3.
We now explore the stability conditions for phases 1, 2, and 3. The stability
analysis reduces to determining those values of the parameters.,., Ul, and U2 for
which a given type of solution to the equations of state corresponds to a local
minimum of the potential (18.1). To this end, as was noted in §15, we should
require that all the principal minors of the matrix D constructed from the
second-order derivatives {)~/{)1J)..{)1J1' be positive definite. In the model being
considered (18.1) this matrix has the form
(18.6)
The first stability condition follows from the requirement that the first principal
minor of the matrix (18.15) be positive definite. Allowing for equation (18.6),
this gives
(18.7)
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 163
The requirement that the second principal minor be positive definite gives us
the second stability condition for phase 3:
(18.8)
(18.9)
the second system of inequalities being incompatible in view of the fact that
Ul > O.
Finally, the third stability condition reduces to the requirement that the
determinant of the matrix D be positive definite. This leads to the inequality
(18.10)
Substituting the equation of state (18.6) into the inequality (18.10) yields
r
2Ul - U2
> 0. (18.11)
2Ul + U2 -
Taking into account the first and second stability conditions (18.7), (18.9), it
follows from this inequality that r ~ O. However, this requirement contradicts
the requirement that the solution to the equation of state (18.6) be real. Thus
phase 3 has no stability region in the T}4 model.
Consider now phase 1. On inserting the form of the solution T}l = T}2 =
T}3 = T}, corresponding to phase 1 in equations (18.2), we obtain the equation
of state for the parameter T}: r + 2( Ul + U2)T}2 = O. From this equation, we find
(18.12)
The first stability condition gives Ul > O. Allowing for Ul > 0, the second
condition leads to the inequalities
(18.13)
(18.14)
From the conditions for the solution (18.12) to be real, it follows that r < O.
Thus the stability region of phase 1 is determined by the inequalities
(18.15)
164 CHAPTER 5
Now we pass on to phase 2. The equation of state for phase 2 has the form
r + 2U1 'f/2 = 0 and the OP value for this phase is
2 r
'f/ = - - . (18.16)
2U1
The first stability condition gives U1 ;:::: O. The second condition reduces
to the inequality
(18.17)
which, allowing for the fact that the expression (18.16) for 'f/ is real, gives two
inequalities. In final form,
(18.18)
The stability regions of phases 1 and 2 are bounded by the inequalities (18.15)
and (18.18) respectively, as shown in Fig. 5.16. The dissymmetric phases 1
and 2 border on each other on the U1 = U2/2 plane. Just as in the case of a
two-component OP, it is impossible in the 'f/4 model to solve the problem of the
phase transition between these phases. Indeed, the equality of the potentials
<'P('f/ 0 0) = <'P('f/ 'f/ 'f/) leads to the expression
r 2 ( 2u 1 - U2) _ 0
(18.19)
U1(U1 + U2) - ,
which does not give the coupling between all the three coefficients r, U1 and
U2. From the condition for the potentials of phases 1 and 2 to be equal to zero,
Fig. 5.16. Stability regions of phases 1 and 2 for the potential (18.1)
with a three-component order parameter.
Exploiting the concept of the IRBI enables one to view the phase diagram
construction methods in a different light. We wish to enumerate the chief
advantages that the IRBI technique can provide in this problem. First, finding
the IRBI permits the thermodynamic potential to be immediately expanded
up to any power desired. Second, a knowledge of the IRBI enables one to single
out those terms of a given degree in the potential ~ that account for qualitative
changes in the phase diagram. Third, the utilization of the IRBI may come
to play an essential role in probing low-symmetry dissymmetric phases. As
shown in the foregoing sections, such phases occur in phase diagrams only
when invariants of sufficiently high degrees are allowed for in the potential.
The technical difficulties that arise here naturally lead us to the desire to learn
to what lowest degree we can confine ourselves without losing the phase of
interest.
166 CHAPTER 5
(19.1)
(19.2)
We write out the equations of state, regarding <I> as a composite function of 771
and 772, in terms of the invariants 11 and 12 :
where the derivatives with respect to the invariants of the series for <I> are
{)<I> 2
= {)I1 = r + 2U1h + wI2 + 3z 1 I 1 ,
I _
<I>1
(19.4)
I {)<I>
<I>2 == {)h = v + wl1 + 2z 2 12 .
(The use of the minimization of the potential with respect to invariants was
first proposed in [10]). Without placing in advance any constraints on 'T}1 and
772, it is easy to see from the equations of state (19.3) that in the case of <I>~ -:f. 0
and <I>~ -:f. 0 the formal solution has the form
(19.5)
which, just as in the 'T}4 model (§17), corresponds to a solution of the type
(±77 0) (phase 1±).
To obtain equations for finding a general-type solution, we should require
As follows from the explicit form of the expressions (19.4) for <I>~ and <I>~, the
system of equations (19.6) may yield solutions for 'T}1 and 'T}2 provided only
that at least the fifth-degree invariants are allowed for in the potential (19.2).
However, truncating the potential at degree five results in the possibility that
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 167
As another example, we consider the IRBI for the group 1 = Oh. This basis
consists of three invariants (§14):
(19.7)
(19.8)
(With allowance for fourth-degree invariants only, this potential was treated in
detail in §18.) We write out the equations of state
in terms of the derivatives of cI> with respect to the invariants h, 12, and h:
cI>~ == ocI>/oh = r + 2U1h + 3W11[ + w212 + ... ,
cI>~ == ocI>/oI2 = U2 + w212 + ... , (19.10)
cI>~ == ocI>/oI3 = W3 + ....
All the solution types possible for the potential (19.8) are listed in §18. Corre-
sponding to the lowest-symmetry phase 6 (solution type (1]11]21]3)) is the system
of equations
cI>~ = 0, cI>~ = 0, cI>~ = O. (19.11)
Note that in the notation of the equations of state (19.9) a general-type solu tion
always reduces to the requirement that cI>: = O. It follows immediately from
the expressions (19.10) for cI>i that the lowest degree that should be taken
into account in the thermodynamic potential in order to obtain a general-type
sol u tion (1]11]21]3) is the eighth.
168 CHAPTER 5
The equations of state for phase 4 (solution type (171 172 0)) reduce to
<I>~ = 0 and <I>~ = O. Again, it is immediately seen from the expressions (19.10)
that allowance for invariants of degree four alone, as made in §18, gives no
unambiguous solution for the parameters 171 and 172. To obtain the relevant
solution, it is necessary, at least, to allow for invariants of degree six.
Let it be emphasized that even on determining the lowest degree at which
the series for <I> should be truncated to obtain the solutions of interest, we still
have no guarantee that a given phase will have no finite stability region, as was
the case with phase 3 (17 17 0) in the 174 model.
(20.2)
1. 7]
2 Tl
= --,
2Ul
e= 0;
2. 7]- 0
-,
e - --'
2
-
T2
2U2' (20.3)
2 T2 W - 2U2Tl c2 Tl W - 2U1T2
3. 7] = d , .. = d '
with d = 4Ul U2 - w 2 . From the expressions (20.3) we find the regions in which
the solutions (20.3) for phases 1 to 3 are real:
1. < 0, Ul > 0;
Tl
2. T2 < 0, U2 > 0;
(20.4)
3. T2W - 2U2Tl > 0, T1W - 2U1T2 > 0 (d > 0);
T2W - 2U2Tl < 0, T1W - 2U1T2 < 0 (d < 0).
The stability regions of these phases are determined by the inequalities
(20.5)
Simultaneously with the conditions of existence (20.4), they have the form
2 II 1; 0
-~~. . L=-~=~2
/'1 f
./ /
i
(a) (b) (c)
Fig. 5.17. Phase diagram in the case of two order parameters: (a)
.6. > 0, 2(U1U2)1/2 > w > 0; (b) .6. > 0, -2(U1U2)1/2 < w < 0; (c)
w > 2( Ul U2)1/2. The dot-and-dash line represents the phase stability
boundary; the dashed line represents second-order transitions; the
solid line represents first-order transitions.
(20.7)
It was shown in the foregoing that at certain ratios for the thermodynamic
potential coefficients a phase diagram with a bicritical point is realized (Fig.
5.17c). This type of phase diagram corresponds to the situation where the
sequence of transitions occurring in the crystal when the external parameters
are varied is first 0 +-+ 2 and then 2 +-+ 1. Some sort of a supplanting of one
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 171
2 r,
...........
J ---
\ 1
'2
---:; f
I
(a) (b)
e
Fig. 5.18. Phase diagrams in the "76 and 6 models for: (a) Ul < 0,
U2 < 0, -2(U1U2)1/2 < w < 0; (b) Ul > 0, U2 > 0, W> 2(U1U2)1/2.
The other equations are obtained from the equations of state (20,9) by cyclic
permutation of the parameters rJ).. and 6, The solutions to the equations of
state describe phases 0 through 7, which have different symmetry:
3 ,' c - c - 0'
,,2 -
"I -
,,2
,,2 =
,,2 _
,,3 - - 2( U1 + U3)'
r1
.. 1 - .. 2 - ,
4 ', "
"I -
- ,,
,,2 -
- " c 2 + "2
,,3 -- 0 '''1 c2 -_ r2,
-- ,
2U2
5: 2 - 2U 2r 1 + wr2 0 c2 c2 r1 w - 2U1r2
rJ1 = .6. ' rJ2 = rJ3 = '''1 +"2 = .6. ;
6: 2 2 wr2 - 2U2r1 0 c2 c2 2wr1 - (2U1 + U2)r2
rJ1 = rJ2 = .6. ' rJ3 = '''1 + "2 = .6. 1 ;
(20,11)
(20,12)
The ratio of the structural and magnetic OP's on this line depends only on the
ratio UI/U2, which can be determined experimentally,
\Ve conclude by making one more statement concerning the example con-
sidered, From experimental data on the magnetic transition in MnAs, it follows
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 173
that the crystalline symmetry of the magnetically ordered state coincides with
the symmetry of the initial phase. The possibility itself of there existing a
magnetically ordered phase with the same crystalline symmetry as that in a
non-magnetic phase is known to be present only in the case of a magnetic
transition according to a one-dimensional representation (in keeping with the
Niggli-Indenboom theorem [16]). At the same time the analysis performed
above shows that the relevant representation at a magnetic phase transition
in MnAs is two-dimensional rather than one-dimensional (the collinear ferro-
magnetism in the basal plane in a uniaxial crystal is always associated with the
two-dimensional representation). True, the potential <I> as a function of two OP
components up to fourth-order terms inclusive depends only on the sum of the
a
squares of these components +e~, that is, is quasi-dependent on the quantity
e alone. This corresponds to the situation where the magnetic anisotropy in
the basal plane in a hexagonal crystal is described only by invariants starting
with degree six. Thus the one-component ness of the magnetic OP is justified
only to within fourth-order invariants. On the other hand, although the fact
of the crystal symmetry being restored at the magnetic transition is held to be
experimentally established, it is clear that this fact is true only to within some
experimental accuracy, namely up to anisotropic sixth-order invariants again.
No exact symmetry restoration may occur here, for we deal with the collinear
ferromagnetism in the basal plane, and therefore, on account of the lattice ad-
justing itself inevitably to the magnetic order that arises, the symmetry of the
crystal cannot become higher than the orthorhombic symmetry.
• - K
II- - Mn
O-F
only one star arm, for example ~1, participating in the second transition. The
magnetic structures arising at the points 111 and T.l are characterized also by
the star {~13}.
In order to construct the potential describing the phase transitions in
KMnF 3 , it is necessary to determine by which IR each of the transitions is
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 175
<I> = ~1''1 (1]r + 1]~ + 1]~) + tUI (1]{ + 1]i + 17i) + ~U2( 1]r1]~ + 1]r1]~ + 1]~1]~)+
+~1'e(~r + ~i + ~~) + t Vl(a + ~i + ~i) + ~V2(~~~~ + a~~ + ~~~~)+
+~wl(1]i~r + 1]~~~ + 1]~~~) + ~w2[1]i(~~ + ~D + 1]~(~i + ~~) + 1]~(~i + ~~)]+
+W3(1]1172~16 + 1721]366 + 1]31]166), (20.13)
where 1''1 = 1'~(T - T~) and 1'e = 1'€(T - T~,) with T~ and T,~, being the "bare"
temperatures of the structural and the magnetic phase transition in the absence
of a coupling between them.
Each phase is characterized by a six-component vector (1]1, 1]2,1]3,6,6,6).
We investigate the conditions for the succession of the three phases (1 to 3)
observed experimentally:
Phase 2
2 -1''lVI + 1'e W I ('2 _ _ 1'e U I + 1''lWI (20.18)
1]11 = Ul VI - wi '''II - UI VI - Wf '
176 CHAPTER 5
(20.19)
(20.20)
Phase 3
(20.22)
(20.25)
[( U2 - Ut)171. + (WI - W2)~1.][( V2 - vt)~1. + (WI - W2)17rJ - w~17H1. > O.
(20.26)
The phase transition at the point Tel is, as can be seen from the expression
(20.15), of second order, like the 1 f-t 2 transition at the point 111' The phase
transition 2 f-t 3 at the point Tl. turns out to be a first-order transition. The
transition temperature (20.23) is found from the equality of the thermodynamic
potentials <1>2 and <1>3. The OP jumps at T = Tl. should be related by the
equations [17,18]
17IT _ 1- (1 + DJ:I)(wIr~)/(vIr{) ~~
(20.27)
171. 1-(1+D.LI)(w2r~)/(vIre)· ~1'
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 177
C4 2
"'II _ UI VI - W2 _ -2 1 (20.28)
e1 - UI VI - Wi - t <.
This applies to the point T = Tl.. At the point T = 111 the magnetic OP
that arises is of second order, the structure parameter being continuous. The
temperature behavior of the OP in KMnF 3 is sketched in Fig. 5.20. Dis-
continuities of the structural OP at the point Tl. were observed in a neutron
diffraction study on a KMnF 3 single crystal, where a magnetic peak intensity
hysteresis was also detected indicating a first-order phase transition at the same
point [17]. This theory of the coupling of two OP's enables one to understand
the regularities observed in the structural and magnetic phase transition in
KMnF 3 .
invariants since the law of conservation of momentum holds (to within the
parent-crystal reciprocal-lattice vector) for such terms. Displacements of F
may bring about displacements of other light atoms, primarily those of K.
Such light-atom displacements were detected experimentally at T < Tc2 [19].
Let some OP X>. correspond to this type of displacement. The occurrence of
these displacements in the phase where the OP's R>. and M>. were condensed
calls for invariants such as RM X. It follows from the law of conservation of
momentum that the parameter X should be specified by the star {,dOl:
(20.29)
where r '" (T-Tcd, r' '" (T-Tc2)' and all the other parameters depend weakly
on temperature. Minimization of <I> with respect to X>. yields the equation
(20.31 )
which shows that the displacement of K atoms should accompany the displace-
ment of F atoms in the R + M phase at T < T c2 . Since the coefficient J( is
temperature independent, the parameter X>. is not an OP in the strict sense
of the word; the X>. may be called a concomitant OP. The crystal distortion
that is associated with it does not change the phase symmetry at T < T c2 . The
space groups of the various dissymmetric phases are as follows:
(20.32)
ORIENTATION TRANSITIONS
<I> = 1'1(17r + 17~ + 17~) + U1(17r + 17~ + 17~)2 + U2(17r17~ + 17i17~ + 17~17~)+
+V1(17i + 17~ + 17~)3 + V217i17~171+
+V3[17r17~(17r + 17~) + 17i17~(17i + 171) + 17~171(17~ + 171)]' (20.33)
where 171, 172, 173 may be viewed as magnetization vector components. Po-
tentials of this kind were explored in detail in the preceding sections. In the
construction of a phase diagram, no restrictions are imposed in the general case
on the value of the vector 11 = (171,172,173) or its angular position. A distinctive
feature of a thermodynamic description of orientational phase transitions is
that the magnitude of the magnetization vector is normally assumed to be in-
dependent of temperature. Only the angular dependence of the magnetization
vector is investigated. It is therefore convenient to choose direction cosines Cti
as the variables: 17i = 17Cti and to write the potential (20.33) in the form
(20.34)
with <I>(17) the isotropic part of the potential; J{l and J{2 anisotropy constants,
which are some combinations of the starting parameters Ui and Vi.
The potential (20.34) describes three phases: phase I with the magneti-
zation vector along [100], phase II with the magnetization vector along [110],
and phase III with the magnetization vector along [111]. The phase diagram
180 CHAPTER 5
portrayed in Fig. 5.21 indicates at what values of the anisotropy constants ](1
and ](2 each of the aforecited phases occurs. Thus the model (20.34) under
consideration describes orientational first-order transitions (If-tIII and IIf-tIII)
and orientational second-order transitions (If-tII) between phases with fixed
magnetic moment direction.
~K.~I(=DI
"", j _1. I
" !! I I
" I
I
________~~I~-------- J'l
to the potential (20.34) will lead to the appearance of two angular phases in
the phase diagram: phase IV with the magnetization vector along the [u v 0]
direction and phase V with the magnetization vector in the general direction.
The stability regions of phases IV and V are determined by the inequalities
[22]:
Phase IV:
(20.36)
Phase V:
e
where TJ and may be regarded as two magnetization vector projections. Since
the symmetry group of RF0 3 is a rhombic one, each magnetization vector
projection transforms by its individual one-dimensional IR, and this is what
determines the choice of the potential in the form of equation (20.37).
300
200
100
o 02 04 06 08
(20.38)
For J{2 > 0 all the three phases have stability regions. In the case of J{2 < 0
phase III is unstable.
The results cited above coincide with the more general results (20.35) and
(20.36). Phases I and II are virtually (77 0) and (0 e) type phases. Phase III
is a phase of type (77 e). Just as in the preceding case, the directions of the
magnetization vector in phases I and II coincide with one of the crystallographic
directions. Phase III is said to be angular; it corresponds to an arbitrary
magnetization vector direction. The temperature dependence of the anisotropy
constants J{i leads to continuous rotation of the magnetization vector in phase
III as the temperature is varied. Experimental data on rare-earth orthoferrites
are reviewed in [22].
We consider the main properties of induced orientation transitions. As
already stated, induced transitions are orientation transitions induced by an
external field H. The influence of an external field on a magnetic structure
depends on the type of structure considered.
Consider orientation transitions in antiferromagnetics. The major features
peculiar to the behavior of antiferromagnetic structures in an external field
may be illustrated by the example of a two-sub lattice anti-ferromagnet. This
is usually described by the magnetization vector M = MI + M2 and antiferro-
magnetism vector L = MI - M 2, where MI and M2 are the magnetization
vectors of the sublattices 1 and 2. As in the previous example, we assume the
effect of an external field to reduce only to rotation of the vectors Ml and M2
and to a change of the angle between them, the length of the vectors remaining
constant: MI = M2 = M. We assume also that the vectors M I , M 2 , and H
lie in the same plane (Fig. 5.23) and the anisotropy axis coincides with the
z axis. It is convenient to write the thermodynamic potential in terms of the
variables MI and M 2:
(20.40)
field on the occurrence of that structure. Instead, we are concerned with the
relative position of magnetic moments in a magnetic structure of a given type
as a function of the orientation of the applied field H. In the general case a
potential such as (20.40) can be obtained in the usual way. That is, using the
IRBI method (see §12), a potential symmetric under the paramagnetic-phase
symmetry group is constructed for a specific crystal. Then we substitute into
this potential the OP corresponding to the magnetic structure type chosen,
impose the conditions for the magnetic moments to be constant and obtain, as
a result, a potential of the form (20.40).
To start with, we consider the cases H " z and H ..L z (1jJ = 0, 7r /2). Since
the position of the vectors L, M 1 , and M2 is unambiguously prescribed by the
angles () and e/; (Fig. 5.23), we seek the equilibrium position of the vectors Ml
and M2 in the equations
8'P 8'P
8() = 0, 8e/; = O. (20.41)
(20.44)
(;~ - ~, -cos 2'1jJ) sin 28- (sin 2'1jJ-2;~ ~, sin 28) cos 28 = ~ sin 2'1jJ, (20.45)
We differentiate equation (20.46) with respect to the angle e. Using the ex-
pression obtained, we may eliminate the angle e from equation (20.46). As a
result, we obtain the equation for Hz and Hx:
The resulting equation determines two astroids on the (x, z) plane (Fig. 5.24).
Substitution of Hx = 0 into equation (2.47) yields the expression (20.44) for Hl
and H 2 . Over the field intervals enclosed within the astroids, an orientational
first-order phase transition is possible in which the vector L rotates from the z
axis to the x axis. The maximal angle at which such an orientation transition
is possible is determined by the critical point of the astroid [25] and is equal to
7/Jcr = 2~'
-f--------+-------l---::- Ii j.
HS1 ,}(
The ellipse on the critical-field diagram (Fig. 5.24) determines the values
of Hsf at which a 'spin-flop' transition takes place. The major and minor ellipse
semi-axes are found from the stability conditions (20.43) and are equal to
Hsfx~M(2A+a+b),
, ,
Hsfz~M(2A-a+b). (20.48)
References
1. Yu.M. Gufan, and V.P. Sakhnenko: Fiz. Tverd. Tela 14, 1915 (1972).
2. Yu.M. Gufan, and V.P. Sakhnenko: Zh. Eksp. Teor. Fiz. 69, 1428 (1975).
3. Yu.M. Gufan: Termodinamicheskaya Teoriya Fazovykh Perekhodov (Ther-
modynamical Theory of Phase Transitions), Izdat. Rostov. Univ. (1982).
4. A.P. Minina, and N.V. Proskuryakov: Vysshaya Algebra (Higher Algebra),
Fizmatgiz (1962).
5. Yu.M. Gufan, V.P. Dmitriyev, V.P. Popov, and G.M. Chechin: Fiz. Tverd.
Tela 21,554 (1979).
6. Yu.M. Gufan, V.P. Dmitriyev, V.P. Popov, and G.M. Chechin: Fiz. Met.
Metalloved. 27,787 (1979).
7. Yu.M. Gufan, and E.S. Larin: Izv. Akad. Nauk SSSR, Ser. Fiz. 43, 1567
(1979).
8. K. Aizu: Phys. Rev. B 23, 1292 (1981).
9. V.E. Yutkevich, B.N. Rolov, and A.I. Medovoy: Czechoslovak J. Phys. B
31,77 (1981).
10. Yu.M. Gufan, and V.P. Sakhnenko: Zh. Eksp. Teor. Fiz. 63, 1909 (1972).
11. K.S. Aleksandrov, A.T. Anistratov, B.V. Beznosikov, and N.V. Fedoseeva:
Fazovye Perekhody v Kristallakh Galoidnykh Soedinenii ABX3 (Phase Tran-
sitions in Crystals of Halide Compounds ABX3 ), Nauka (1981).
12. E.M. Lifschitz: Zh. Eksp. Teor. Fiz. 14, 353 (1944).
13. I.F. Lyuksyutov, V.L. Pokrovskii, and D.E. Khmel'nitskii: Zh. Eksp. Teor.
Fiz. 69, 1817 (1975).
14. N.P. Grazhdankina: Uspekhi Fiz. Nauk 96,291 (1968).
15. V.E. Naish, Yu.N. Skryabin, and V.N. Syromyatnikov: Fiz. Met. Met-
alloved. 52, 1147 (1981).
16. V.L. Indendom: Kristallografiya 4,619 (1959).
17. Yu.A. Izyumov, F.A. Kassan-Ogly, and V.E. Naish: Fiz. Met. Metalloved.
51,500 (1981).
18. X. Convent, and N.M. Plakida: Acta Phys. Polon. 63,779 (1983).
19. M. Hikada: J. Phys. Soc. Japan 39, 180 (1975).
20. N .M. Plakida, A. Podolska-Strycharska, and V. Sikora: Kristallografiya 85,
2189 (1983).
21. H. Horner, and C.M. Varma: Phys. Rev. Lett. 20, 845 (1968).
22. K.P. Belov, A.K. Zvezdin, A.M. Kadomtseva, and R.Z. Levitin: Orient-
atsionnye Perekhody v Redkozemel'nykh Magnetikakh (Orientational Tran-
sitions in Rare Earth Magnetics), Nauka (1979).
DIAGRAMS IN THERMODYNAMIC PARAMETER SPACE 187
23. L. Neel: Izv. Akad. Nauk SSSR, Ser. Fiz. 21, 890 (1957).
24. S. Foner: Phys. Rev., 130, 183 (1963).
25. G.K. Chepurnykh: Fiz. Tverd. Tela 10, 1917 (1968).
26. I.E. Dikshtein, V.V. Tarasenko, and V.G. Shavrov: Zh. Eksp. Tear. Fiz.
67,816 (1974).
27. V.A. Lvov, and D.A. Yablonskii: Fiz. Niz. Temp. 8, 951 (1982).
28. V.G. Bar'yakhtar, A.N. Bogdanov, V.T. Telepa, and D.A. Yablonskii: Fiz.
Tverd. Tela 26, 389 (1974).
CHAPTER 6
188
MACROSCOPIC ORDER PARAMETERS 189
Table 6.1
List of tensors frequently used in phase transition theory [2]
DV Magnetization Md
Strain f. 0I{3
Susceptibility Xo:{3
where A is a complete tensor of specified rank with all the components being
non-zero. As an example, we wish to find the symmetric form of the strain
tensor for the group C2h(~)' This group consists of the elements
where the brackets specify the transformation of the radius vector r = (xyz)
n
components for each element. The formula (21.1) then gives
:::) + ( :::
fzz -fzx
fxy
fyy
- f zy
=:::)]
fzz
= (:::
0
:::
0
(21.2)
Suppose now that an Oh(m3m) -+ C2h(2/m) transition takes place. The strain
tensor, which is symmetric under the group Oh, has the form: (f)Oh = hxx,
where I is the identity matrix. On comparing (f)Oh and (f)c2h (21.2), we see
that the Oh -+ C2h transition should be accompanied by the occurrence of
spontaneous deformation f xy .
(21.3)
Using this formula and tables of space group IR's for the wave-vector", =
o (these IR's coincide with the representations of the corresponding point
groups), we can readily find the constitution of the tensor representation, for
example, for the group Oh:
(21.4)
192 CHAPTER 6
Indicated in the brackets are the frequently used notations of the point group
representations. For the representations Tl, 1"5, and T7 the tensorial basis func-
tions can now be found by the projection operator method. Proceeding from
the general formula (21.1), we write the expression for the basis functions ¢'5.
transforming by the IR of the point group Co:
(21.5)
(21.6)
Similarly, we obtain the basis functions for the two-dimensional Ts(Eg) and
three-dimensional T7(T2g ) IR's of the group Oh:
A complete list of basis functions constructed from polar and axial vector com-
ponents and second-order tensor components is presented in Table 6.2.
Table 6.2
Macroscopic basis functions constituted by polar (r) and axial (S)
vector components and by second-rank tensor for point group IR [5].
C 2h T2(Au) z. (Eg)
T6(B 2u ) y. (fxz,fyz )
S4 T2 + T4 (x,y),(Sy,Sx),
(E1 + E 2) (fyz,fxz).
T3(B) Z, fxx - fyy, fxy
194 CHAPTER 6
and construct a potential that is invariant under the group Oh. In this case
the straightforward method is the more convenient. We construct the various
quadratic combinations from the variables f;:
6
<1> = L rit:; + L rijf;fj. (21.8)
;=1 ;j
On operating on <1> with the generators I, C4z , C3 (111) of the group Oh and
requiring the invariance of <1>, we obtain
(21.11)
We consider the scheme of constructing potentials when there are two param-
eters (in the general case these are multi component parameters). Assume that
one of these parameters is microscopic and its transformation properties are de-
scribed by space group irreducible representations, whereas the other parameter
is macroscopic and its transformation properties are described by point group
irreducible representations.
In constructing the potential, we may encounter two cases, namely: the
micro- and macroparameters may have similar and different transformation
properties. We label the microscopic OP components by 1}).., and the macro-
scopic OP components by Xa or Xi (Xa = Pa , M a , Xi = fi)' The thermo-
dynamic potential will contain two terms quadratic in the parameters 1}).. and
Xa: Tl L).. 1}~ and l{ La x~. In keeping with Landau's theory (§1) it will be
assumed that only Tl and l{ depend on temperature: Tl = T10(T - Tel), l{ =
l{o(T - Te2)' The form of the other terms involved in the thermodynamic
potential is determined by the form of the invariants contained in the corre-
sponding IRBI. To find this IRBI, we need to construct the reducible represen-
tation D"Y EB D"=o according to which the variables 1}).. and Xa transform and
to proceed henceforth in keeping with the recommendations outlined in §12.
The thermodynamic potential obtained as a result of the relevant computations
198 CHAPTER 6
may be represented as
(22.1 )
or in simplified form:
(22.5)
Here gl (1]) and g2 (1]) are polynomials relative to the components of the para-
meter 1], and since 1] should not transform according to the identity represen-
tation, this would imply the absence of a transition; the first term involved in
the polynomial g2(1]) is quadratic in 1].
As a rule, the outlined procedure of constructing the IRBI from the vari-
ables 1]>. and x'" is somewhat laborious to realize in practice. On the other
hand, in most of the cases one can restrict oneself to the first terms in <l>int, so
a simpler approach may be used in the construction. The gist of this approach
is as follows. It is necessary to find the least degree n such that the decomposi-
tion of the representation (DI<v)n into irreducible representations contains the
representation DI<=o. Then the terms describing the coupling of the parameters
1]>. and x'" in the potential <l>int will have the form 1]).. X",.
IMPROPER TRANSITIONS
The presence of a coupling between the parameters 1]>. and X", is responsible for
the occurrence below the phase transition point of new macroscopic properties,
which are described by quantities x",. Such transitions are normally referred
to as improper (§27). In recent years these transitions have been investigated
extensively in connection with their utilization in instrument engineering. Our
goal is to determine the form of the simplest potential describing an improper
transition. We start by considering the case when the quantities 1]>. and x'" are
one-component micro- and macroparameters. Depending on the transforma-
tion properties of these parameters, their coupling may be described by terms
of the form 1]X, 1]2X, 1]X2, 1]2x2. An interplay of the form 1]X occurs when the
parameters 1] and x transform in the same way. Such transitions are said to be
MACROSCOPIC ORDER PARAMETERS 199
where the parameters 1J>. describe the transition at the point M and the para-
meters 6 the transition at the point R [7]. In the examples considered, the
first structural transition in CsBrCla and CsPbCl3 is a first-order transition,
so it is necessary in the potentials (22.6) and (22.7) to allow for the sixth-
order terms in the parameters 'fJ>. and 6. In the following we analyze only
a particular approach rather than the entire phase diagram and therefore, to
simplify calculations, not all of the invariants of degree six will be taken into
account in the expansions (22.6)-(22.8). Having the potential (22.6)-(22.8),
one can study interaction effects of two structural transitions. We concentrate,
however, on the effects of the interplay of the micro- and macroparameters in
the vicinity of one of the structural transitions.
An important role in the investigation of structural transitions attaches to
ultrasonic measurements since analysis of ultrasound velocity anomalies pro-
vides information on the behavior of the OP's. For such experiments to be
described theoretically, it is necessary to add to the potentials (22.6)-(22.8)
the terms specifying the elastic energy of the crystal:
and the interaction energy of the strain tensor with the OP [7]:
Xi being the standard notation of the strain tensor components (§21). The
expression (22.10) is easy to construct with the help of the straightforward
method described in §11.
Consider the first structural transition described by the modes of the point
M ({,d I}). The symmetry group of the dissymmetric phase, according to [7],
is D~h' These data suffice to determine the solution type (0 0 1J) corresponding
to the 0]; ---+ D!h transition. Interested in this particular transition only, we
write by virtue of the formulas (22.6)-(22.10) the effective potential
(22.11)
MACROSCOPIC ORDER PARAMETERS 201
(22.12)
( u' )2
T. - TO
c2 - c2
+ 4_1_'7 _
' (22.14)
r20 v'7
,
il'TJ = 'TJ. IT=T
U1'7
c2 = -2;"". (22.15)
'7
Using these results, we can readily calculate the anomalies of the elastic moduli
at the transition point by the formula [8]
(22.16)
~ _ ( 02~eff )
(22.17)
)../-1 - O'TJ)..O'TJ/-I •.
As a result, we can write the values of the elastic moduli above the tran-
sition temperature (T > TC2):
Cij = C~ (22.18)
23. Ferroics
The previous chapters were concerned with different aspects of the symmet-
ric and thermodynamic description of phase transitions, based on an analysis
of the transformation properties of the macroscopic OP's. In particular, we
have ascertained (§9) that the crystal symmetry lowering at a phase transition
leads to the formation of a domain structure from which rotational, arm, and
translational domains may be singled out. The presence of rotational and arm
domains is due to a decrease in rotational crystal symmetry and the presence
of translational domains is associated with a change in translational symmetry.
Correspondingly, all the phase transitions may be broken up into two types.
The first type includes transitions in which the dissymmetric-phase point group
G~ does not coincide with the initial-phase point group GO (G~ C GO). The
second type comprises transitions for which the point groups of the initial and
dissymmetric phases coincide (G~ = GO). Crystals that undergo transitions
of the first type are often said to be ferroics [9]. Because of the lowering of
the point symmetry, the dissymmetric phase in ferroics should be described by
macroscopic variables, alongside the microscopic OP's. For example, if there
occurs in the crystal a phase transition with tl, = 0 from an initial cubic-
symmetry phase to a tetragonal dissymmetric phase, then the dissymmetric
phase should consist of three domains. Each domain differs from the other in
the strain tensor describing the spontaneous deformation of the parent cubic
cell along one of the edges x, y, z. The physical nature of the spontaneous mi-
croscopic variable arising at the phase transition underlies the classification of
crystals that undergo phase transitions. Thus materials in which a spontaneous
polarization or magnetization vector occurs at a phase transition are called fer-
roelectrics or ferromagnetics. A crystal in which the dissymmetric phase is
described by the spontaneous strain tensor is called a ferroelastic. Similarly,
other types of ferroics are distinguished. A crystal classification based on these
principles is presented in Table 6.3. To determine to which item of Table 6.3 a
particular phase transition GO -+ G~ belongs, we need to find the point group
GO representation D V describing the symmetry change GO -+ G~ and then
to calculate the basis functions of this representation from the material tensor
components. If the basis functions can be constructed from the polarization
vector components, we have a ferroelectric transition. But if the non-zero basis
MACROSCOPIC ORDER PARAMETERS 203
Table 6.3
Classification of ferroics [10]
lVlacrgv.anables
Type of ferroic shecI ylllg DO,I)1~in Examples
t e domams SWItC over field
Spontaneous
Eerro~
electrIC polarization ~~ratric BaTi0 3
FERRO ELECTRICS
This class of ferroics has received the most study. As is seen from Table 6.3, the
macroscopic variable describing the ferroelectric transitions is a polar vector,
that is, a first-rank tensor. Since a vector transforms according to a represen-
tation of dimension not higher than three, every proper ferroelectric transition
should be described by one of the twelve thermodynamic potential types cor-
responding to the crystallographic point I groups (§14). The IRBI's of these
groups are outlined in §14.
In the case of improper ferroelectrics the procedure of constructing the
potential becomes somewhat more complicated. Namely, it is required, by an-
alyzing the relation of the space groups of the initial and final phases, to find
the relevant IR DV. Then, by analyzing the point symmetry change at the
transition, we need to find the representation D"=o describing the transfor-
mation properties of the polarization vector. The IRBI should be constructed
according to the technique (outlined in §12) for the representation D V EB D"=o.
A detailed analysis of the improper ferroelectrics which are not simultaneously
ferroelastics is contained in [11]. It has turned out that transitions of this kind
are described by three potentials:
(23.1)
where
<1>2 = r(7]r + 7]5 + 7]5) + Ul(7]r + 7]5 + 7]5)2 + u2(7]r7]5 + 7]r7]5 + 7]57]5)+
+KP2 + (8P)7]17]27]3; (23.2)
MACROSCOPIC ORDER PARAMETERS 205
(23.3)
FERROELASTICS
Table 6.4
Examples of compounds undergoing a ferroelastic transition
Proper
V3Si O~ D~h 21
InTI O~ DI7 320
4h
KN0 2 0 9h Dg d 295
TbV0 4 DI9 D24 33
4h 2h
Improper
HIGHER-ORDER FERROICS
Table 6.5
Point symmetry change types for higher-order ferroics [13]
Notation: diJj , piezoelectric tensor; C)..Jj' elasticity coefficient; L)..Jjv,
polar sixth-rank tensor
Spontaneous tensor
Ferroic type Symmetry change components
Ferroelasto-
electric mmm 222 d 14 , d25 , d36
4/mmm 422 d 14
3m 32 d 14 , dl1
6/m 6 d l1 , d22
Dvx[DvF 6m2 6 d l1
6mm 6 d 14
6/mmm 6 d l1 , d22
6/mmm 6m2 d 22
6/mmm 622 d 14
m3 23 d 14
432 23 d 14
m3m 43m d 14
m3m 23 d 14
[[D~F] 3m 3 C 15
6/m 3 C 14 , C25
6/mmm 3m C 14
6/mmm 3 C 14 , C 25
208 CHAPTER 6
Spontaneous tensor
Ferroic type Symmetry change components
Ferroelasto-
electric and
ferrobielastic 42m 4 d 15 , d31 , C 16
4mm 4 d14 , C 16
r
4/mmm 4 d 14 , d15 , d31 , d36 , C 16
6 3 dll,d22,CI4,C25
6mm 3m dll, C 14
quantities of rank three and higher than three belong to higher-order ferroics.
According to the nature of the spontaneous third- and higher-rank quanti-
ties, the higher-order ferroics are subdivided into ferroelastoelectrics, which
are characterized by the occurrence of spontaneous components of the piezo-
electric tensor dip, ferrobielastics, which are characterized by the occurrence of
spontaneous components of the elasticity tensor CAP' and higher -order ferroics
proper.
Just as with ferroelectrics and ferroelastics, we can enumerate all types
of point symmetry change for higher-order ferroics (Table 6.5). The left-hand
column of Table 6.5 gives the name of a ferroic and shows the physical repre-
sentation according to which the corresponding tensor components transform.
MACROSCOPIC ORDER PARAMETERS 209
R3 at T = 648K is LaCo03.
Table 6.6 presents materials that, by virtue of their structural data, may
be higher-order ferroics [13]. The principle of selecting these materials is based
on the fact that the structure of each compound may be viewed as a weakly
distorted version of a more symmetric hypothetical structure Go. The ratio
of the point groups GO and Gg should not contradict the data of Table 6.5.
The principles of searching for the groups Go are the same as those used in the
selection of the paraphase (§1O).
210 CHAPTER 6
Table 6.6
List of compounds in which one may expect phase transitions accom-
panied by spontaneous occurrence of high-rank tensors (r 2:: 3) [13]
Dissymmetric Transition
Type Compound Paraphase phase type
N2 Pa3 P2 1 3 P
KAlO 2 Fd3m P2 1 3 I
24. Non-ferroics
As stated in the foregoing, the very fact that a spontaneous macroscopic quan-
tity arises at the phase transition may, irrespective of the physical content of
this quantity, provide a basis for a classification of phase transitions. Thus the
concept of the ferroic was introduced to denote a vast class of crystalline ma-
terials in which the phase transition may be accompanied by the occurrence of
spontaneous macroscopic variables. The rest of the materials should evidently
belong to the opposite class of 'non-ferroics'.
A major group-theoretic indication of ferroics is that the point symme-
try of the crystal should necessarily change at the phase transition, as it is
the symmetry of the final and initial phases that determines the nature of
the spontaneous macroscopic quantities. Major group-theoretic indications of
non-ferroics are the obligatory preservation of the crystal point symmetry at
the phase transition and the obligatory change of the translational symmetry.
Along with the term 'non-ferroic', introduced in [16], we will therefore speak
also of a translational phase transition.
The various translational phase transitions have actually been enumerated
in tables of subgroups of space groups [17]. From these tables, it is seen that
most of the possible translational transitions are described by one-arm stars.
In the case of many-arm stars the translational transition is described by a
full-arm channel.
To describe exhaustively all potential non-ferroics, it is necessary to find
for each translational transition the relevant IR and its image (I group), using
tables [17], and to construct the IRBI. This program has been fully implemented
in [16], but only for the case of second-order translational phase transitions.
As follows from the results of that paper, the translational transitions are
comparable in number with so extensive a class of ferroics as the ferroelastics
(§23). It turns out that of the 230 space groups, 167 may be taken as the
symmetry group of the initial phase of non-ferroics. For these 167 groups a
total of 1800 possible translational transitions, described by 1713 active IR's
of sixty-five Brillouin zone points, have been revealed. Of the 1713 active
representations, 1191 IR's are one-dimensional and belong to the case of a
one-arm star. The phase transitions in these cases should be accompanied by
doubling of the primitive cell. For two-arm stars of thirteen Brillouin zone
points, 330 active IR's have been found. The corresponding transitions should
212 CHAPTER 6
EXAMPLES OF NON-FERROICS
References
12. P. Toledano, and J .C. Toledano: Phys. Rev. B 21, 1139 (1980).
13. P. Toledano, and J .C. Toledano: Phys. Rev. B 16, 386 (1977).
14. L.D. Landau, and E.M. Lifschitz: Statistical Physics, Addison-Wesley (1958).
15. R. Bline, and B. Zeks: Soft Modes in Ferroelectrics and Anti-ferroelectrics,
Elsevier-North Holland (1974).
16. P. Toledano, and J .C. Toledano: Phys. Rev. B 25, 1946 (1982).
17. V.E. Naish, S.B. Petrov, and V.N. Syromyatnikov: Podgruppy Prostranst-
vennykh Grupp. II. Podgruppy s Ismeneniyem Yacheiki (Subgroups of Space
Groups II. Subgroups with Alteration of Cells), Dep. VINITI, No. 486-77
(1977).
18. A. Chattergee, A.K. Singh, and A. Jayaraman: Phys. Rev B 6,2285 (1972).
19. A. Jayaraman, A.K. Singh, A. Chattergee, and S. Ushe Devi: Phys. Rev.
B, 9 2513 (1974).
20. M. Hidaka, LG. Wood, B.M.R. Wanklyn, and B.J. Gerrard: J. Phys. C
12, 1799 (1979).
21. M. Hidaka, LG. Wood, F.R. Wondre: J. Phys. C 12,4179 (1979).
CHAPTER 7
215
216 CHAPTER 7
(25.4)
The generalized variable x in this representation of the potential is described
by the expression
(25.5)
X
-1
=
({)X)
{)xs
({)2<1»
= ()x 2 s·
(25.8)
The symbol s here indicates that the corresponding quantity is taken for X = O.
Analysis of the x versus X curves sketched in Fig. 7.1 enables us to make
a plot of the potential <I> versus the field X at T < To. This relation is shown
in Fig. 7.3.
Examination of the x(X) and <I>(X) relations shows that as the field X is
varied, the OP x and the energy of the system should exhibit discontinuities
between the states corresponding to the points 4-8 and 2-6. A hysteresis loop
2-3-4-8-7-6 should be observed in experiment. The coercive field is equal to
HX6 - X4)'
PHASE TRANSITIONS IN AN EXTERNAL FIELD 217
,x
T>To
..X
By analyzing the x(X) (Fig. 7.1) and <I>(X) (Fig. 7.3) plots, one can con-
struct a phase diagram such as that depicted in Fig. 7.4. As can be seen from
Fig. 7.4, the kind of the phase transition changes at the point 0 (tricritical
point).
9 82
Fig. 7.3. Field X dependence of the minimized potential <I> at T < To.
Fig. 7.4. Phase diagram for the system with the potential (25.6). The
dashed line refers to second-order phase transitions; the solid line
refers to first-order phase transitions; the dot-and-dash line refers to
the stability boundary of the phases.
In describing such transitions, one allows for the invariants of degree six in-
volved in the potential and assumes that the coefficient of the fourth-order
invariant is negative (§ 16 and also [1,2]). Consider the effect of an external
field on these transitions, proceeding from the expression for the potential
(25.9)
enables one to make plots of the OP x versus the temperature r = rh(T - To)
and the field X, which are presented in Figs. 7.5 and 7.6. Analysis of the x(X)
PHASE TRANSITIONS IN AN EXTERNAL FIELD 219
x<x ,~
The second interval is given by the inequalities Tc < T < T cr , with Tc the
temperature of a first-order phase transition in the absence of a field (§16). In
220 CHAPTER 7
3u 2
rcr = 5v ' (25.13)
(25.14)
(25.15)
_ x
X=-,
Xcr
_ r
r=-,
-
X=-.
X (25.17)
rcr Xcr
Fig. 7.7. Phase diagram in the variables (T, X) [3] for the 1]6 model.
The solid lines refer to the first-order transition; the dot-and-dash
lines refer to the boundaries of the regions where the phases coexist
(see Table 7.1).
Table 7.1
List of regions of existence of different phases on the phase diagram
of the 1]6 model (Fig. 7.7)
Initial
Region phase Phases Phases
BDG S M -
CDG M S -
BGF S M M
AFGC M S M
ACE - S M
the potential <I> for each value of the parameter ii:, we can locate boundaries of
existence of the initial phase, boundaries of coexistence of several phases, and
phase transition points [3]. It is in this way that the authors of [3] found the
regions of coexistence of phases (phase stability overlap regions), delineated
by a dot-and-dash line in Fig. 7.7, and the first-order phase transition line,
222 CHAPTER 7
shown by a thick solid line. The first-order phase transition line terminates at
a critical point D with coordinates r = 1, X = 1 [3].
With u > 0 the critical values determined from the conditions (25.11) are
1'cr = Xcr = Xcr = O. This point is normally referred to as a central critical
point [3]. Critical points in the case of u < 0 (D-type points), the values of
which are given by the expressions (25.12)-(25.14), will be called critical points
of phase diagram wings. Assuming the parameters l' and u of the potential
(25.9) to depend on T and the external field X ext (for example pressure P), we
can construct a phase diagram in the space of T, X, X ext , where X is the field
conjugate to the OP (Fig. 7.8) [3]. The diagram portrayed in Fig. 7.7 may be
viewed as one of the cross-sections X ext =const of the diagram in Fig. 7.8. The
point t (Fig. 7.8) at which three second-order phase transition lines converge
is said to be a tricritical point. The Ft line and its continuation lie in the
(T, X ext ) plane; the surface spanning the Ft and tD lines and the surface
symmetric relative to it are the surface of first-order phase transitions.
"- '\
A \
\
\ Xed
Fig. 7.8. Phase diagram in the space of the parameters T,X, and
X ext for the 1]6 model. The dashed line represents a second-order
phase transition; the solid line shows the phase diagram section cor-
responding to Fig. 7.7.
In the (1', X ext ) plane the point t can be obtained from the condition
1'(T, X ext ) = u(T, Xexd = 0 [2].
(25.18)
(25.19)
(25.20)
These three phases have the same energies, as can be readily verified by sub-
stituting the corresponding values Of'T}l and 'T}2 into equation (25.19), and cor-
respond to three different domains of the same phase, for example ('T}, 0). An
applied pressure should, however, lift this degeneracy. Indeed, we add to the
expression (25.19) the term
(25.21 )
where 0"1 and 0"2 are external forces conjugate to the OP. These quantities
are constructed on the basis of the expressions (25.19) from the strain tensor
components 0"0:(3:
(25.22)
224 CHAPTER 7
and a phase of the ('fJ1 'fJ2) type is determined from the system of equations of
state &<P/&'fJ1 = 0 and &<P/&'fJ2 = 0, which can be rewritten in the form
0"1 0"1
r - 3V'fJ1 + 2u(4'fJ12 - v
2 2
-3 ) = 0, 'fJ2 = 3'fJ1 - - .
3v
(25.24)
A numerical analysis of equations (25.23) and (25.24) was made in [5]; the
result is presented in Fig. 7.9. On the (0", T) plane there exists a point t1 at
which the second-order transition from phase ('fJ, 0) to phase ('fJ1, 'fJ2) changes to
a first-order transition. The point t1 therefore is a tricritical point. The phase
diagram on the (0", T) plane is shown in Fig. 7.10.
''It
(a) (b)
The phase diagram given in Fig. 7.10 (uniaxial stress 0"2 = 0) may be
regarded as the cross-section of a more general phase diagram (Fig. 7.11) cor-
responding to arbitrary stresses 0"1 =1= 0 and 0"2 =1= 0 [6]. The diagram presented
in Fig. 7.11 may be viewed as consisting of three phase diagrams such as those
sketched in Fig. 7.10, which are located at an angle of 1200 relative to each
other. Such diagram symmetry is a consequence of the fact that the initial
PHASE TRANSITIONS IN AN EXTERNAL FIELD 225
(5
0
~'I'tO)
it'
(If,'I'2) .\
T
Fig. 7.11. Phase diagram in the case of spontaneous strain in A-15 com-
pounds. q, quadrupolar point; tl, t2, and t3, tricritical points.
chapter.
the form
(25.26)
The effective potential contains two quadratic invariants, r1 'f}f and r2'f}~. This
implies that under the effect of the external field X 2 the two-dimensional rel-
evant IR has split up into two one-dimensional representations, as a result of
the lowering of the 'crystal plus field' system's symmetry. Thus we have passed
on to the problem of two coupling OP's (treated in detail in §20). At certain
values of the field X 2, as well as of the other coefficients of the potential (25.26),
a phase diagram with a tetracritical point (Fig. 5.1) may be realized. As a
result, as the temperature is lowered a ('f}1, 0) or (0, 'f}2) type phase should
arise first; it is only when the temperature is decreased further that a ('f}1, 'f}2)
PHASE TRANSITIONS IN AN EXTERNAL FIELD 227
form:
<P = <Po(T) + <PI) + <Px + <Pint, (26.1)
where 'f/ == {'f/1,'f/2, .. .}, X == {X 1,X2, ... } is the contracted notation of the OP
and of the generalized force acting on the system. The starting point in our
analysis is the equation of state for the OP (o<P/O'f/ = 0)
(26.5)
(26.6)
(26.7)
where the symbol s signals the fact that instead of 'f/ we should substitute a
value of 'f/. that satisfies the equation of state. We transform the second term
in equation (26.7) by representing it first as
The relation between the variables 'f/).. and X is given by the equation of state
(26.5), which will be written in the form 1/;('f/, X) = O. From this it follows
immediately that
01/; + 2:(
01/;) (O'f/)..) = O. (26.9)
oX ).. O'f/).. oX
Substitution of the explicit expression for 1/;( 'f/, X) into this equation yields the
epxression for the quantity (O'f/)../ oX).
(26.10)
PHASE TRANSITIONS IN AN EXTERNAL FIELD 229
(26.11)
Inserting equations (26.11) and (26.8) into equation (26.7) gives the final ex-
pression for the generalized susceptibilty x:
(26.12).
(26.13)
When substituting spontaneous values of the parameter 1], part of the terms
involved in the polynomial gl (1].) may cancel out in the case of a multi compo-
nent OP. We introduce the index nF, which is equal to the lowest degree of the
o P components contained in the polynomial gl (1].). Following Aizu, we call
these indices faintness indices. If, by virtue of the symmetry of a particular
problem, the polynomial gl(1]) == 0, then, according to Aizu, we will assume
that nF = 00 and nM = 00.
Since the classification of susceptibilities under consideration relates to
second-order phase transitions, it will be assumed, as usual, that only one
coefficient in the potential (26.1)-(26.4) depends on temperature: r = r~(To -
T). Then, evidently, the following relations hold:
(26.15)
230 CHAPTER 7
where the <Pi ('fl.) are the eigenvalues of the matrix 11> AI' ('fl.). Again following
Aizu [9], the indices Pi will be called catastrophe indices. Finally, as shown in
§25, the lowest degree of the 'fli components involved in g2('fl) is equal to two,
so we may set
(26.16)
in the vicinity of the phase transition. We now analyze the temperature depen-
dence of the second term in the generalized susceptibility (26.12). Allowance
for the relations (26.14), (26.15), and (26.7) yields
For the indices P = 1 equation (26.19) coincides with the Curie-Weiss law.
We now let T be less than To. The expressions (26.12), (26.16) and (26.17)
then yield
X - L '" (To - T)-l, (nM - P = 0), (26.20)
that is, the Curie-Weiss law holds in the dissymmetric phase. Note that,
according to the relation (26.19), the Curie-Weiss law also holds in the initial
phase if nM = 1. For nM > 1 the susceptibility in the initial phase, according
to the relation (26.17), does not depend on temperature. From the expressions
(26.16) and (26.17), it follows also that
CATASTROPHE INDICES
We consider possible values of the catastrophe indices Pi. In the initial phase
(T > To) <I>'xI'(1].) = </>8,X1' and Pi = 1 for any i in view of the fact that 1]. = 0,
so the temperature singularities of the generalized susceptibility at T -+ To + °
are determined only by the value of the faintness index nM and do not depend
on the catastrophe indices. It can be readily verified that in the dissymmetric
phase T < To near the second-order phase transition point (T -+ To - 0), in
the case when the phase transition is described by a one-dimensional IR, the
catastrophe index is equal to unity: P = 1. Thus catastrophe indices with values
other than unity may occur only for phase transitions with a multicomponent
OP.
As an example, we wish to find the catastrophe indices for a 2-component
OP 1] = {1111]2} whose transformation properties correspond to the potential
;r.
'ol' = r (2
1]1 + 1]22) + U (2
1]1 + 1]22)2 + V11]12( 1]12 - 31722)2 + V21]22( 1]22 - 31]12)2 . (26.22)
The matrix <I>,XI' for a solution of type 1]1 = 1], 1]2 = °has the form
(26.23)
(26.24)
It might be as well to point out that the values for the catastrophe indices
depend on the choice of the variables. Indeed, let the potential <I> have the form
(26.25)
(26.26)
(26.27)
(26.29)
(27.1)
(27.2)
(27.3)
The condition for the dissymmetric phase to be stable, 0 2 <I> /0"1 2 > 0,
reduces to the requirement that the inverse susceptibility be positive:
(27.4)
Substituting the spontaneous parameter "Is (that is, the equilibrium value
of the parameter "I in a zero field) into equation (27.3), it is easy to find the
isothermal inverse susceptibility (T < To):
Above the transition temperature (T > To) we have in the initial phase
(where TJ = 0)
(27.6)
The resulting expressions lead to Landau's 'law of 2' [12] which means that
at the same distance from the transition temperature the susceptibility in the
symmetric phase is twice as large as the counterpart in the dissymmetric phase.
The temperature dependence of the susceptibility X satisfies the Curie-Weiss
law on either side of the transition.
(27.7)
(27.8)
In the absence of a field, the spontaneous values of the parameters TJs and Xs
(27.9)
2 r
TJ s = - 2u' (27.10)
whence it follows that the variable Xs depends on temperature in a linear
fashion (Fig. 7.12).
We determine the catastrophe index p, that is, the condition for the dis-
symmetric phase to lose stability in a zero field X:
0 2 If>
- = -41'~(T - To). (27.11)
0172
It follows from the expression (27.11) and the definition of the index p (§26),
that p = 1.
PHASE TRANSITIONS IN AN EXTERNAL FIELD 235
Thus if the transformation properties of the parameters 1] and x are such that
a coupling of type 1] 2 X occurs, then the susceptibility X with respect to the
quantity x in terms of the model (27.7) undergoes a discontinuous change at
the transition point:
.x
We now introduce the field XT/ conjugate to the parameter 1] and calculate
the susceptibility XT/' The field XT/ is defined by the equation
XT/ = (~<l»
u1] x,T
= 2r1] + 4U1]3 + 201]x, (27.14)
(27.15)
236 CHAPTER 7
Above the transition point (T > To), "1. = 0, x. = 0 and the susceptibility
conforms to the Curie-Weiss law
In the dissymmetric phase (T < To) the Curie-Weiss law works well too, but
the 'law of 2' no longer holds:
2
XIj-1 = 4roI ( To - T ) ( 6 u) .
1 - 8f{ (27.17)
From the results obtained, we thus see that the susceptibility corresponding
to the parameter "1, which is viewed as a primary parameter (§21), diverges at
the phase transition point To, whereas the susceptibility corresponding to the
secondary or improper parameter x remains finite within the framework of the
model (27.7).
An interaction of the type "12X (or "12 X), which contains the potential
(27.7) considered above, arises in those cases where the symmetry of the initial
phase allows a linear invariant x. Such a situation may be encountered, for
example, in cubic crystals when account is taken of the interplay of the strain
(x = ta/3) with the OP "1.
Of all the cases that may be described by the potential (27.7), the case
of translational transitions is the most interesting. As already stated in §24, a
special feature of translational transitions is that the symmetry change at the
phase transition point arises only from a change in the translational symmetry
of the crystal while the point symmetry remains unaltered. As a consequence,
such transitions are not accompanied by the occurrence of new macroscopic
spontaneous variables x, so the temperature anomalies in the macroscopic vari-
ables x, allowed by symmetry in each particular case, as well as the generalized
susceptibilities X are important sources of information in the investigation of
translational transitions.
In the foregoing example it was assumed that the generalized thermo-
dynamic variable x was an independent variable in the representation of the
potential (27.7). This way of representing the potential is convenient in finding
inverse susceptibilities, as can be seen from the above computations. In §25
we noted that in some cases it is more convenient to choose generalized forces
X conjugate to x as independent variables. As will be clear from the example
that follows, such representation of the potential <I> is natural in the calculation
of the susceptibilities X.
PHASE TRANSITIONS IN AN EXTERNAL FIELD 237
(27.18)
with u > 0 and v > o. An interplay of the type 7]2 X 2 corresponds to the
indices nM = nF = 00. The catastrophe index, just as in the previous example
(potential (27.7)), is equal to unity.
The formula (26.6) says that in the present case the quantity Xs for the
dissymmetric phase is equal to zero. The susceptibility is given by the expres-
SIon
(27.19)
where g2(7]s) = ,7];. On substituting in equation (27.19) the quantity 7]., which
is determinable from the equation of state, we find the temperature dependence
of the susceptibility X [13]:
u, [(
X = L -2 3v 1 - ,
1 - 3vro(T - To)u -2)+t] . (27.20)
"r >0
L ----)-
/ Iy<o
,/ I
I T
Tc '
An 7]2 x 2 type coupling exists in all the cases where the linear invariant
x is absent. As in the previous example, the results obtained are particularly
important in the case of translational transitions.
In the foregoing we have considered two models of potentials, (27.7) and
(27.18), which describe improper transitions with a one-component OP 7] and
one parameter x, and also improper transitions with two interaction types,
7] 2 X and 7] 2 X 2 respectively. These two models may be employed to describe
improper ferroelastics. However, as shown in [14], they are inapplicable in the
238 CHAPTER 7
iF. _
'!' - r (2
'rJl + 'rJ22) + Ul (4 4) 2 2 LX- 2 + 'Y'rJl'rJ2 X ,
'rJl + 'rJ2 + U2'rJl 'rJ2 + (27.21)
where the quantity X coincides with one of the components of the electric field
E.
In the absence of a field the equation of state allows three solutions:
O. 'rJls = 'rJ2s = 0;
2 _ r.
1. 'rJls = 0, 'rJ28 - - 2Ul ' (27.22)
We can readily see that it is only in phase 2 that a spontaneous quantity may
occur alongside non-zero parameters 'rJls and 'rJ2s:
Xs = 'Y'rJls'rJ2s' (27.23)
X = L (T> To);
(27.24)
X = L + 'Y2(4ul + U2)-1 (T < To).
For T < To the susceptibilities are independent of temperature.
(27.25)
PHASE TRANSITIONS IN AN EXTERNAL FIELD 239
X.=,'f]s· (27.26)
The susceptibilities X above and below the transition point obey the Curie-
Weiss law:
,2
X= L + - (T> To);
2r
,2 (27.27)
X = L - 4r (T < To).
The phase transitions described in terms of the model (27.25) are said to be
pseudo-improper [1]. As an example, we may quote the ferroelastic transition
mmm --+ 21m which occurs at T = -62°C in (NH 4hS04 [15].
As already stated in §16, as a simple model of the potential <I> describing first-
order phase transitions we may take a potential of the form
(28.1 )
where u < 0 and v > O. The equation of state allows two solutions (see
equations (16.5) and (16.6)):
2 U[ ( 3Vr)1/2]
'f]2 = - 3v 1+ 1- U2 . (28.3)
In §16 the first solution was shown to be an analytic continuation of the unique
solution in the 'f]4 (v = 0) model and to describe a second-order transition. The
second solution (28.3) is absent in the 174 model and occurs for the first time
in the 'f]6 model. This solution describes a first-order transition. Evidently, as
invariants of higher degrees are added the solutions to the equation of state will
increase in number and part of the solutions will be an analytic continuation
of the preceding models, but there will also be 'novel' solutions. It stands to
reason that all the solutions can be classified according to types, laying at the
240 CHAPTER 7
(28.4)
y3 _ 3py - 2q = 0, (28.5)
where
(28.6)
The equation (28.5) has three solutions if p3 - q2 > 0. These may be written
as
v v ( 8uw) 1/2 1
1]
2
sn
= -4w
- - -4w 1- - -
3v 2
cos -(a + 2mr),
3
(28.7)
sin a = 4uw (1 _ 8UW) -3/2 (1 _ 3rv _ 32uw + 12rw _ 12r2w2) 1/2, (28.8)
v'3v2 3v 2 u2 9v 2 uv u 2v 2
Exploiting the definition of the transition type (§16), we find that the solution
with n = 1 is a 1 - s type solution and describes a second-order transition; the
solution with n = 2 is a 2 - s type solution, while the solution with n = 3 is
PHASE TRANSITIONS IN AN EXTERNAL FIELD 241
(28.10)
2 V [1- ( 1- -
= -3w
- 3UW) 1/2] , (28.11)
1] 2
s, v 2-
2
= --
1] 0
V [1 + ( 1- -3uw) 1/2] . (28.12)
" 3w v2
Substitution of equation (28.11) or (28.12) into the equation <I> = 0 yields the
value of r at the point T = Tc:
__ 2v 3 [ _ 9uw ( _ 3UW)3/2]
rc - 27 w 2 1 2v 2 =f 1 v2 •
(28.13)
-1 _ _
XO -
4v 3 [ _ 9uw
27w 2 1 2v 2 =f
(1- 3UW)3/2]
v 2 •
(28.14)
-1 8v3{){
= 9w )2 , (28.15)
Xd 2 ±R 1 =f R
-1 2v 3 2
Xo = 27w 2 (I ± 2R){1 =f R) , (28.16)
We begin with the 2 - s type transition. The expression for R leads to the
inequality
(28.18)
(28.19)
On the <I> = 0 transition line the inequality for 2 - sand 3 - s type transitions
takes the form
32v 3 R(l _ R) > 0 (28.20)
9w 2 -
From the requirement that 1J;,2 > 0 at T = Tc and from the conditions (28.18)-
(28.21), it then follows that a 2 - s type transition is possible when u < 0 and
v > O. For a 3 - s type transition the region of existence is bounded by the
inequality (28.18) and v < O.
We consider in more detail the case v < o. The temperature of a phase
transition of 1- s type (second-order transition) is determined by the condition
'r = o. The 3-s transition temperature is described by theequation (28.13), in
which the plus sign should be retained. From this, it follows immediately that
if the bracketed expression in equation (28.13) is greater than zero at v < 0,
then the temperature Tc for a 3 - s transition is higher than the temperature
To for a 1 - s transition and vice versa. As a result, we may write down the
following constraints on the coefficients u, v, and w for 1 - s, 2 - s, and 3 - s
type transitions:
1- s. v> 0, u> 0 and v < 0, u > v 2(4w)-lj
(28.22)
2-s. v>O, u<Oj 3-s. v<O, u<v2(4w)-1.
It follows from these inequalities that the value of R for 2 - s type transitions
should lie within the limits 1 < R < 00 and for 3 - s transitions, within the
limits ~ < R < 00. By virtue of equation (28.17) we may then write down the
range of variation of the ratio XO/Xd:
Thus we have obtained a very simple criterion which permits the type of a
first-order transition to be determined from susceptibility measurements. The
results thus obtained refer to susceptibility in terms of proper parameters 1].
(28.25)
setting u < O.
The expression for the susceptibility X = - ({)2~ f {)X 2 ). has the form
(28.26)
(28.27)
x (a) x (b)
L-~ I I T
L
Using the standard techniques outlined in §8, it is easy to find that the IR
sought is the one-dimensional representation T4 of the star of the wave-vector
~1 = ~b1 (star {~12}). Since the star {~12} has three arms, the Dth --+ D~~
transition is described by a three-component OP (171,172,173). The transforma-
tion properties of the OP are described by the three-dimensional matrices of
the representation D induced from the representation T4. By constructing the
matrices of the representation D it can be readily verified that a set of different
matrices of this representation generates an I group isomorphic to the point
group O. Therefore, the potential may be immediately written out in the form
n.
~11 = r (2
171 + 1722 + 1732) + Ul (4
171 + 1724 + 1734) + U2 ( 171172
2 2 2 2 + 172173
+ 171173 2 2) . (29.1)
The 0 P (171, 172, 173) refers to a transition with", i 0, so it does not describe
the macroscopic properties of the dissymmetric phase. To analyze the influence
of an external field on domains, macroscopic variables should be included in the
potential. We choose as these variables the strain tensor components Xi. For
hexagonal crystals the strain-induced contribution is described by a potential
of the form
(29.3)
(29.4)
The symmetrized combinations of the representations T1, T9, and Tll are
T1(A 1g ) :
T9(E1g) : (29.5)
Tll(E2g ) :
246 CHAPTER 7
respectively.
The data obtained enable us to write the form of the coupling type for the
parameters Xi and TJ>.:
bll\)Domlli'>l ~/
e" /
/
a" . /
------~ e,D
a' b'
Doma.i?l I
{J
e"' \
ail' \\
Doma.';. J \
b'" \
To determine what form the solutions to the equations of state should have
for the different domains, we use the condition for the density function 8p to be
invariant under the symmetry group of the low-symmetry phase. Using three
crystallographic ally equivalent directions, we choose the subgroups D~~ of the
group Dth. This corresponds to three pairs of domains (Fig. 7.17). From the
PHASE TRANSITIONS IN AN EXTERNAL FIELD 247
condition gop = op, where g E D~~, we obtain the equation of state for the OP
components, which gives the form of the solutions for the different domains:
The plus or minus sign refers to the pair: domain - translational domain.
In the following we restrict ourselves to an exploration of the solutions
(29.7). On substituting the relations (29.7) into the equations of state fJif> / fJ1]).. =
o and fJif>/fJxi = 0, we find the expressions for the spontaneous OP 1]).. and
strains Xi for each of the domains, 1 to 3, [17,18].
1.
(29.8)
2.
(29.9)
where the coefficients C 1 , C2 , C3 and the parameter 1]2 have the form
03
C 1 = 2(02 C13 - 01 C33) [ (C11 + C12 )C33 - 2
2C13 }
-1
,C2 = - C 66 '
(29.11)
C3 = [201C13 - 02(C11 + C12)] [(C 11 + C12 )C33 - 2Cf3r1,
(29.12)
for explicitly, we need to add to the potential (29.1), (29.2), (29.6) a term ~int
of the form
which describes the interaction energy of all the three domains with the exter-
nal fields Xi (in the present example Xi stands for the external-stress tensor
components).
Consider the case of uniaxial pressure: Xl < 0, X 2 = X3 = X6 = O. From
the expressions (29.8)-(29.10) and (29.13) it follows that ~:~t < ~:~t = ~int.
In the case of uniaxial pressure the domain boundaries in the crystal should
therefore be expected to shift in such a way that the domains 1 and 2 vanish
and, as a result, a single-domain state may be reached (the coefficient C 2 is
assumed to be positive [17,18].
By applying uniaxial pressure along the other axis in the basal plane of
the initial phase D~h X 2 < 0, Xl = X3 = X6 = 0, we get ~i~t > ~i~t = ~int'
that is, the domain to vanish in this case is the domain 3, and the domains 1
and 2 should occupy equal volumes of the sample.
The results of experimental investigations into the effect of pressure on
the domain structure in (NH 4hS04 are in good agreement with the picture
described above. This provides further conclusive evidence that the paraphase
D~h has been chosen correctly in these compounds (§10).
From the example considered, it is clear what one should expect in the
general case. It is evident that application of an external field should lead
to shifting of the domain boundaries and to conversion of the sample to a
single-domain state in all the cases when the change of the point symmetry
of the crystal allows the occurrence of spontaneous macroscopic properties:
strain, polarization, magnetization [19,20]. Owing to the difference of these
quantities in different domains, directional influence may then be exerted on
the domain structure by applying to the sample an external field conjugate to
these quantities: pressure, electric field, magnetic field. What has been said
above holds good also for higher-order ferroics, when the low-symmetry phase is
characterized by the occurrence of macroscopic properties that are described by
high-rank tensors (§23). For the domain structure to be affected, the sample in
this case should be subject to certain combinations of external fields of different
nature: pressure and electric field, pressure and magnetic field, etc.
PHASE TRANSITIONS IN AN EXTERNAL FIELD 249
References
Martensite Transformations
The previous chapters dealt with first- and second-order transitions at which
the symmetry group G 2 of the low-symmetry phase is a subgroup of the parent-
phase symmetry group G 1 : G 2 C G 1 . Such transitions are called group-
subgroup related transitions. In addition to these, there exist also so-called
reconstructive phase transitions, at which the condition G 2 C G 1 is not fulfilled,
that is, the low-temperature-phase symmetry group G 2 is not a subgroup of
the high-temperature-phase symmetry group G 1 .
Two reconstructive phase transition types are distinguished [1]. One of
them includes transitions due to diffusion processes. Transitions of the second
type include those realized by cooperative displacement of crystal atoms. In
the following we concentrate on the second type of transitions.
The consistent diffusionless character of displacements at reconstructive
phase transitions of the second type provides a basis for constructing a variety
of schemes that describe the crystallographic relation of the phases G 1 and G 2
[2,3]. Two cases may evidently be singled out. In the first case the groups G 1
and G 2 are subgroups of one common supergroup Go : G 1 eGo, G 2 eGo,
and G 1 and G 2 are not group-subgroup related. Then, choosing the phase G 2
as the parent phase, standard Landau theory methods (see Chap. 5) enable
one to construct a phase diagram. The relative position of the phases on the
phase diagram is sketched in Fig. 8.1. It is seen that a chain of transitions
Go ~ G 1 ~ G2 is possible under certain conditions (thermodynamic path
A - B). The last link in this chain is a reconstructive transition G 1 ~ G 2 .
Such a situation occurs in many crystals that undergo a series of struc-
tural phase transitions. For example, the cubic perovskite BaTi0 3 exhibits
250
MARTENSITE TRANSFORMATIONS 251
Gs C G2 ·
By convention, the phase transition G 1 -+ G 2 may be broken up into two
stages by introducing a hypothetical intermediate phase G. : G 1 -+ G s -+ G 2 •
In a number of cases such an intermediate phase has been detected experimen-
tally. For example, CU2Te exhibits a reconstructive phase transition D~h -+
O~. Using various experimental techniques [4], several intermediate phases are
recorded over the temperature interval between 30 and 40°C. A similar picture
is observed in Cu 2S, ZnS, and GdS [5].
Reconstructive transitions via an intermediate phase embrace transitions
in interstitial solid solutions [6,7]. As a function of interstitial atom size and
concentration, either ordering-type transitions giving rise to a superstructure
(see §4) or reconstructive transitions may occur in interstitial solid solutions.
Superstructures occur predominantly in transition metal hydrides since hydro-
gen weakly distorts the metallic lattice, whereas nitrides and carbides display
chiefly reconstructive transitions. For example, with interstitial solid solutions
of carbon in b.c.c. metals of composition Me2X (Me=V, Nb, Mo, W; X=C) a
reconstructive transition gives rise to a structure with an h.c.p. lattice [6].
Group-subgroup related structural phase transitions, as a rule, satisfy Lan-
dau's concept of the transition going over one IR, that is, the concept that the
restructuring of the crystal structure at such transitions is described by modes
of one of the parent-phase symmetry group IR's. At reconstructive phase tran-
sitions the crystal lattice restructuring cannot be described by modes of a single
IR. A feature peculiar to reconstructive transitions is that along with displace-
ment modes, lattice deformation plays an important role in the restructuring
of the crystal structure. Thus shear deformation of (110) planes along [110] [8]
directions plays a leading role in the b.c.c. - f.c.c. transition.
Phase transitions in which deformation plays a determining role show typi-
cal anomalies in physical properties, a fact which provides grounds for isolating
such transitions into an individual class, the class of martensitic transitions [9].
Characteristic features of martensitic phase transitions are [1]: (1) the phase
transformation is realized when the temperature is varied, (2) the martensitic
transitions are accompanied by a substantial change in sample shape, (3) the
formation of a new phase proceeds at a high rate, and (4) deforming the sample
promotes a martensitic transition.
It should be emphasized that both reconstructive transitions, for example
the a - 'Y transformation in iron, and continuous transitions, for example the
structural phase transition O~ -+ D~h in A-15 compounds, may belong to
MARTENSITE TRANSFORMATIONS 253
(a) (b)
Fig. 8.2. Projection of the b.c.c. structure on the (001) plane (a) and
on the (110) plane (b).
of the reference plane are displaced in one direction by the same amount. In
both cases the adjacent planes are shifted relative to each other by the same
amount ~.
Q
II
L, •'6
0
,2) i
-
1/ ]~----
'7 16 --1/
/!J" ./ c / fI' r> !J
9
.,/'
,.:'5
X
0
3 "1/
,
_C
'
II
2
Let us see what kind of structure results from the realization of type A dis-
placements in the b.c.c. structure. For this type of displacement we assume the
Po plane (Fig. 8.3) to be fixed. We denote the displacements of the atoms lying
in the P+ plane by solid arrows (Fig. 8.5) and those of the atoms lying in the
P _ plane by dashed lines. The numbering of the atoms presented in Fig. 8.5
coincides with that of the atoms of the b.c.c. structure sketched in Fig. 8.3. As
a result, a three-layer structure such as that shown in Fig. 8.6 arises which, by
the type of layer arrangement, is reminiscent of an f.c.c. structure. The result-
ing three-layer structure differs from the h.c.p. structure in that the triangles
isolated in Fig. 8.5 are not equilateral. We can satisfy ourselves that for an
h.c.p. structure to be obtained, it suffices to compress the resulting three-layer
structure along the z axis and to extend it along the [110] directions. These
additional atomic shifts are shown in Fig. 8.5 by dotted lines.
The magnitude and direction of extra shifts at the b.c.c.-f.c.c. transition
may be obtained using the solid spheres model [11]. One of the conditions
that the solid spheres model satisfies is the requirement that the number of
abutment joints of spheres be maximal for a given type of packing. When the
atoms of one of the (110)-type planes are displaced along [110] by scheme A this
256 CHAPTER 8
Fig. 8.5. Projection of the b.c.c. structure on the (110) plane. The
solid arrows show displacements of atoms lying in the P+ plane; the
dashed arrows show displacements in the P _ plane; the dotted lines
represent extra displacements. The numbering of the atoms is the
same as that in Fig. 8.4. 0, atoms of the P+ plane; *, atoms of the
P _ plane; 0, atoms of the Po plane.
Fig. 8.6. f.c.c. structure and its projection on the (111) plane. 0,
atoms of the Po plane; 0, atoms of the P+ plane; *, atoms of the P_
plane.
~.-
/
/
/
o
•
/
/
I /
/
/.;:;-
~"'//
/
/
/ '-; / """/o-::---+--:::---tl·-"---:M
1/ /
Fig. 8.8. Projection of the b.c.c. structure on the (110) plane. The
solid arrows show displacements of atoms in the P+ and P_ planes;
the dotted lines represent extra displacements. The numbering of the
atoms is the same as that in Fig. 8.4; the symbols are the same as
those used in Fig. 8.5.
&4
@-
Fig. 8.9. h.c.p. structure and its projection on the (001) plane .• and
0, atoms lying in the planes z = 0 and z = ~ respectively.
ORIENTATION RELATIONS
A feature typical for orientations of b.c.c., f.c.c and h.c.p. lattices relative to
one another at reconstructive transitions is the presence of invariant planes
and parallel crystallographic directions of these lattices. It is this fact that
provides a basis for introducing the concept of orientation relations. Namely,
MARTENSITE TRANSFORMATIONS 259
Fig. 8.11. Projection of the f.c.c. and h.c.p. structures. The symbols
are the same as those used in Figs. 8.6 and 8.9. The arrows indicate
C-type displacements of the atoms.
:2 3
P;---i\
1\'1/ \
4 I - - \~I
Q,- "
0 - - -I" 6
olvIs' J1",
\ \ I
\/.--v
(a) (b)
8.6, the orientation relations for the b.c.c. - f.c.c. transition have the form [17]
Similarly, we can determine from Figs. 8.8 and 8.12 the orientation relations
for b.c.c. - h.c.p. transformations:
_I_l~ ,\
Gi l __LL I
I
y
j-, ~ 0 1\ ,I
A
1\ I -.!..\, f ;)
V1 w~ i ~ r
K ~ 0
_JJ_LL1L_
/
/ I
\ I
I
,
I I
I ,
I
Gs. lLLU ,
I .J;LF 0
I
~l ~V
t 'V
Y ~ '{ 'f ~
LLJ __
I
l_J~ G:l.
cell edges:
Me: (O,-D.,O), (!,D.,!), (O,!-D.,!), (!,!+D.,O);
(30.4)
X: (~,~, 0), (~,~, 0), U,~, !), (~,~, !).
Then the following expressions may be written for the structural amplitude [2]:
(30.5)
fMe, fx being the atomic scattering amplitudes and h, k, I the indices of the
reciprocal lattice sites in the (x', y') coordinate system.
FI = 2fMee-i4'Trn2fl (1 _eiS7rn2fl);
h = 2nI + 1, k = 2n2, 1= 2n3; (30.6)
F2 = 2fMee-i27r(2n2+I)fl (1 + ei47r (2n +I)fl);
2
The rest of the reflections describe the scattering from metallic and interstitial
264 CHAPTER 8
2 6
t-_,_
-.
J.
z
-0
The requirement that the displacement Ll be small (Ll ~ 1) limits the set of
possible m and n2 values: 12m + 11 ::; 12n2 + 11- As a function of the values of
m and n, the critical displacements are equal to t,
112 , 210' etc. Each such Llc
has its own period, which specifies the extinct subsystem of sites. For example,
MARTENSITE TRANSFORMATIONS 265
(30.9)
The condition (30.9) bounds from below the value of critical displacements:
~c min = 112 , The system of reciprocal lattice sites that corresponds to ~c min is
depicted in Fig. 8.18, where extinct sites are marked with crosses. To identify
the lattice type of the final phase G2 by the system of sites that remains, we
have to allow for the fact that the structural amplitudes degenerate at ~c = /2'
The resulting equations imply that the intensities of the corresponding reflec-
tions are equal in magnitude (Ii '" IFi 12 ), the intensity of each of the reflections
F3 - F6 taking on four different values. As a result, the relative position of the
reflections that remain and their intensity prove typical for the h.c.p. struc-
ture. Figure 8.18 shows the orientation of the h.c.p. reciprocal lattice relative
to the initial f.c.c. lattice. The obtuse angle of the hexagonal cell which has
been obtained on the basis of only displacements ~, without taking account of
the homogeneous deformation, is equal to '" 109 0 • Therefore to obtain an ideal
h.c.p. structure, allowance must be made also for the homogeneous deforma-
tion of the reciprocal lattice, as was the case in analyzing atomic displacements
at the b.c.c. - h.c.p. transition (Fig. 8.8). The interrelation between the unit
cells of the b.c.c. and h.c.p. phases is demonstrated in Fig. 8.8.
(31.1)
for the extra displacements. The last line describes the displacement field
corresponding to additional atomic displacements along the z axis at the b.c.c.
- f.c.c. transition; these displacements are shown by dotted arrows in Fig. 8.5.
MARTENSITE TRANSFORMATIONS 267
From the definition of the strain tensor components, it is then easy to find the
corresponding components
(a) (b)
in a somewhat different form, the way they are shown in Fig. 8.20. As can
be readily seen from the formulas (5.1)-(5.6), the type B displacement field
portrayed in Fig. 8.20 is described by the optical mode of the IR T4 with the
wave-vector", = ~b3 (where b 3 is the reciprocal lattice vector) of the group
O~. Extra displacements at the b.c.c.-h.c.p. transition, similar to those in the
case of b.c.c.-f.c.c. transformations, are described by strain tensor components
fa{3,2,3 (31.3).
if:
'I' = 1'1- (2
e2 + e32) + 1'2
- (2e4 + e52 + eii") + 3we3
1 - (2
e3 - 3e22) +
(31.8)
We determine the stability regions of these phases. The stability conditions for
the phase (00) are described by the inequalities r1 > 0 and r2 > O. For the
phase (eO) we have the expressions for the OP:
(: - 1
"-2
+ (14 _ 2r 2 )1/2 , (31.11)
(31.12)
(31.13)
The (00) -+ (eO) transition line is found from the condition that the energies
of the phases be equal in magnitude, <1>(00) = <1>(eO), and has the form
r2 = 1/9. (31.14)
The phase transition lines (00) -+ (ep) and (eO) ;::: (ep) were found numerically
and the final form of the phase diagram for the cases (7 0.5 and (7 =
-0.5 =
is given in Fig. 8.21. The potential is treated similarly in the theory of quasi-
crystals (see §42).
'~
Ai
G = 0.5
(00 )
- - - ----::-7
!! 2
~ ~
Uf)
Fig. 8.21. Phase diagram for (7 = ±0.5. The solid lines refer to first-
order transitions; the dashed lines refer to second-order transitions.
-7f-~
C.;-- - -7fY-
~~ f
('--/-- ~ J
,[<i T~()
J 6 1
Fig. 8.22. Shape of the P1(0 function at different values of the para-
meter r2.
e
terms of the form -P1 and -P2 p. When the applied stress is conjugate to the
parameter e, the condition that the potential be minimal yields
P1 = e - e + 2r2e· (31.1G)
The shape of the function P1 is shown in Fig. 8.22. For r2 > ithe function P1
has no point of inflexion (Fig. 8.22b). In the interval ~ < r2 i
< the function
P 1 has extrema at the points (Fig. 8.22b):
In this interval segments of negative slope arise on the right-hand wing of the
function P1 (0. Consequently, on reaching a certain value of P1 , a first-order
phase transition characterized by a hysteresis loop should take place in the
crystal (see §25). At r2 = ~ the function P 1 (e) is tangent to the abscissa axis
at the minimum point. Finally, the shape of the function P1 (e) for r2 < ~ is
shown in Fig. 8.22d,e,f.
272 CHAPTER 8
P{ =e-e+2r;e, (31.18)
p2 = -2(O"e + rd,
where P{ = =
PI + 0"1'1 and 1'; 1'2 - 0"2. The shape of the function P{ (e) is the
same as that in the previous case. A point of dissimilarity is that the condition
(p2 = 0) as a function of the parameters 1'2, 1'1 and 0" separates the region of
e
all values of into two intervals, one with p2 > 0 and the other with p2 < O. In
e
the first interval (p2 > 0) the P{ versus curve is described by the expression
(31.15), while in the second interval (p2 < 0) the relation is described by the
expression (31.16). On the boundary of these intervals the function PI (e) has
a cusp.
Consider the case where the external field is conjugate to the parameter
p. From the condition that the potential be minimal, we obtain
}'--.,.....
• .P
Fig. 8.23. Variation of the shape of the P2(P) function at (J" > 0 as
the temperature is lowered.
----.1'----1
+-~l-_f_7· /
Fig. 8.24. Variation of the shape of the P2(p) function at (J" < 0
as the temperature is lowered. The difference of the external field
effect on strain at (J" < 0 lies in the P2 (p) relation exhibiting S-shaped
anomalies.
274 CHAPTER 8
The shape memory effect implies the ability of certain distorted-shape alloys
to recover their original shape spontaneously as the external conditions are
changed [24]. Examples of such alloys are Cu-20.4Zn-12.5Ga (at.%), Ag-45Cd
(at .%) and Cu-39.0Zn-4.6AI (at .%). The essence of the original shape -+
distorted shape -+ original shape cycle is elucidated by the diagram presented
in Fig. 8.25 [25]. At the first stage of this cycle an austenitic phase of a given
shape is cooled down to the martensitic transformation temperature. At the
point T = Tm a martensitic phase transition takes place as a result of which
the initial (austenitic) phase is replaced by a new (martensitic) phase. The
martensitic phase arises in the form of domains. The volume of each domain
differs from the volume of the corresponding amount of austenite; however,
since these domains are appropriately aligned relative to one another, the shape
of the entire sample changes insignificantly. All domains are iso-energetic. By
applying pressure to such a sample, we disturb the energy equality of different
MARTENSITE TRANSFORMATIONS 275
In,,",)- J'mc.;l1
m",~I(n" Ie
t
--L - - --1,--
Ht:tl.Lng- ~ ioaoi;'I1f}
1111111111111[1 ~
ill UnevvJ, ~g-
1IIIIIIIIIIIIIIIi
\-----1--1
1--L-t:.L --L -ilL
'::"nlJfe -doma;Y/ S;'l1aee -Joft,""'n
'marfenr," .. rn"~ te"r ife
With the crystal as a whole rotated homogeneously, the energy of the crystal
should not change. Therefore, the thermodynamic potential is independent
276 CHAPTER 8
of the tensor components wOl {3, but it may depend on their derivatives, which
should be included when writing down the gradient terms.
Also we need to invoke the compatibility equations coupling the tensor
components fOl{3 and WOI{3. This is because not all of the tensor components
are independent, for they are expressed in terms of derivatives of the three-
component quantity U OI ' that is, the displacement vector. Mathematically,
these relations follow from the condition that the mixed derivatives be equal
in magnitude:
U OI ,{3-y = u OI ,-y{3' (32.2)
Elimination of W from this expression gives the relation between the second-
order derivatives of the strain tensor components only:
called the compatibility condition. In explicit form, this equation splits up into
six equations:
fxx,yy + fyy,xx = 2fxy ,yx;
However, only three independent equations remain [26]. Using equations (32.3)
and (32.6), we may eliminate w from the gradient invariants of the thermo-
dynamic potential and find the independent invariants. Homogeneous invari-
ants' for the cubic group in terms of the symmetrized combinations (31.15)
MARTENSITE TRANSFORMATIONS 277
<Pinh = ~ J
dr{ alel~el + a2(e2~e2 + e3~e3)+
+d3el(~2e2 + ~3e3) + a4(e3~3e3 - e2~3e2 - 2e3~2e2)+
+d5(e4~e4 + e5~e5 + e6~e6)+
+a6(e4~5e6 + e5~6e4 + e6~4e5)}. (32.8)
(32.9)
which transform according to the same IR as the quantities el, ... , e6 (31.15)
do.
EQUATIONS OF MOTION
6L d 8L
(32.10)
8uO/ - dt 8uO/'
278 CHAPTER 8
(32.11)
and the potential eI> is determined by the sum of the expressions (32.7) and
(32.8). The equation (32.10) may now be rearranged to read
.. 8eI>
pUO/ = --c-' (32.12)
uUO/
The explicit form of the Lagrange equation may be obtained by writing out
the variational derivative of the potential with respect to strains:
rest of the strains are absent in the entire crystal, that is, el = e4 = es = e6 = O.
Suppose also that the state of the crystal remains homogeneous along the z axis.
The general compatibility equations (32.6) here reduce to
82e3 82e3 8 2e3 ( 82 82 )
8x2 = 8y2 = 8x8y = 0, 8x2 - 8y2 e2 = O. (32.14)
(32.17)
Minimization with respect to ¢J yields three values of the phase, ¢J = 11', 11'/3,
and -11'/3 (for definiteness, we assume that V4 > 0), which correspond to the
elongation of the crystal along one of the cube axes. Minimizing with respect
to P, we determine the phase transition point rc = VV4U4, the OP Po jump
at this point and the temperature dependence of energy as a function of the
OP (Fig. 8.26). The phase transition from the initial cubic (p = 0) to the
tetragonal (p i 0) phase occurs when the potential's minimum 1>h(p) touches
the abscissa axis.
280 CHAPTER 8
We now write the equations of motion for the OP components, using the
general form (32.12) and the potential (32.16). We have
.. 82 82 8 8
pe2 = 3 ( -2 + - ) -
8x 8y2 8e2
v'3(-8~
- -)
-
8x 2 8y2 8e3
-8~2 2
(32.18)
.. 8 8 8 8
pe3 = -v'3(-2 - -)-+ (-+ -)-
2 8~ 2 8~ 2 2
Recall that the solutions of equations (32.18) must satisfy the compatibility
equations (32.15). We now take the form of the solution with e2 depending on
x - y:
e2 = e2(x - Y - vt) == e2(r).
For solutions of this particular form (allowing for e3 = const) equation (32.18)
reduces to
(32.19)
8~ = 0, (32.20)
8e3
where e20 = e2 (r -> ±oo).
The second of these equations may be represented as
(32.21 )
Only with this quantity is the condition e3 = const fulfilled. Thus the above
condition is fulfilled only at one temperature point. At all other temperatures
we have to relinquish the condition e3 = const.
MARTENSITE TRANSFORMATIONS 281
Considering the limit T ..... ±oo, we see that if e2(+oo) 1= e2(-oo), then the
equation is satisfied only at v = O. Thus a domain wall at rest exists at temper-
ature (32.21) which separates two phases with e2(+oo) and e2(-oo) (Fig. 8.27)
[27]. These phases are two martensite domains with different tetragonal axis
orientations. The corresponding solution of equation (32.22) is
\
\
\ Jl I
From the formulas thus found we can reproduce the displacement fields.
The complete set of strain-tensor components in our case has the values:
(32.24)
From the first three equations we obtain expressions for the spatial derivatives
of the displacement vector:
where
F(x,y) = fo X
-
y
dte2(t) = C1'2~0:4r/2Incosh(X~Y)
and h, h and fa are arbitrary functions of their own arguments. The arbitrary
functions may be determined from the last trio of equations (32.24):
(32.26)
MARTENSITE TRANSFORMATIONS 283
Taking into account the boundary conditions e6(Y = ±oo) = 0, the compati-
bility equations
{} 2e 6 _ {} 2e 6 _ {} 2e 6 _ {} 2e 6 _ 0
{}x{}y - {}y{}z - {}x{}z - {}y2 - (32.27)
e6 = f(z) + j(x).
Thus we shall try to solve dynamic equations that depend only on one coordi-
nate z or x.
The potential (32.9) + (32.10) under the conditions (32.26) reduces to the
functional
(32.28)
contains only one literal parameter r, that is, the temperature (the prime de-
notes the derivative with respect to the coordinate). This functional and the
associated dynamic equations have been investigated by Falk [28,29].
The homogeneous part of the functional fo = re 2 - e4 + e6 as a function of
the OP e exhibits different behavior in four temperature intervals (Fig. 8.28):
(a) r> 1/3. There is one minimum fo at the point e = O. This minimum
corresponds to the stable point of the austenitic phase.
t
(b) < r < ~. Two extra degenerate minima arise at the points eo. These
minima correspond to the metastable martensitic phase; the minimum e = 0
corresponds to the stable austenitic phase.
284 CHAPTER 8
(c) 0 < r < ~. As before, there are three minima, but the energy of the
martensitic phase becomes lower than that of the austenitic phase, which is
metastable in this interval.
(d) r < O. There are only two minima ±eo, which correspond to two
domains of the stable martensitic phase.
llL
\
o e e
- eo 0 eo)
(a) (b)
10 10
e
(c) (d.)
eo = ± (
1 + VI
3
- 3r) 1/2 . (32.30)
2 1/ dlo
v e + e - de = + Cl,
2
C2 Z (32.34)
where the prime now denotes differentiation with respect to T. From the bound-
ary conditions (32.33), we find
The last relation means that (a) e oo = e_ oo and Cl = v 2 eoo is a moving soliton
in one phase and (b) e oo ::p L oo , Cl = 0 and v = 0 is a domain wall at rest.
Integrating equation (32.34) with respect to de yields the integral of motion
(32.35)
where Fa is the integration constant. Using this equation we can readily write
1: 1:
the total soliton energy E corresponding to the solution
We find the localized solutions of equation (32.25) in the four different cases
determined by the boundary conditions.
(1) Domain boundary between the austenite and the martensite (e_ oo =
eo, e oo = 0, v = 0). The conditions e oo = 0 and e~ = 0 give Fo = O. From
the conditions Coo = eo and e~oo = 0, we find fo(eo) = 0, which is fulfilled at
the point r = ~. Thus an austenite-martensite domain wall at rest exists only
at the phase transition point. Under these conditions equation (32.35) reduces
to
(32.37)
-1/2
e(x) = [2(1 + eX-XO) ] (32.38)
(32.39)
The condition for a local solution to exist is e'2 > 0 over the entire range
-eo < e < eo, so that ei should be greater than zero; from this it follows
that r < ~. We arrive at the natural conclusion that the domain boundary
exists only below the phase transition point. The shape of the boundary is
determined by the solution of equation (32.39):
e(x)
X-Xo][ a + sinh 2 -~-
= [eo sinh -~- X-XO]-1/2 , (32.40)
where
3e 02 - 1
C 1 =eo(3e 6- 1)1/2, a= 22 l'
eo -
The width of the wall is determined not only by the quantity ~, but also by
the parameter a involved in the denominator. Specifically, when r --+ ~ (as
the transition temperature is approached) e6 --+ ~; therefore the parameter a
diverges, leading to an infinite wall width.
MARTENSITE TRANSFORMATIONS 287
mf2iastQbfe
a L( s t e ni i e
o "-----------p> lJ2
(32.41 )
t=
For a localized solution to exist, we must have a value of e 0 such that e' = 0,
that is, a turning point must exist. This means that we must have e~ > 0 and
e~ > 0, whence the velocity v is included in the interval
(4) Localized austenitic region in the martensite (e_ oo = eoo = eo, vi- 0).
The equation of motion becomes
(32.44)
this root being closest to eo. An analytic solution of equation (32.44) is difficult
to obtain since here we need to know the roots of the quartic equation (32.45).
In the metastable-martensite region (~ < r < ~) we can obtain a zero-velocity
soliton solution:
1 - 2efi ] 1/2
e(x)= [ . , (32.46)
1 - ((3efi - l)/efi) tanh 2 ((x - xo)/e)
where
For finite velocities as well as for the stable-martensite region, numerical solu-
tions of equation (32.44) have been obtained in [29].
We have considered separately the reconstruction of a cubic lattice due to
tetragonal and shear strains. As was shown in detail in §30, the martensitic
b.c.c. --+ f.c.c. transition involves both strains simultaneously and is described
by the general energy functional (32.7) + (32.8). The first approximate treat-
ment, however, may incorporate only fundamental displacements of type A
(Fig. 8.19a), which are described only by the shear strain in the system of
coordinates directed in the (001) plane along the lattice diagonals. Therefore
the solutions obtained under the restriction (32.36) may be a first step toward
a discussion of the inhomogeneous states at the b.c.c. --+ f.c.c. transition.
Since this approximation disregarded tetragonal strains (and also the change
in volume), the question of a complete description of this transition remains
open.
SQUARE LATTICE
The general functional for three-dimensional space is too complicated to
permit an understanding of all the regularities in the interplay of inhomoge-
neous strains. Therefore it is reasonable to go over to the two-dimensional case,
MARTENSITE TRANSFORMATIONS 289
where the martensitic transition is modeled by the transition from the square
lattice to the triangular lattice (Fig. 8.30). The necessary displacements are
described by the following four strain tensor components:
The quantity el is the change in volume (area), e2 describes the strain along
one of the axes, e3 describes the shear strain and w describes the torsional
strain.
~~.--4'~'~-·~~
;, ~ ~
At..'" ~ ~
----:~: ~:-----..!---...:
Fig. 8.30. Geometry of the transition from the square to the triangular
plane lattice. The symbols are the same as those used in Fig. 8.5.
(32.48)
(32.49)
This equation is a particular case of the compatibility conditions (32.5) for the
two-dimensional space.
We write the thermodynamic potential in the form
<I> = J
dr¢(r), ¢(r) = ¢h + ¢inh(r), (32.50)
290 CHAPTER 8
where
'l'h
A-.
= rl e2
l + r2e2+ r3e3+ weI + Ul el + U2 e2+ U3e3+
223444
[( ae 1 2 2
</Jinh(r) = 11 ax )2 + (ael)2]
8Y [(ae
+ 12 ax )2 + (ae
ay )2] +
ae3)2 + (ae3)2]
+/3 [( - -ay + vI [ael
i: ae2 ael ae2]
-- - - (32.52)
ax ax ax ay -ay .
Thus only four inhomogeneous invariants are independent. The other three
invariants
(32.53)
!:!:...
8e3
= 0, (32.54)
strain e2 or e3' The equations (32.54) in the two-dimensional case are too
complicated:
0
(oxcj22 + Oy2)
2 3 2
13 e3 = r3 e3 + 2U3e3 + 2v3e2e3.
r2 = r3 == r, 12 = 13 == I' U2 = U3 = V3 == U,
this system reduces to a two-dimensional non-linear Schrodinger equation for
(32.55)
This equation was used earlier in the theory of vortices in a non-ideal Bose gas
[30]. The above equation was shown to have vortex solutions, which in our
case describe a localized austenitic region in the martensite. Indeed, at infinity
the function 1/J tends to the equilibrium value 11/J12 = e~ + e5
= r/2u, while at
the center of inhomogeneous distribution (r = 0) it has an asymptote 1/J '" r V ,
where v is an integer specifying the azimuthal dependence of 1/J.
We now investigate the solutions in which two of the four strains el, e2, e3
and ware equal to zero in the entire plane. In four cases,
el = w = 0, e2 = w = 0, e3 =w = 0 and e2 = e3 = 0,
the minimization equations yield homogeneous solutions for two extra strains.
In two cases, el = e2 = 0 and el e3= =0, a localized plane solution arises,
either in the direction of the edge of the parent-lattice square or in the direction
of the diagonal. For example, el = e2 = 0, w = e3 - 20',
(32.56)
6~R~!
9
1'r / ! 0 _~
~
3 4·
initial b.c.c. structure, Gint is the symmetry group of the intermediate w-phase,
and G 2 = D~h is the symmetry group of the final phase, called an ideal phase
or an aging w-phase [32]. The w-phase is realized in disordered quenched Zr-
and Ti-based alloys, as well as in Nb-Mo, Mo-Re, and other alloys [32].
According to X-ray structure analysis data, the w-phase is described in
various systems by either the wave-vector "'w = ~(111) or the wave-vector
'" = "'w + 8"', which is close to the former vector. In this context we wish to
note that since the point "'wis not a symmetric point of the reciprocal lattice,
the structure of the w-phase is independent of the wave-vector type ('" or "'w)
by which it is described. For definiteness, we will consider the vector "'w. Let
us reveal the IR's that describe the w-phase structure.
The wave-vector "'w = ~(b1 + b 2 + b 3 ) belongs to the eight-arm star {~7}
of the b.c.c. lattice [33]. The h.c.c. lattice contains one atom 1'1 = (000) per
unit cell. From the general formulas (5.1)-(5.3), we readily obtain that the
mechanical representation decomposes into two IR's of the group G"w'
(33.1 )
The modes of the representations 71 and 73 for the arms K-w and -K- w , found
by the formulas (5.5), are given in Table 8.1. The modes for the other arms
K-L of the star {K-7} are easy to find by the formula (5.6). The values of the
modes on the nth atom are determined by the expression
'!/JOt ( "LVI)
>. n >. I0)e'''Lt
= '!/JOt ("<tV . n, (33.2)
'!/JOt e In)
LV being the atomic component of the basis function in the nth cell.
Table 8.1
Displacement modes for group G" arms ±K-w
7 K- r1(0 0 0)
71 K-w 1 1 1
-K- w -1 -1 -1
K-w 1 (2 (
73
(2 1 (
_ (2 - (
-K- w -1
_ (2
-1 -(
The structure of the w-phase is described by two modes 71 with arms ±K-w .
Since the vector K-w is parallel to the [111] axis, the displacements of all atoms
lying in the (111) plane are equal in magnitude. Therefore, the formation of
the w-phase may be viewed as resulting from displacement of (111)-type planes,
as shown in Fig. 8.31. Indeed, the atoms of the P1 plane are not displaced, the
atoms of the P2 plane are displaced by the amount u and the atoms of the P3
plane are displaced by the amount -u along the [111] direction. The resulting
structure of the w-phase has the symmetry D~d. On reaching the magnitude
of displacement u = a-/3/12, the atoms belonging to the P2 and P3 planes
emerge in the same plane (Fig. 8.31) and the symmetry of the crystal rises to
D~h. Figure 8.31c shows that no additional strain is needed to form an ideal
w-phase.
MARTENSITE TRANSFORMATIONS 295
THERMODYNAMIC POTENTIAL
As the OP at the b.c.c. - w phase transition we may choose the mode mixing
coefficients T}1, T}2 for the IR 71. The form of the matrices of this representation
[33] determines the transformation properties of the components T}1 and T}2 and
the form of the invariants involved in the potential:
(33.3)
(33.4)
where
u - u3 u2 (33.5)
P=T}2iiJ' <I> = <I>(2iiJ)4' 7'=r(2iiJ)3'
From the condition 8<I>/8<jJ = sin 3<jJ = 0 we find the equilibrium values of the
angle <jJ = 0, 7r /3, 27r /3. These three solutions correspond to three translational
domains which differ in the phase of displacement waves of (111) planes along
the [111] direction.
We set <jJ = 7r /3 and explore the form of the potential (33.4) at different
temperatures 7' = 7"(T-Tc). To this end, we investigate the condition 8<I>/8p =
0, from which we find
(33.6)
(33.7)
INHOMOGENEOUS STATES
i'*'i. -_ -1
V
J {iii. + - ~ (01]2
dr '*'w U ~
a=l
01]1) + 'Y- ~ -
1]1-- -1]2--
OXa OXa
~
01]1
-- 01]2
a=1 OXa OXa
-} (33.8)
and introduce the dimensionless form of the potential in terms of the variables
p and ¢:
<I> = J
~ dr{rp2+~lcoS3¢+~p4+up2~~ +'Y[l(~~r + (~:r])· (33.9)
In equation (33.9) we have passed over to the derivative along the z direction
parallel to the body diagonal (111) and neglected the inhomogeneity in the
other directions.
MARTENSITE TRANSFORMATIONS 297
which is the sine Gordon equation. This standard equation of the theory of
incommensurate structures will be investigated in detail in the next chapter.
Here we restrict our attention to a known solution of this equation in graphic
representation (Fig. 8.33a). According to this solution, the bulk of the crystal
breaks down into fixed-phase regions within which the structure of the w-phase
is described by the wave-vector "'w (see Fig. 8.31). The boundaries between
these regions correspond to the change of the phase, while the structure of
the regions is described by a soliton solution of the sine Gordon equation. In
the equilibrium state the solitons are spaced out at equal distances, forming a
soliton lattice. In Fig. 8.33b the internal structure of domains is represented
only by two periods. Actually, each domain contains many such cells. The same
applies to the intermediate region (shaded area). The wave-vector displacement
8" is determined by the distance between solitons L : 8", = 27r/3L.
Consider now the approximation <p =const in the functional (33.9). The
equation of state
(33.11)
here is similar to equation (32.34) and has a localized solution [21] for
P2 - P1 2 Z - Zo ] -1
P = P2 [1 + P1 cosh -~- , (33.12)
where
1 1 F2
P1,2 = "2(y'r2=F~, ~ = V"2'
This solution may be viewed as the critical nucleus of the phase in the initial
metastable (supercooled) b.c.c. structure. The shape ofthe nucleus is described
by a bell-shaped curve (33.12), thus distinguishing it from nuclei in heterophase
fluctuation theory [21,34].
298 CHAPTER 8
fi
J
~-L --J
--------.-- t
(a)
111111I11111 ~IIIIIIII
1'1~P;P'~~P1~ ~~~~Pl~~ ~~~ ~P,.~
(b)
References
1. C.S. Barret, and T.B. Massalski: Structure of Metals, McGraw-Hill (1966).
2. V.A. Somenkov, A.V. Irodova, and S.Sh. Shil'shtein: Fiz. Tverd. Tela 20,
3076 (1978).
3. P. Toledano, and G. Pascoli: Ferroelectrics 25,427 (1980).
4. A.B. Plyusinin, A.N. Dubrovina, and S.M. Finarev: Soviet Phys. Cryst.
24, 344 (1979).
5. A.F. Bol'shakov, A.O. Dmitrienko, and B.V. Abalduev: Izv. Akad. Nauk
SSSR, Neorg. Mater. 15, 1528 (1979).
6. A.G. Kachaturyan: Teoriya Fazovykh Prevraschenii i Struktura Tverdykh
Rastvorov (Transforrnations and Structure of Solid Solutions), Nauka (1974).
7. V.A. Somenkov, and S.Sh. Shil'shtein:Z. Phys. Chern. Neue Folge 117,125
(1979).
MARTENSITE TRANSFORMATIONS 299
32. V.V. Kondratyev, and V.G. Pusin: Fiz. Met. Metalloved. 60,629 (1985).
33. B. Horovitz, J .L. Murray, and J .A. Krumhansl: Phys. Rev. B 13, 3549
(1978) .
34. J.W. Christian: The Theory of Transformations in Metals and Alloys,
Pergamon (1975).
CHAPTER 9
Thus far we have considered phase transitions that are specified by IR's of space
groups with wave-vectors lying at Brillouin zone symmetric points (Lifschitz
stars). As a result of such transitions, structures arise in the parent crystal; the
periodicity of these structures is strictly commensurate with that of the parent
crystal. Thus, for example, a magnetic ordering over Lifschitz stars gives rise
to magnetic structures with black-and-white (Shubnikov) symmetry lattices
whose periods are multiples of the paramagnetic-crystal Bravais lattice periods
with a multiplicity of 1,2,3, ... (the maximal increase in unit cell volume is by
a factor of 32).
At phase transitions over non-Lifschitz stars the wave-vector modulus may
vary continuously without a change in symmetry, so the period of the structure
that arises at such transitions may be arbitrary with respect to the period of
the initial structure. Formally, these periods may be incommensurate and
therefore phases characterized by non-Lifschitz wave-vectors are said to be
incommensurate (IC). A multitude of such structures, arising at structural
or magnetic phase transitions, are known (Fig. 9.1). Examples are the various
spiral (helical) magnetic structures (SS) or spin wave (SW) type structures.
In vie\v of restricted experimental accuracy in the determination of the wave-
vector, the problem as to the commensurate ness or incommensurateness of
these structures is very difficult to solve on the basis of a purely structural
investigation. It would therefore be more correct at the present stage to talk of
modulated or long-period structures rather than incommensurate structures,
bearing in mind that in practice the wave-vectors of modulated structures differ
little from the wave-vectors "-0 lying at the Brillouin zone symmetric points,
301
302 CHAPTER 9
(34.1 )
The phase transition occurs when the temperature T = To and the wave-vector
q = "-0 are such that
minA(q) = O. (34.2)
It is convenient to take the origin of the wave-vector to be at the nearest
symmetric point. The quantity A( q) may then be expanded in a power series
of the small vector q. The form of this expansion depends on the symmetry of
the Lifschitz point itself. Two different expansion types may be encountered
for A( q) in the vicinity of the Lifschitz point:
A(q)
q
o ko o ko
(a) (b)
(35.1 )
(35.2)
non-linear term U1]3 generates odd harmonics, so the solution of equation (35.3)
should be tried in the form of a series:
where the amplitudes Ap and the wave-vector I\, should be determined from the
non-equilibrium energy minimum obtained on substituting the solution (35.5)
into the functional (35.1). The calculation of the non-equilibrium energy is a
trivial one and with the help of the simplest integration
IJ
V drcosl\,z = 0 (I\, =P 0), IJ
V drcos 2 K,Z = 2'
1 V1 J 3
drcos 4 I'Cz='8""
(35.6)
306 CHAPTER 9
3u 2 2 3u 2
+2(A 1A 3 + A1A5 + A3 A 5) + '2A1A3 + 2(A1A3 A 5 +A1A3A5) + ...
2 2 2 2 U 3 2
(35.7)
(35.8)
(35.9)
(35.10)
We now need only to minimize this expression with respect to Al and K. The
best way of minimizing with respect to K is by using the general expression
(35.7), whence we obtain the equation for determining the wave-vector:
(35.11)
where the prime denotes the derivative with respect to the argument. Retaining
only the contribution of the first multiple harmonic, this equation yields
K
2
= 2(
KO
A5)
1 - 24 A2 ' (35.12)
1
(35.13)
INCOMMENSURATE PERIODICITY PHASES 307
The quantity L(K), on substituting equation (35.12) into its definition (35.8),
becomes
L(K)=r- ;:[1- (24~~f] =1'- ;: +O(A~). (35.14)
Up to terms At the expressions (35.12) and (35.13) thus yield equations that
give the temperature dependence of the quantities Al and K:
(35.15)
(35.17)
k2 ::::: k2
o
[1 __1 (Te re- T) 2] '
500
(35.18)
The energy of the incommensurate phase is found from the expression (35.10)
on substituting into the latter the formulas (35.15) and (35.16). In terms of
the quantity l' we find to within (Tc - T)3
The energy equality <I>IC = <I>c determines the first-order phase transition point.
The latter is given by
2 2 2 4 (1 + x)3
-1' = xrc , x = 3(1 + x) + 27 63 _ x .
Thus the incommensurate phase in the (1', /) plane lies in a sector defined by
the two parabolas:
/2
r=-x-. (35.22)
4a
A complete phase diagram, corresponding to the functional (35.1), is presented
in Fig. 9.3. This diagram was first constructed by Michelson [1] in the approx-
imation of one harmonic in the incommensurate phase. Allowance for multiple
harmonics changes the phase boundaries but does not alter the qualitative form
of the diagram [2].
From the diagram, it follows that as the temperature is lowered an incom-
mensurate-to-commensurate phase transition, IC -> C, occurs at / < 0 (ther-
modynamic path a). In many cases, however, an inverse sequence of phases is
observed: C -> IC. This sequence may arise for / > 0 if allowance is made for
the renormalization of / by including into the potential ~ an invariant of the
form -5'1}2(d'l}/dz)2. Instead of /, the potential ~ then involves the effective
parameter
(35.23)
and at a sufficiently large 5 in the condensed phase, where 'I} increases with
decreasing T, the quantity t will decrease with the temperature and if it be-
comes negative, a transition to an IC phase will take place. This will lead to
the following temperature-induced change of the wave-vector in the IC phase:
/\'0 ( t)
= - 2a
1/2
'" (Tc - T)
1/2
. (35.24)
Thus the reverse phase transition sequence C -> IC may possibly be due to
displacement along the thermodynamic path b.
INCOMMENSURATE PERIODICITY PHASES 309
e
where "I and are complex conjugate OP components. Minimization of this
e
functional with respect to the components "I and yields the following pair of
complex conjugate equations:
(35.27)
+ 2: (Apn+1ei(pn+l)"z + A_pn+1e-i(pn-l)"z).
00
Substituting this series into equations (35.26), we can readily see that the
multiple harmonics have the following smallness with respect to the parameters
u, w, and (Tc - T):
On minimizing with respect to the amplitude and the wave-vector, the equlib-
rium values of the harmonics are
c
o
A - _ (n - 1)wA~-1 A _ 2uAiA-n+1 .
-n + -
1 +1 - (35.32)
L({n _ 1),..) + 4uAi' n - L({n + 1),..) + 4uAi'
1'2
<I>c = --
4u
- 21wl (1'
- )n/2 .
-2u
Neglecting the contribution by multiple harmonics to the IC phase, on equating
the expressions for <I>IC and <I>c, we obtain the equation for the transition
INCOMMENSURATE PERIODICITY PHASES 311
temperature:
r
r2=(r-rc)2-4ulwl ( -2u
)n/2 .
So far we have assumed that the inhomogeneity of the structure arises
along one direction in the crystal, that is, the star of the wave-vector k has two
arms (~, -~). Structures for which the wave-vector lies in the symmetry plane
and the star of the wave-vector has more than two arms should be described by
functionals with derivatives relative to several directions. A typica.l functional
with a two-component OP for such situations will be
(35.35)
1: Kx
2
= - 2al'
/
Ky = ° or Kx = 0,
2
Ky
/1
= - 2a1 ;
(35.36)
2 2 /
2:
Kx = Ky = - 2a 1 + 2a2 + a3 '
which are stable under the conditions / < 0, a2 + ~a3 > al and / <
0, a2 + ~a3 < al respectively. The symmetry of both solutions corresponds
to a four-arm wave-vector star, both solutions corresponding to a phase with
inhomogeneity in one direction with ~ = (K, 0, 0) or ~ = (K, K, 0).
(35.38)
are found from the following considerations. First, we must require that the
expression (35.37) be positive definite at large values of 1],x. This reduces to
finding positive-definiteness conditions for the higher-order (quartic) form in
terms of energy density:
(35.39)
It suffices to find the conditions under which <P4 takes on positive values at all
the extrema on a circle of constant radius p2 = E,x1]~, since <P4 = p4f, where f
is a function only of the angles determining the spatial position of the vector
'IJ = (1711]21]3).
Thus we need to find the extrema of the functional <P4 - J.tE,x1]~, with J.t
the Lagrange multiplier. The equation
has the following solutions (the value of the extremum for <P is also indicated):
(35.40)
Like the wave-vector, the modulus and phase of the OP component are found
from the energy minimization condition. On substituting equation (35.40)
into equation (35.37) and integrating, we obtain the expression for the non-
equilibrium energy of the incommensurate phase:
(35.41 )
where
(35.42)
'" being the modulus of the wave-vector", = (x:x, X: y , X: z ). The requirement that
the quartic form in A(",) be positive definite at large x: is fulfilled in exactly the
same way as in the previous analysis and we arrive at the conditions <1'1 +<1'2 > 0
and 3<1'1 + <1'2 > 0 enumerated in the complete list of conditions (35.38) for <1:>.
The expression (35.41) says that the minimization with respect to", is not
associated with that with respect to P>. and <P>., that is, with the minimization
with respect to directions of magnetic moments.
The equation
(35.43)
1. (35.44)
2. (35.45)
311 CHAPTER 9
which are stable at 0:2 < 0 and 0:2 > 0 respectively. (It stands to reason that
there exist two more type 1 equivalent solutions with cyclic permutation of
coordinates.) Thus the wave-vector of the structure may lie either along the
cube edge or along the body diagonal of the cube. At an equilibrium value of
/" we have the following values of the coefficient r == ~A("') of the quartic form
in the expression (35.41):
-
r = '21 [ r - 3,2]
4(30:1 + 0:2) , 0:2
0
> . (35.47)
(35.48)
(35.49)
1p: pi = p2, p~ = p~ = 0,
(35.50)
'J'IC
;0..
= rp 2 + '3(
-
8 Ul + U2 ) P4 .
The phases <POt are not defined. The 1p describes the SW structure with polar-
ization along the cube edge.
2_2_122_
2p: PI - P2 - '2 p , P3 - 0,
These angles describe the helix in the plane perpendicular to the cube edge.
3p:
This solution describes the helix in the plane perpendicular to the body diag-
onal of the cube.
u- 1 u=o
2 2 1
2J(2)
k--- Ss ---- U j
U +2..U =0
1 S 2
Fig. 9.4. Phase diagram in the (Ul' U2) plane for the Ginzburg-Landau
functional (35.37). The boundaries of the phases are first-order tran-
sition lines.
316 CHAPTER 9
V1 J
drw (OTJl
OX + OTJ2
oy + OTJ3)
OZ =
2 1 (~
2"w L:P)..K,).. )2 = 2"wp
1 2 2 2
K, cos"p, (35.55)
where "p is the angle included between the vector." = (TJ1TJ2TJ3) and the wave-
vector "'. Depending on the sign of w, either a state with "p = 0 or a state
with "p = 7r /2 is favorable. In the 1p phase, for example, this will be the
LSW or TSW structure. Thus the coupling of the wave-vector to the magnetic
moment vector in the model (35.37) may be due to the inclusion in <I> of new
inhomogeneous terms of a non-exchange nature.
LIFSCHITZ POINT
With the example of the simplest thermodynamic potential (35.1) with gradi-
ent terms, we saw that as the temperature is lowered a modulated IC phase
specified by the wave-vector "'0 arises from the initial phase at , < 0 , whereas
an ordered homogeneous C phase arises when, > O. In both cases the phase
transition to the ordered phase is a second-order transition.
If the system is subject to external effects X (pressure, field, concentration,
etc.), the quantity r in the potential (35.1) will depend not only on T, but also
on X and the 0 ...... C phase transition will proceed along the r(T, X) = 0 line
in the (T, X) plane. Suppose that the quantity, also depends on T and X and
let us reverse the sign on the ,(T,X) = 0 line. The point (n,XL) at which
both lines intersect is called the Lifschitz point [3]. This point is determined
by solving the two equations:
(35.56)
INCOMMENSURATE PERIODICITY PHASES 317
The above features of the phase diagram in the vicinity of the Lifschitz
point, which have been derived from an analysis of a system with a one-
component OP, are general. With a two-component OP, only the shape of
the lines themselves changes, while the topological structure of the phase dia-
gram remains the same as before.
Note that the homogeneous phase 1 is not necessarily a ferromagnetic
phase. The thermodynamic potential (35.1) (or (35.25)) describes a phase tran-
sition to a state with a wave-vector near some symmetric (Lifschitz) point and
we deal with a modulation of the corresponding Lifschitz structure. If the sym-
metric point lies at the center of the Brillouin zone, then the modulated struc-
ture is a ferromagnetic one; otherwise the structure is an anti-ferromagnetic
one.
We conclude by noting that [4,5] point out the existence of Lifschitz points
due to gradient invariants of degree three made up of quantities of the form
(35.57)
(36.1 )
with the conditions u > 0 and 'Y > 0 ensuring that the extremals of the func-
tional are stable. The signs of wand (J' may be arbitrary. We have immediately
assumed that the structure is homogeneous in the x and y directions, so the
volume integral reduces to a one-dimensional integral, L being the size of the
crystal in the z direction.
The potential <I> can be minimized with the aid of the methods used in the
previous section, where the solution is tried in the form of a harmonic series
and the amplitude of the harmonics and the wave-vector are found from the
condition that <I> be minimum. This procedure, however, is valid not far from
the phase transition point. If we wished to go deeper into the condensed-phase
region, we would have to allow for a large number of harmonics. Another
approach can be used in which one has to introduce a certain simplification
into the functional (36.1) and to obtain a minimization equation with a known
exact solution. This program dates back to Dzyaloshinsky [7] who suggested
using the OP constant modulus approximation
p = const, (36.2)
e
where p is defined by the equations 1] = peicf> and = pe-icf>. In this approxi-
mation the functional (36.1) depends only on the distribution of one quantity,
namely, the phase
(36.3)
d 2 (n¢) .
~ + vsm(n¢) = 0, (36.4)
INCOMl\1ENSURATE PERIODICITY PHASES 319
(36.5)
In the absence of anisotropy, equation (36.4) has the solution ¢ = liZ, which de-
scribes a one-harmonic incommensurate structure, for example, a simple helix.
On the other hand, equation (36.4) has also a homogeneous solution, which cor-
responds to an incommensurate structure. At finite v, equation (36.4) should
describe an inhomogeneous OP distribution. We can readily find one exact
periodic solution to this equation.
To begin with, we obtain the first integral of the equation by integrating
the left-hand and right-hand sides with respect to d¢:
~(ds;)r -vcosn¢ = f,
c being the integration constant. The resulting equation may be solved ill
quadratures:
vv r"'/2 dB
(36.6)
-;-z = Jo Jl _ re2 sin 2 B'
2
¢(z) = ~am (vv )
-;-z, re , (36.7)
where re is the modulus of the elliptic function (0 :::; re :::; 1). The quantity x
must be found from the energy minimum of the system; the second integration
constant, that is, the initial coordinate z, is set equal to zero.
The energy of the state in which the OP distribution is described by the
formula (36.7) is expressed in terms of the complete elliptic first- and second-
order integrals J( and E:
<{lIe = rp
2
+ up 4 - 2p
2 ¢VV + 2p 212
IO"I-}, V (re - 2 2 4E)
- - 2 - +}"? 2 . (36.8)
nre \. n re re~
(36.9)
320 CHAPTER 9
p
~=f
311
L
-n~~~T------- __
(36.10)
The soliton picture described above is easy to see if we make use of the
asymptotes of complete elliptic integrals. The soliton lattice period diverges
logarithmically as v --t Vc (that is, as re --t 1):
4re 4
L= r.7ln~. (36.11)
Vv 1 - re 2
Since the modulation wave-vector K, ' " 1/ L, it is clear that for v --t Vc the wave-
vector tends to zero. That is to say, the wave-vector of the incommensurate
phase tends to its commensurate value "'0'
We wish to illustrate the evolution of the spatial distribution of the OP
described by the solution (36.7), with an example of a modulated SS structure.
Let the wave-vector of the helix be directed along the principal crystal axis (z
INCOMMENSURATE PERIODICITY PHASES 321
axis) and let the magnetic moments lie in the basal plane (x, y). The magnetic
anisotropy in this plane is described by an invariant of the form M+ + M:!..
The nth-order anisotropy singles out n equivalent directions in the basal plane,
the atomic magnetic moments tending to adjust themselves to these equivalent
directions. If there is no anisotropy, the magnetic moments rotate uniformly
during displacement from layer to layer in the z direction. If the anisotropy is
infinitely large, the magnetic moments approach one ofthe n specific directions.
In keeping with the solution described by the function am ({i z, re) (Fig. 9.6),
the orientation of atomic magnetic moments at v values sufficiently close to
Vc (that is, when sufficiently removed in temperature from the point Tc) is
shown in Figure 9.7. There are seen to arise stacks of ferromagnetic planes
whose magnetic moments are successively aligned with the anisotropy axes.
The transition between two stacks occurs over a small length compared to L
and corresponds to a soliton. For v = Vc there is only one commensurate-phase
domain of infinite length.
Fig. 9.7. Spiral structure in the presence of an anisotropy in the basal plane.
¢(z) (36.12)
p=l n cosh P""'j(
by virtue of the expression (36.7), where
K=--
7rVv (36.13)
nreI<
322 CHAPTER 9
is the wave-vector of the structure. Thus, in full accord with the results of §35,
where a two-component OP system without a Lifschitz invariant was consid-
ered, extra harmonics e i (n±l)I<z and multiple counterparts arise alongside the
fundamental harmonics eil<z when an anisotropy is present in the basal plane.
i 1----___
oI--------~------~~
o,~ i
x~
o os i
_l~_L_
J __
---.-1-------- ;r' = 1
Next we need to establish the range of variation for the parameter of the func-
tional (36.3) over which the solution (36.7) of the minimization equation is
stable. The solution of interest arises from the condition that the first varia-
tion of the functional be equal to zero: 6<1> = O. The stability of the solution
determines the sign of the second variation: 02<1> > O. We write out the expres-
sion for the second variation in the form [8]:
(36.15)
324 CHAPTER 9
where
~ ( d
D</></> = 'Yp2 dz
)2 - Wpn cos n¢, (36.17)
D~ p</> = 2(1P dz
d d¢ d 2 n 1 .
+ 2'YP dz dz - wn P - sm n¢. (36.18)
In the expression (36.17) we have used the symbolic notation (djdz)2, which
implies
1
-D
2 H
=
(12
r - -
'Y
+ -L1 J (d¢
dz'Y -
~
+ -(1)2
'Y
> O. (36.19)
(36.21)
where "po is a function with a 'poor' asymptote at z -+ ±oo, and 8¢' is bounded
within these limits. The contribution to 82 <1>, with which we are concerned,
INCOMMENSURATE PERIODICITY PHASES 325
Consider now the eigenfunctions of the operator - fi.r - v cos <jJ. The corre-
sponding eigenvalue equation
(36.24)
1/Jo = dn u l
tio
ti du
-2-'
dn u
(36.26)
(36.29)
This expression should replace the second term on the right-hand side of the
expression (36.23), which now may be represented as
1
2L J A 1
dzD¢¢(O</»2 = 'Yp2I J dzo</>' (-d
2
dz 2 - v cos</>)
2<1>Ic
o</>' + 2"1 aare 2 (ore)~.
?
(36.30)
Similarly, we calculate the contribution to P<I> by the mixed variation
1
2L
J 2 n - 2 1
dzDp¢o¢>Op = -2'Yp v-n-I
A J. dz Sll1 </>o</>' + a 2 <1>IC
apore opore. (36.31 )
Thus we have arrived at the result that the variation of a functional with
a linear asymptote with respect to z is equivalent to the variation of <I> with
respect to the parameter reo The derivatives a 2<1>IC/are 2 and a 2<1>IC/areap can
be calculated using the explicit expression (36.8) for the incommensurate-phase
energy:
a 2 <1>IC = 4v 'Yp2 1 E (36.32)
are 2 n 2 re 4 (1 - re 2 ) J{'
a<l>IC 4 n - 2 E2 - J{2(1 - re 2 )
- - = - V'YP--
2
(36.33)
areap n re 3 J{2(1 - re 2 ) .
The problem then arises of diagonalizing the quadratic form with respect
to o</>' and op. This problem is solved by the linear substitution o</>' = o¢+ fOp
and the requirement that the function f obey the inhomogeneous differential
equation
d2
( dz ) n- 2
2 + v cos</> f + -;:;pvsin</> = 0,
(36.34)
f = --dn
n - 2 u ( u + Cl + C2
np
1 0
u
-du
2 -) '
dn u
(36.35)
INCOMMENSURATE PERIODICITY PHASES 327
has the requisite periodicity f(u + 2/C) = f(u) and the required periodicity of
the functions snx and cnx if we choose C2 = -l{(1- re 2)/E.
On calculating all the integrals, the completely diagonalized quadratic
form (36.15) for the incommensurate phase is
The new variable Ore results from the diagonalization of the integral-free terms
(36.30) and (36.31) and is related to Ore by the equation
_ n - 2 E2 - l{2(1 - re 2 ) op
Ore = Ore - -2- El{ -p .
For the phase to be stable, all the coefficients of the quadratic form (36.36)
must be positive. We first transform the integral term in equation (36.36) by
expanding oJ in Lame equation eigenfunctions:
OJ = J dlo:(l)1jJ(l).
Then
(36.39)
since min l = 0 for equation (36.24). Therefore, the state of the soliton lattice
is stable relative to the phase fluctuations oJ. Also
12 3 4 7r 2 2 0" 2
re In re' = 32(n-2) (-r),' (36.41 )
328 CHAPTER 9
(36.42)
As is seen from equation (36.41), the stability of the soliton lattice with
respect to the variation of the amplitude p is perturbed at values of re less
than unity (but in a logarithmically close vicinity of unity). This means that a
continuous transition to a commensurate phase cannot be realized for re = 1.
By comparing the energies of the incommensurate and commensurate phases,
it may be shown that a first-order transition between these phases occurs at
re' ~ 1. The quantity re' is determined by the equation [8]:
(36.43)
The IC-C phase transition, which is described by the soliton mechanism, oc-
curs owing to the competition of two terms in the thermodynamic potential
(36.3). One of them, containing a phase gradient, tends to bring about an
inhomogeneous OP distribution in the system, that is, a modulation with the
wave-vector KO. The other term, allowing for the anisotropic energy, tends to
make a homogeneous structure with the phase ¢, which is determinable from
the relation cos n¢ = ±1, where the sign is chosen depending on the sign of the
parameter v.
We have assumed that the starting thermodynamic potential (36.1) or its
simplified expression (36.3) describes a phase transition with a wave-vector
near some Lifschitz point of the Brillouin zone. We call a phase with a wave-
vector at such a point a homogeneous phase. If we consider the behavior of
our system as the temperature is varied, we can see that when T < Tm a
modulated structure arises which has a wave-vector KO' As the temperature is
elevated the KO decreases gradually and, as a result, a homogeneous structure
arises of which the period is commensurate with the initial-phase period since
it is determined by a Brillouin-zone Lifschitz point. Therefore, a homogeneous
phase is a commensurate phase in the aforementioned sense and an inhomo-
geneous phase is, generally speaking, incommensurate. The chief result of the
previous subsection is that the commensurate phase is stable in some region of
the parameter KO.
INCOMMENSURATE PERIODICITY PHASES 329
Dzyaloshinsky was the first to adduce arguments supporting the idea that
variations of the wave-vector with temperature should occur not smoothly but
as discontinuous jumps between individual commensurate values in which the
corresponding phase is stable. Such stepwise variation of the wave-vector came
subsequently to be called the devil's staircase [10].
We have seen what role the anisotropic Dzyaloshinsky invariant cos n¢J
plays in transitions between IC and C Lifschitz phases. The integer n is a
characteristic of a two-dimensional IR at a Lifschitz point, and for all the four-
teen Bravais lattices it may be equal only to 3,4,6,8,12 (§14). Let us consider
some non-Lifschitz point that lies on a symmetric direction, say Ii, = flhl,
where 0 < fl < ~. In a crystal with a finite number N of cells, closed ac-
cording to the Born-Karman condition, the wave-vector runs over the values
IC = (lIN)(27rla), where 0 < I < N - 1; thus the parameter fl runs over ra-
tional values min only. Thus we take the wave-vector Ii, = flh 1 with the only
component
27rm
IC = - - (m < n) (36.44)
na
(this vector correpsonds to a point in common on the flh 1 line) and consider the
corresponding two-dimensional IR of the parent-crystal space group G" (the IR
is realized on the arms Ii, and -Ii,).
It is normally assumed that the wave-vector symmetry does not change on
the line and that the group G" is the same for all fl. This is really the case if
we consider the wave-vector group poil'lt symmetry alone. But the translation
subgroup will be different as a function of the number fl = min. The trans-
lation subgroup contains minimal translations by the quantity na. The basis
functions of the IR on the arms Ii, and - I i , vary with varying translation vector
tn according to the law: W" '" exp( i/d) and '1/J-" '" exp( -ili,t). The quantities
('1/J,,)n and ('1/J_,,)n are invariants under translations for the wave-vector (36.44),
so the invariants should consist of powers of the OP's 'T}n and ~n. vVhatever
the rotation group G", only two invariants may be set up from these powers:
(36.45)
In form, these invariants are the same as those existing for two-dimensional
representations at Lifschitz points; however, the number n may be arbitrary
for non-Lifschitz points. The difference between these invariants lies in the
circumstance that the number n characterizes the I group (c n or cnv ) rotation
axis for Lifschitz points and minimal translations for non-Lifschitz points.
330 CHAPTER 9
In moving along the line in the Brillouin zone, the quantity n varies from
point to point according to the wave-vector (36.44). Alongside the variation
of n, there is a change in the translational symmetry of the group G" and in
the energy invariants. The latter circumstance makes the energies of phases
with close wave-vector values different, thereby ensuring the energetic favor-
ableness of certain wave-vectors. It is on the strength of these considerations
that Dzyaloshinsky arrived at the conclusion that the wave-vector may vary
stepwise between sequential rational values with varying temperature [11].
All the arguments concerning the stability of phases with a Lifschitz wave-
vector may be applied to a phase with a non-Lifschitz but commensurate vector.
With varying KO the modulation wave-vector will abruptly pass to a value that
corresponds to a commensurate phase with some value of min and this phase
will be stable over a certain interval of functional parameters, then there will be
a jump to a sequential rational min value, etc. In each cascade of these phase
transitions certain values of min occur that are dependent on the parameters
of the thermodynamic potential.
VVithin the framework of the mean-field approximation for an Ising mag-
net with positive values of exchange integrals for nearest neighbors and with
negative values for next nearest neighbors, Bak [12] constructed, with the help
of a computer, phase diagrams on the temperature-exchange integral plane
(Fig. 9.10). It has turned out that a large number of modulated phases ex-
ist. These are characterized by the values nlm = i, i, i, ... , /7 and each of
them has an individual region of existence. If we vary the temperature (with
the exchange integral being fixed), we will obtain a sequence of phases with
some of these nlm values. This is a model example of the devil's staircase. A
variety of systems exist for which the devil's staircase (that is, a discontinuous
wave-vector variation) has been observed experimentally [13-17].
A number of physical factors should destroy the devil's staircase, espe-
cially at large values of n. Here belong primarily cooperative motions in the
soliton lattice, which describes the structure of the incommensurate phase near
the boundary of its stability. Domain boundaries, described by solitons, are
infinite planes perpendicular to the wave-vector. Elementary excitations of
such boundaries are phasons, that is, flexural waves that propagate in the
plane of the boundaries, and also oscillations of the boundaries themselves rel-
ative to one another. Allowance for these excitations leads to shifting of the
commensurate-phase stability boundaries. Phasons narrow the stability range
of the incommensurate phase [12].
INCOMl'vIENSURATE PERIODICITY PHASES 331
Fig. 9.10. Phase diagram for ANNNI model [12]. .:71 > 0 is the
exchange integral between nearest neighbors in the direction of the
modulation; .:72 < 0 is the exchange integral between next nearest
neighbors.
STOCHASTIC REGIME
Recently it has been found that the physical models described by an energy
functional of the type (36.3) (before proceeding to a continual limit in which
the lattice parameter tends to zero) exhibit peculiar behavior near the IC-C
phase transition boundary. Along with the commensurate and incommensurate
structures, some chaotic structures with random distances between solitons
(describing in the incommensurate phase the wall between the domains of the
commensurate phase) may exist in these models. This was first disclosed by
Aubry (see his review lecture [18]) in the Frenkel-Kontorowa model [19,20]
for an atomic array in a periodic potential, in [21] in the three-dimensional
anisotropic Ising model allowing for the interaction with next nearest neighbors
(the so-called ANNNI model), in the discrete </;4 model [22,23], and in the
discrete Peierls model [24,25].
332 CHAPTER 9
37. Multi-N.-structures
(37.1)
where the amplitudes '1]1, '1]2 and the wave-vectors ~1, ~2 are found from the
condition that the potential be minimal, ~1 and ~2 lying in the (x, y) plane.
The problem reduces to one of minimizing the form
(37.3)
"1 and "2 being the moduli of the vectors ~1 and ~2.
The term -a1 allows the wave-vectors to be referred to the crystal axes;
a favorable position of the vectors at a1 > 0 is along the bisectors of the x and
y directions, for a1 < 0 the favorable position is along x or y. Assume, for
definiteness, that a1 > 0, so that for the minimum (37.3) we have "rx "'r
= y
and "~x = "~y, where the term proportional to a1, involved in the expression
(37.3), cancels out. Minimizing with respect to the moduli of "1 and "2, we
determine the values of these vectors:
(37.4)
334 CHAPTER 9
(37.5)
°
with 1~ structure.
To make the multi-~-structure energetically favorable ("11 =1= and "12 =1= 0),
it is therefore necessary to make the coefficient of "1f"1~ in the effective potential
(37.5) negative. This can be achieved only by including extra terms into the
functional (37.1), terms that would be of order "14. These should inevitably
involve spatial derivatives of "1. For example, the presence in the energy of an
invariant of type W"12( {)2"11 ()x{)y) 2 changes the coefficient of 17f"1~ by an amount
'" wKiK~. If this quantity is of the order of u, the overall coefficient of 17i17~
is likely to reverse sign at an appropriate value of w. In this case the multi-
~-structure will be favorable. However, since the wave-vectors of modulated
MULTI-~-STRUCTURE IN CeA12
(37.6)
with a small value of the parameter p,. Thus the wave-vector star in the struc-
ture is a 24-arm star. A 3~-structure is assumed to occur in each domain
corresponding to the wave-vectors (37.6) which group around the point ~o.
INCOMMENSURATE PERIODICITY PHASES 335
<I> = ~ Jdr{ 117 2 + U1]4 + (-y + a17 2)( 1]; + 1]; + 17;) + (r1 + a11]2)( 1]x + 1]y + 1]z )2+
+(a + )..1]2)(1]xx + 1]yy + 1]zz)2+
+( a1 + )..1772)( 1];x + 1];y + 1];z - 21];y - 277;. - 21];z)+
+(a2 + )..21]2)1]xy.(1]x + 77y + 1].)+
+(a3 + )..31]2)(1]xxx + 1]yyy + 1]zzz)(1]x + 1]y + 1]z)}, (37.7)
s(r) = 1]exp(i~or)(111).
(37.8)
exists; the vectors "'1, "'2, and "'3 are determined by the empirical relations
(37.6). On substituting equation (37.8) into the functional (37.7) and carrying
out the integration, we obtain the expression for non-equilibrium energy in a
form isomorphic to a homogeneous phase with a three-component OP:
(37.9)
r (37.10)
-2a·
The stability of the phase specified by 1]f = 1]~ = 1]5 ::P requires that °
the coefficient of the term (1]r + 1]~ + 1]5)2 be positive and the coefficient of
(1Jf1]~ + 1]f1]5 + 1J~1]5) negative (see §18). Thus the following inequalities should
be fulfilled for the 3", structure to be favorable:
(37.11)
336 CHAPTER 9
MULTI-",-STRUCTURE IN Nd
The experimental situation in the case of Nd is much richer than that for
CeAh (Fig. 9.11) [32]. Nd pertains to a group of rare-earth metals having
an h.c.p. structure in which the atoms occupy the positions 2(c) and 2(a) of
the group D~h (Fig. 9.11). At TNl = 19.9 K there arises on position 2(c)
atoms a longitudinal spin-wave structure with modulation wave-vectors ±"'1 =
J-lb 1 , ±~2 = J-lb 2 , ~3 = J-l(b 2 -bt). These correspond to six satellites that lie in
the basal plane near each crystal reciprocal-lattice site. As the temperature is
lowered down to TN2 = 19.3 K a second transition occurs which is characterized
by a small deviation of the magnetic moments and modulation wave-vectors
from the symmetric directions in the basal plane. On the neutron diffraction
patterns this is manifest in the splitting of magnetic satellites. In the following
we restrict our attention to the temperature range 7.5 K< T < 19.9 K in which
a third phase transition comes about at TN3 = 7.5 K.
In the past few years three versions of magnetic structure have been pro-
posed to decipher neutron diffraction patterns of Nd. It was originally assumed
[32] that a 3",-structure (with wave-vectors "'1, ... ''''6) takes place in this case
which is accompanied by distortions of a lattice of the same periodicity. How-
ever, an ad hoc experiment refuted this assumption [33]. Then a 1",-structure
model was suggested [34] in which the large number of magnetic satellites in
the structure is attributed to the presence of domains. This model, however,
does not account for certain temperature dependences that are observed exper-
imentally; this applies, in particular, to the temperature behavior of satellite
splitting. Finally, the third model [35] presupposes a participation of two arms,
~1 and "'2, in magnetic structure (2",-structure) formation. As shown in [36],
the presence of an inhomogeneous anisotropy in this case leads necessarily to
the vectors ~i deviating from the symmetric directions. Thus the latter model
offers the most organic explanation of the effects observed in Nd.
INCOMMENSURATE PERIODICITY PHASES 337
1(,,)
2ie)
l[a)
.lie)
21")
(37.14)
338 CHAPTER 9
The real amplitudes "1i determine the magnitude of the corresponding arm
contribution, the <Pi are the angles between the directions of the magnetic arm
contribution and the axis h. We express the wave-vectors "'i in terms of the
polar coordinate system, where the angles 1/;; too are referred to the h1 axis:
"'1 = ",(cos 1/;1, sin 1/;1, 0), "'2 = K(COS(1/;2 + 2; ),sin(1/;2 + 2311'),0),
(37.16)
211'. 211'
"'3 = "'(COS(1/;3 - "3),sm(1/;3 - "3),0).
Then, using the experimental fact that the angles <Pi and 1/;i are small, we
obtain the following non-equilibrium energy expression
<I> = r( "1I + "1~ + "11) + u( "1 I + "1~ + "11)2 + v( "1I"1~ + "1I1J1 + "1~1J1)+
The first three terms in the potential (37.17) define the structures to which
the transition from the paramagnetic phase belongs. Just as in the case of
CeAI 2 , we have a 3",-structure for v < 0 and a l",-structure for v > O. In terms
of the quartic OP model the 2",-structure ("11 =
1J2 =I 0, "13 0) is unstable; =
however, as first suggested in [36], non-equilibrium energy terms associated with
wave-vector and magnetic moment orientation may render this phase favorable.
These are the last three terms in the expression (37.17). The coupling of
the wave-vectors and magnetic moments to the symmetric directions (that is,
the tendency of the wave-vectors and magnetic moments to align with the
symmetric directions) is described by the two terms next to last in the potential
(37.17), and their relative position by the last term.
INCOMMENSURATE PERIODICITY PHASES 339
4u 2 2
1/'1 = 3v'3WII: 6 (772 - "73) for "71 '# 0 (37.20)
(the expressions for 1/'2 and 1/'3 result from equation (37.20) by cyclic permu-
tation of the indices 1,2,3). The equation (37.20) tells us that only in the
2~-structure are the angles ¢ and 1/' non-zero. Thus deviation of wave-vectors
from symmetric directions is a distinctive feature of the 2~-structure in the
model (37.17).
It is a simple matter to see that on minimizing with respect to the angles,
the last three terms in equation (37.17) are equivalent to the following sextic
invariant:
- 8 ["712( rh2 - "732)2 + "722( "712 - "732)2 + "732( "712 - "722)2] , (37.21)
where
u2 ( 9WII:4) u
WII:36 +
8=- 1 32 - ' -6~1. (37.22)
/1 WII:
From a general analysis of the thermodynamic potential for a 3-component
OP ("71, "72, "73), it follows (see §§17 and 18) that the phase ("7 "7 0), which is un-
stable in the "7 4 model, is stabilized by terms of degree six. The inhomogeneous
terms of the model (37.12) lead to such sextic terms in the effective potential.
Note that, in addition to the invariant (37.21) obtained, the non-equilibrium
energy potential doubtless contains also other sixth-degree invariants, which we
have omitted at the very outset (see equation (37.14)). Adding these invariants
entails no fundamental changes in the results. It might be well to point out that
by virtue of the condition stated in equation (37.22), anomalous large values
of the parameter 8 are required for the quantity W to be sufficiently large.
As a result of minimizing with respect to amplitudes, we obtain three
magnetically ordered phases:
(37.23)
3~ : (37.24)
(37.25)
340 CHAPTER 9
If(
r-J
Fig. 9.12. Phase diagram for the functional (37.12) describing the
multi-~-structurein Nd.
Thus the foregoing thermodynamic analysis shows that only the 2~-struc
ture is accompanied by deviation of magnetic moments and wave-vectors from
the symmetric directions. Experimental observation of this deviation provides
ponderable supportive evidence for the 2~-structure in Nd. Another indirect
confirmation might be the predicted singularities in the excitation spectra of
2~-structures in the vicinity of the phase transition point [30].
Summarizing, the above analysis of whether multi-~-structures may exist
in modulated phases shows that the conditions for such structures to occur are
difficult to fulfill. This is because it is required that the parameter u (that
is, the coefficient of the fundamental 174 term) be small compared to terms of
order w,,4 or Wi ,,6, where wand Wi are the coefficients of the corresponding
invariants describing the inhomogeneous anisotropy.
In this section we deal with various aspects of the behavior of modulated struc-
tures subject to external effects. We start by considering an SS structure, which
INCOMMENSURATE PERIODICITY PHASES 341
(38.1)
The behavior of the system will depend largely on whether or not the symmetry
allows a Lifschitz invariant. In the first case the functional (36.1) describes, for
n = 1, a helix (spiral) in the field applied in the basal plane, the anisotropy
constant w needing to be identified with H. All the SS structure deformation
effects that arise as the field is varied are described by the solution (36.7) for
n = 1. Clearly, as the field is increased the SS structure deforms; when H = He
(which corresponds to re = 1) it transforms to a ferromagnetic structure. The
transition to this phase proceeds via the soliton lattice, when the change in
magnetic moment orientation along the field is extremely non-uniform.
In the case of an exchange helix, that is, in the absence of Lifschitz in-
variants, the behavior of the SS structure in an externally applied magnetic
field was earlier investigated within the framework of the Heisenberg micro-
scopic model in a number of papers [37], and it was shown that in some critical
field He the helical structure transforms to a fan phase. We now explore this
problem in terms of phenomenological theory using the functional
(38.2)
The natural crystalline anisotropy in the basal plane is disregarded. The min-
imization equation for 1] is of the form
(38.3)
L
00
1] = Apeipl;;z, (38.4)
p=-oo
342 CHAPTER 9
that is, it contains all multiple harmonics with all p values, the smallness of
the amplitude being defined by the relations
(38.5)
rAo+2uA~+H=0. (38.7)
(38.9)
while the amplitude of the zero harmonic (magnetization) satisfies the cubic
equation
2 10
[r - 3L(~)lAo + "3uAg + H = 0, (H < He). (38.11)
INCOMMENSURATE PERIODICITY PHASES 343
Combined with the equation Al = 0, this equation determines the line of the
phase transition to a homogeneous-magnetization state on the (T, H) plane.
In explicit form the phase transition line is given by the pair of equations:
The wave-vector of the IC phase near this line depends on Hand T through
the amplitudes Ao and A l :
x;
2= X;o2[ 1- P(2x;)
192u 2 22]
AlAo . (38.13)
to the wave-vector, and also by applying other physical effects, for example
deformation [38,39], to the system.
If the field is applied in an arbitrary direction to an IC structure wave-
vector oriented, say, in the z direction, then the symmetry may sometimes
admit gradient invariants along another direction, thereby leading to instabil-
ity of the phase with the original wave-vector orientation. An instability will
appear, for example, if invariants of the form
(38.15)
are allowed. The occurrence of such an invariant will cause the wave-vector to
deviate in the x direction. The character of this deviation depends on whether
the parent structure is commensurate or incommensurate. If the structure is
commensurate, an energy of commensurateness exists in the functional <1>; this
energy is described by the invariant wert + en). In the approximation of a
constant modulus of p for the OP the equilibrium distribution problem for the
o P in a field is evidently described by the functional
(38.16)
which, as a matter of fact, coincides essentially with the functional (36.3) con-
sidered earlier. The solution of the corresponding variational problem has the
form [40] (see equation (36.7))
const, HxHz:::; H'6
<P ()
x - { (38.17)
- ~am(qx, re), Hx ·Hz ~ H'6,
where
(38.18)
Thus a certain threshold value of the quantity HxHz exists up to which the
commensurate structure (this may be a normal ferro- or anti-ferromagnetic
structure or a modulated structure with a commensurate wave-vector) is pre-
served owing to the energy of commensurateness. As the threshold is elevated,
a modulation in the x direction arises in the system which is described by the
formula (38.17). From the same formula, it is seen that if the initial structure
is incommensurate, then the modulation in the x direction is induced by the
skew-oriented field in a threshold-free fashion.
We wish to indicate other fundamental possibilities of controlling the wave-
vector by a magnetic field through invariants of the type
(38.19)
INCOMMENSURATE PERIODICITY PHASES 345
d17 de
da{3Ha H{3 dz dz· (38.20)
contains the possibility that not only the value of the wave-vector but also
its direction may vary. It is clear that a skew-oriented field relative to the
vector Ko should inevitably rotate the latter since the symmetry of the crystal
plus field system no longer reflects the 'good' wave-vector orientation. For an
incommensurate structure this rotation will take place in a field as weak as
desired; for a commensurate structure it should start with some threshold.
Similar effects should arise also when the system is subject to other ex-
ternal effects, for example in the presence of a deformation. The role of these
effects may be illustrated by one good example. At TN = 21 K there arises
in the cubic crystal ZnCr2Se4 a helical magnetic SS structure in which the
wave-vector lies along the cube edge and is close to the vector Ko = (OO~) [41].
The magnetic phase transition here is of first order and is accompanied by the
occurrence of a spontaneous deformation in the direction of the wave-vector.
As the temperature is lowered the wave-vector of the helix varies in strict pro-
portion to the lattice parameter ratio specifying the magnitude of strain. This
linear relation between K and cia may be shown to arise from the symmetry
of the system. ZnCr2Se4 has a spinel structure and the magnetic atoms of
Cr occupy the positions 16( d) of the space group Ok. The magnetic modes
of an anti-ferromagnetic structure with Ko = (OO~) were enumerated in [26].
Comparison of these modes with the observed SS structure shows that the SS
structure is a weak modulation of the collinear anti-ferromagnetic structure
described by two two-dimensional IR's 1"1 and 1"2 of the vector Ko group. With
(1]11]2) denoting the OP components that transform by 1"1 and with (66) rep-
resenting those transforming according to 1"2, we can obtain a functional [or a
four-component OP. The functional contains a Lifschitz invariant that involves
346 CHAPTER 9
d'f/1 d6 d6
rr ( 6 - -'f/1- +'f/2-
dz dz dz
- 6d172
-) .
dz
(38.22)
Also there exist two invariants that contain the strain tensor component (zz:
(38.23)
For this class of solutions the problem of determining the equilibrium structure
reduces to one of minimizing the functional with a one-component OP that
interacts with the strain (zz:
(38.25)
This gives the value of the wave-vector K and the magnitude of the spontaneous
deformation (~z:
11,=-
rr + (3(~z , (38.26)
2,
o
(zz
1( rr)2 2 r
= -2L 6- 2,(3 'f/, 'f/ ~-2u· (38.27)
It is the equation (38.26) that demonstrates the linear relationship between 11,
and (~z, a relationship that is due to the invariants (38.22) and (38.23).
Eliminating the quantity (zz from the expression (38.25) leads to the renor-
malization u ---+ u - ft(6 - {,(3); as a result of this renormalization the u may
become negative, which would lead to a first-order transition. Indeed, a first-
order transition is observed experimentally. However, no unambiguous conclu-
sion may be drawn concerning the mechanism involved. This is because in the
presence of a magnetostriction the magnetic fluctuations, which are not taken
into account in this calculation, also result in the phase transition changing
from second order to first order.
It is noteworthy that the crystal ZnCr2Se4 is a rare example where mag-
netic structure modulations are due to Lifschitz invariants. In spite of the high
symmetry of this crystal, the linear invariant with respect to gradients exists
owing to the participation of two irreducible representations. Such invariants
INCOMMENSURATE PERIODICITY PHASES 347
are known to describe the energy of small relativistic interactions, so the modu-
lation which they bring about should be small too. In fact, the angle of rotation
of spins on two adjacent crystal planes in ZnCr2Se4 is 42 0 at T = 4.2 K [41].
If this angle were exactly 45 0 , this would correspond to an anti-ferromagnetic
structure with the wave-vector Ko = (OO!); from this, it is clear that the
modulation vector '" is very small. Other examples of modulated magnetic
structures due to Lifschitz invariants are the isomorphous compounds MnSi
and FeGe. These two compounds have a space group T that contains no center
of inversion. In both crystals the modulation vector is also very small [42].
Similar behavior persists also in purely structural modulations of the crystals
[43].
Table 9.1
Irreducible representations of the group Pmnb for the star of the wave-
vector "'- = lib 2
1 0 o -1 _e- i1r /-l 0 0 _ei'lr /-I 0 1 1 0 0 _e i7r /-I _e- i7r /-I 0
74
0 1 -1 0 0 _e i1r /-l _e- i1r /-l 0 1 0 0 1 _e- i1r /-l 0 0 _e i1r /-l
and X2 is assigned to the vector (-"'-3a), then the law of conservation of wave-
vector will be fulfilled in the expressions 17rXl and TJ~X2 since "'-a + "'-3a = b 2 .
When .6. --+ 0 the wave-vector "'-3a --+ 0 and the corresponding quantities Xl
and X2 need to be identified with the spontaneous polarization arising in the
direction of the c axis; from this, it follows that they should transform according
to the IR 74(E 3) [44].
From Table 9.1 we obtain the transformation laws for the quantities Xl
and X2, similar to the relations (39.3):
h = -X2,
2Xl h 3Xl= e -i1r/-l Xl, h
= X2,
2SXl
(39.5)
h2X2 = -Xl, h3X2 = -e i1r /-lX2, h2SX2 = Xl,
where Ii = .6..
It can now be readily verified that it is the quantities XlX2 and TJrXl + 1J~X2
that are invariants. Thus we have the following expression for the thermody-
namic potential describing the phase transitions in K2Se04:
(39.6)
350 CHAPTER 9
This expression contains only one coefficient that reverses sign as the tem-
perature is varied, namely, rj this is because 17 is a vrai OP that describes
a definite atomic displacement in the crystal. The other coe~cients depend
weakly on temperature. The concomitant OP x describes a polarization wave
with wave-vector (39.4).
The thermodynamic potential coefficients (39.6) may depend not only on tem-
perature, but also on the magnitude of the superstructureal wave-vector or
on the parameter 6.. We assume the following rand '" dependence of the
coefficients T and 6.:
(39.7)
(39.8)
where 6. 0 is the value of 6. at the point T = T 1 • The relation between the OP
17 and x leads to the quantity .6. depending on temperature. To determine this
dependence, we need to minimize the potential (39.6).
We rewrite the expression for the potential, separating the modulus and
the phase in the complex OP 17 and x:
(39.9)
We have
(39.10)
Minimization with respect to the phase yields 3¢+1/J = 0, so that cos(3¢+1/J) =
±1 (the sign is chosen as a function of the sign of 6). Minimizing the expres-
sion (39.10) with respect to x gives a relation that expresses the concomitant
parameter in terms of the vrai OP:
(39.11)
r )1/2 1/2
17:::::: ( - 2u '" (T1 - T) . (39.13)
INCOMMENSURATE PERIODICITY PHASES 351
Allowing for this expression, equation (39.11) now gives the temperature de-
pendence of the concomitant OP:
(39.14)
(39.15)
As the temperature is varied over the interval Tl > T > T2 the wave-vector
n, = f.lb 2 of the incommensurate phase runs over a sequence of rotational values
of f.l = min. The set of numbers m and n determines the symmetry of the
corresponding phase. The number n specifies the invariants under the initial-
phase translation group. For a two-component OP {TJITJ2} (with TJ2 = TJn
there are two such invariants: TJ? and TJ~. From these, we may set up two real
invariants:
(39.16)
whence it is clear that in the case of even numbers m the quantity TJ'1 + TJ~
remains invariant, while when m is odd this quantity changes sign. Results
of a similar analysis for all group Pmnb elements are consolidated in Table
9.2. The last column of the table contains combinations of x, y, z coordinates,
which transform in a similar fashion. The results of Table 9.2 tally with those
of [45] after an appropriate coordinate permutation associated with a different
orientation of the group D~~. The authors of that paper were the first to
indicate that macroscopic properties should exist in the incommensurate phase
which are determined by the transformation laws for the quantities TJl ± TJ2'
Table 9.2
Transformation of the quantities TJl ± TJ2 under the operation of the
group Pmnb elements
Indeed, examination of Table 9.2 shows that, for example, the quantities
(TJ"i + TJ2)fyz and (TJ"i - TJ2)PZ are invariant (for even n and odd m), where
fyz and Pz are the components of some second-rank tensor and a polar vector.
Therefore the occurrence of an incommensurate phase with condensed OP's TJl
and TJ2 may induce relevant macroscopic properties in the crystal. When the
parity of nand m changes, as can be seen from Table 9.2, the components of the
macroscopic quantities undergo changes or new quantities appear. It stands to
reason that these quantities may have appreciable values only at small n since
the invariants responsible for their occurrence have smallness", TJ n . From what
INCOMMENSURATE PERIODICITY PHASES 353
has been stated above, it follows that individual steps in the devil's staircase
of transitions may be registered on detecting macroscopic properties of the
incommensurate phase. In principle, there may also exist wave-vector 'lock-
in' mechanisms due to the occurrence of one of the macroscopic quantities
conjugate to r/i ± 17~ .
The quantity 172n cos 2n¢ may be viewed as the square of anyone of the quan-
tities (39.16) and is an invariant since these quantities themselves transform
according to one-dimensional representations.
Minimizing q, with respect to Xl and X2 yields the equation of state
(39.18)
with the help of which we can readily eliminate the variables Xl and X2 from
the thermodynamic potential and write it in the form
(39.19)
354 CHAPTER 9
where
81
W=W _ _
2
+_2.
82
8~1 8~2
We start by finding the equilibrium values of 1] and cjJ in the absence of
the fields Xl and X 2 . To lowest order in 1] the condition 8, Ph/81] = 0 yields
the result 1] ~ (-r /2u )1/2, and the condition 8<1.? / 8cjJ =
0 gives sin 2ncjJ =
0 or
cos 2ncjJ = ±l. The choice of the sign depends on the sign of the parameter w.
For w < 0 we have cos2ncjJ = -1 and the expression (39.19) yields
(39.20)
Thus for a given 1] a phase is possible in which only one of the macroscopic
quantities Xl or X2 arises spontaneously, the temperature dependence of that
phase being determined by the value of n:
(39.22)
(39.23)
Xll = (2~d
-1
+
8r
16 2 - ,
~lW
(39.24)
-1 8~n2 2n-4
X22 = (2~2) + 32u~~ 1] .
References
43. A.D. Bruce, and R.A. Cowley: J. Phys. C. Solid St. Phys. 11,3609 (1978).
44. M. Izumi, J.D. Axe, G. Shirane, and K. Shimaoka: Phys. Rev. B 15, 4392
(1977) .
45. V.A. Golovko, and A.P. Levanyuk: Fiz. Tverd. Tela 23, 3170 (1981).
CHAPTER 10
The space groups were introduced for describing the symmetry of crystal struc-
tures. A crystal structure is a spatially ordered collection of atoms. The relative
position of atoms defines the symmetry of a given crystal, that is, a set of ro-
tations, reflections, and translations that map the structure onto itself. The
crystal structure is given by a set of atomic coordinates, and these coordinates
are the only entity on which the symmetry elements act.
In describing magnetic structures, one needs already two sets of variables,
namely, the coordinates of magnetic atoms l'i and the spin direction Si. The
new variable Si permits introduction of specific magnetic groups to describe
the symmetry of magnetic structures. Take any magnetic structure M, for
example, the structure of the garnet Mn3Al2Si3012 (see Fig. 2.6). Evidently,
if we reverse all spin directions without permuting the atoms as we do so, the
energy of the system will not change and the new structure M' will have the
same symmetry as that of the parent structure M. The structures M and
M' may be viewed as two domains. An element that transforms the structure
All into M' is the spin inversion, which is normally denoted by the symbol I'.
Therefore the symmetry group of the initial phase should contain, in addition
to the usual spatial symmetry elements g, elements of the form g I'. A set of
elements g and gl' constitutes a group H = G X 1', where the symbol l' is used
to designate the spin inversion group I' = {l, I'}. The various combinations of
elements g and I' generate a set of what is known as Shubnikov groups, which
contains 1651 magnetic space groups [1].
358
COLOR SYMMETRY IN PHASE TRANSITION THEORY 359
This representation means that the first factor, which is an element of the
space group, changes the coordinate of the atom and inverts its spin. The
360 CHAPTER 10
second factor changes only the direction of the atomic spin and is due to the
extra element I' included in the Shubnikov group. Groups whose elements have
the form (40.1) are called Q groups [7].
Evidently, the action of symmetry elements of magnetic color groups may
be entirely separated into atomic coordinates and atomic spins:
gp = g(r)tP(S). (40.2)
The first factor, which acts on the atomic coordinates, will be called a base
element. The second factor will be called an element of load or simply a color
load. Groups with elements in the form of equation (40.2) are customarily
called P groups. The above color group for ZnCr2Se4 is an example of such a
P group.
The P and Q groups were obtained when introducing additional rotations
of all crystal spins simultaneously by the same angle tPi. A further complication
of the color groups is the introducion of an individual angle of rotation tP for
each spin. As a result, we obtain W type color groups [8]. Proceeding from the
way the elements of these groups operate on spins, we may, by analogy with
the foregoing, introduce the WQ and Wp groups:
(40.3)
Here the subscript r of the operator tPl.(S) denotes that the rotation of each
spin S depends on the spin coordinate r.
The introduction of color groups was motivated by the problem of classi-
fying magnetic structures [3], a necessity which predetermined the geometric
method of introducing these groups. Later it was shown [9-11] that apart from
an obvious classificatory content, the color groups have also a certain phys-
ical meaning associated with the symmetry of the thermodynamic potential
employed to describe a magnetic phase transition.
4
<Jl~ = L ria; + L Uija;a; + L (a1 a i)(aja,,)+
i=l i~j i##l
i=l
(the quantities a3 and a4 differ from a1 and a2 in the 1]i being replaced by
ei). The expression (40.7) illustrates the relation of the exchange variables ai
to the original variables 1]i and ei.
A further step is to isolate from among the remaining invariants of <Jl - <Jlo
those that correspond to the single-ion anisotropy approximation. To this end,
we need to determine the group of the single-ion anisotropy and to act with this
group on the invariants of <Jl-<Jl 0 that have remained. The single-ion anisotropy
is normally introduced by adding terms :Jzz Slz to exchange invariants of the
type J;j(SiSj). It singles out one of the directions in the crystal (for example,
the z axis), whereas all directions in the perpendicular plane remain equivalent.
364 CHAPTER 10
In the general case this approximation is evidently inadequate since the single-
ion anisotropy is determined by the surroundings of the given atom.
The equivalent directions on the site i are described by the site group Gi,
which for the Fe atoms in FeGe2 is the group D 4z . In the general case the
group Gi may have a different orientation on each site i of a given position, a
situation which corresponds to a distinct set of equivalent directions on each
site. Allowance for the single-ion anisotropy then reduces to isolating equivalent
directions that are common to all sites and are described by the intersection
of all the groups Gi. As can be readily seen, the single-ion anisotropy in the
example under consideration may be represented as a direct product of the
parent group D~~ and the group Go = D 4z : D~~ x D 4.
Acting with the resulting group D~~ x D4 on the potential <1> - <1>0, we
separate the invariants of <1>' corresponding to the single-ion anisotropy ap-
proximation:
For the quantity <1>~Iu we have written out only half the terms; the other half
can be obtained if we replace 1]i byei.
The rest of the invariants of <1> - <1>0 - <1>' are determined by the rest of the
interactions and have the form
(40.9)
Thus blocks have been isolated in the potential (40.6)-(40.9) that corre-
spond to the exchange approximation and the single-ion anisotropy approxi-
mation. In similar fashion, we may separate the invariants corresponding to
the crystalline anisotropy. The potential thus obtained may be employed to
describe magnetic and magnetoelastic properties in the exchange approxima-
tion or the single-ion anisotropy approximation. In this way, the authors of
[14] investigated the spectrum of magnetoelastic excitations in FeSn2 for three
models. The magnetic phase transition to the incommensurate phase was in-
vestigated in [17].
'With reference to the example of FeGe2 and FeSn2, we have shown how
a 'model-less' thermodynamic potential, which includes all interactions and is
COLOR SYMMETRY IN PHASE TRANSITION THEORY 365
invariant under the Shubnikov group D~~ x 1', is reduced depending on the
requirement that it be invariant under the exchange group D~~ x D4 X I'.
In practice magnetic phase transitions are investigated within the frame-
work of a particular model and the problem of the hierarchy of the various
model potentials does not arise. Instead, the necessity arises of constructing a
thermodynamic potential that corresponds to the model chosen and depends
on the OP components. The scheme of constructing the necessary potential
here is altered somewhat but continues to rest on the choice of an appropriate
color group, for example an exchange group. For the chosen group we con-
struct exchange modes, to each of which corresponds its own exchange-group
IR, that is, its own exchange multiplet [13]. Next, as usual, we use exchange
mode mixing coefficients (performing the function of OP's) to construct an
exchange potential, all the terms of which are of an exchange nature. Simi-
larly, we may construct a thermodynamic potential in the single-ion anisotropy
approximation, choosing as the parent group the single-ion anisotropy group.
Summarizing, the idea of the interrelationship between the form of the
thermodynamic potential and the type of color symmetry has proved fruitful
not only in the theory of magnetic structures, but also in the description of
incommensurate structures (§41) and superconducting phase transitions (§43)
and in the analysis of quasicrystals (§42).
the other hand, when described conventially with the help of space groups, the
orientatjonal symmetry of incommensurate structures reduces, as a rule, to a
trivial identity group.
The problem that has arisen can be solved by use of two major methods.
One of them rests on the use of space group IR's [18,19]. It is this method
that is employed in Chap. 9 to analyze phase transitions to incommensurate
phases. The representational approach is based on the use of transformation
properties of the relevant OP with a rational wave-vector "'0 close to the wave-
vector,., of an incommensurate structure. To ensure that a minimum of the
thermodynamic potential q, exists at the point "', spatial derivatives of the OP
along with homogeneous terms are included in the potential. In principle, the
representational approach enables one to describe all the observed anomalies
in physical properties that accompany the formation of an incommensurate
structure, and also to describe the properties of the incommensurate structure
itself. This approach necessitates no identification of the symmetry group CD
ofthe incommensurate structure. Nevertheless, a knowledge of the group CD of
the incommensurate structure proves useful in a number of cases, for example
in investigating the excitation spectrum of commensurate structures [20], the
light scattering on an incommensurate structure [2], etc.
The other way to solve the paradox of the Cheshire cat is by investi-
gating the symmetry of the incommensurate structures. As noted in [21-25],
incommensurate structures may be viewed as the result of projecting a com-
mensurate multidimensional crystal structure on a three-dimensional space.
Thus the problem of finding the symmetry of incommensurate structures re-
duces to one of finding the symmetry of multidimensional crystal structures,
called supercrystals. The search for the multidimensional symmetry rests on
the symmetry of the diffraction picture and is as follows. Suppose that a given
incommensurate structure is described by one wave-vector,., " b 3 (,., = ~b3'
where m and n are integers). Then the set of reflections in the reciprocal space
should be described by the scattering vector re = hb I + kb 2 + lb 3 + ~b3. This
expression may be viewed as the projection of the four-dimensional reciprocal
lattice given by the vectors b I , b 2 , b 3, and b 4 = ,., + e, where e is the unit
vector orthogonal to the vectors b I , b 2 , and b 3 , on the three-dimensional sub-
space e = O. Accordingly, the scattering vector may be rewritten in the form
re = hb I + kb 2 + Ib 3 + pb 4 . Using the inverse Fourier transform
(41.1 )
COLOR SYMMETRY IN PHASE TRANSITION THEORY 367
with rin = riO + tn, riO the position of the ith atom in the initial primitive
cell, tn the translation vector, L the number of the arm of the star {~L}
and .A the number of the basis function of the representation D{I<d v . In a
thermodynamic description of a phase transition to an incommensurate phase,
the mixing coefficients 1]L>. play the role of the OP's. The thermodynamic
potential has the following structure:
+
LlJ···,L m
A,l"",A7n m~3
368 CHAPTER 10
(41.5)
if the following two conditions are satisfied, which place constraints on the
choice of the phases <PD.:
L <PLpAp = 0
111
(mod z) (41.7)
p=1
(z being any integer). The first condition is associated with the quadratic
invariant in <1>, and the second condition is a consequence of the term 8(""L 1 +
... + t"L",) involved in equation (41.3). The conditions (41.6) and (41.7) should
be satisfied for all values of A. Therefore, the phase <PLA may be assumed to be
independent of the subscript A:
(41.8)
The number of phases <PL is related to the number of star {""L} arms.
A feature peculiar to incommensurate stars is that not all of the m arms are
independent. For example, the arms ""L and ""L+1 = -""L. Consequently,
d (d < m) independent arms J(p may be chosen. For the remaining d - m arms
of the star {""L} we may then write
d
""L = L nL"k".
,,=1
(41.10)
COLOR SYMlHETRY IN PHASE TRANSITION THEORY 369
¢L = L:
u:1
nLu"pu. (41.11)
The condition (41.13) gives a system of equations for determining the loads
tPln for each base element involved in the group GD.
The transformation law for the components of the vector 1] is determined
by the expressions (41.5), (41.11), and (41.12) and has the form
( 41.15)
370 CHAPTER 10
(41.16)
whence we have a system for finding the components of the vector "pn:
(41.17)
Using the condition (41.10) and allowing for the relation of the vectors ku and
tn to the primitive translation vectors aa and b a , we have
3 3
As a result, we obtain from equation (41.17) the expressions sought for the
components 1/Jnu:
3
( 41.21)
'Ve let all the six arms of the star {",6} participate in the formation of an
incommensurate structure. In this case the matrix nLu involved in equations
(41.10) and (41.11) has the form
1 0 0
-1 0 0
0 1 0
n= 0 -1 0
(41.22)
0 0 1
0 0 -1
COLOR SYMMETRY IN PHASE TRANSITION THEORY 371
The vectors qL involved in equation (41.17) are the rows of this matrix. Ac-
cording to the expressions (41.11) and (41.12), the same matrix determines the
vector 1/Jn" The representation (41.21) defines the matrix Cqf3:
(41.23)
while the matrix mna for the translations ai, a2, and a3 is
( 41.24)
From the expressions (41.19), (41.23) and (41.24), we find the form of the color
loads for the fundamental translations:
( 41.25)
( 41.26)
The form of the other elements can be found in a similar fashion. Thus we ob-
tain a color group for each incommensurate structure. Using the multiplication
rules (41.15), we can readily verify that the resulting set of elements generates
a group. Part of the base elements in this group may happen to be unloaded,
the rest of the elements have color loads. For example, the expressions (41.25)
tell us that for the three-arm structure all the three translations a", are loaded,
whereas in the case of the one-arm structure, as follows from the expressions
(41.26), we have only two loaded translations, al and a2. The elements that
remain unloaded are those that are involved in the symmetry group of the in-
commensurate structure when it is described in the language of space groups.
Thus, owing to the additional phase symmetry, it becomes possible to restore
the full (both orientational and translational) symmetry of incommensurate
structures. The restored translational symmetry resolves the paradox of 'the
Cheshire cat'.
372 CHAPTER 10
(a) (b)
Fig. 10.1. Examples of filling of the plane with rectangular (a) and
pentagonal (b) unit cells. The shaded area refers to 'voids' in the
filling of the plane with pentagons.
why the report [27] that the electron diffraction pattern of a rapidly quenched
Al86Mn14 alloy consists of spots whose symmetry contains five-fold axes along
with three-fold and two-fold axes (Fig. 10.2) has aroused great interest.
-- .---------~---,
0.·0····0··0
· · · , , · · o o . · g - •• g •••. g •• g· ••• o
Q' • • . II .•.•• I I · • • • II
II ••
. II .• •• "
.-
•• II
.
o· •• ·o ..
-
g·.··o
-
during one hour at 400°C resulted in a transition to the usual crystal phase
A1 6 Mn 4.
Different Penrose cell shapes may be chosen. In Fig. 10.4 we show two
versions of choosing Penrose cells. Common features of these versions are the
values of the acute angles ¢ and obtuse angles 2¢, where ¢ = 27r/5. It is
precisely this value of the angle ¢ that is responsible for the five-fold symmetry
of Penrose tilings.
There are several algorithms for constructing Penrose tilings [30-33]. Of
these algorithms, the deflation method and the matching method are the easiest
to demonstrate.
The gist of the deflation method is that one of the rhombohedra (any
rhombohedron) is broken up into rhombohedra of the same shape but of smaller
COLOR SYMMETRY IN PHASE TRANSITION THEORY 375
(a) (b)
size. For example, let us take an obtuse rhombohedron (Fig. 10.5a), label one
of its vertices by the letter P l , and break it up into rhombohedra smaller in size,
as shown in Fig. 10.5b. We call the vertex PI the deflation pole. We choose
the vertices of the smaller rhombohedra that lie opposite the vertex PI as the
poles P2 oCthe next deflation (Fig. 10.5a). Then again, we choose the vertices
opposite to the vertices P2 as the deflation poles P3 and repeat the deflation.
As a result, we come to one of the versions of Penrose tilings (Fig. 10.3).
The matching method involves different ways of decorating Penrose cells
(tile shapes) and a set of matching rules that determine the algorithm of joining
the tile shapes together. One of the versions of decoration is presented in
Fig. 10.6a [34]. By joining the rhombohedra together in such a way that the
white and black arrows coincide, a tiling such as that shown in Fig. 1O.6b can
be generated. Another version of decoration and the tiling that corresponds to
it are shown in Fig. 10.6c, d.
The Penrose tiling (Fig. 10.3) possesses no translational periodicity. Put
another way, in Fig. 10.3 we cannot isolate a fragment and generate the entire
tiling by parallel translations of that fragment. The periodicity property in
Penrose structures is fulfilled approximately; that is to say, the Penrose tiling
is quasiperiodic.
To illustrate the notion of a quasiperiodic lattice, we mark the Penrose
376 CHAPTER 10
I
t--
(a) (b)
tiles as shown in Fig. 10.7 [32]. A set of five parallel grid lines will then appear
on the tiling assembled from these cells. As can be seen, the spacing of lines in
each set is close to periodic. The separation between adjacent planes is
(42.1)
where
1 n
Xn = n+ -l-
T T
+ Cd + C 2• (42.2)
()
A o
Fig. 10.6. Examples of decorating Penrose cells and the corresponding
tilings.
v,
/\
" ',,:
,
(a) (b)
Fig. 10.7. Decoration of Penrose tilings (a) and cells (b). An illustra-
tion of the quasiperiodicity property of a 2-dimensional quasicrystal.
term in equation (42.2) describes a periodic spacing of planes, and the second
term gives a quasiperiodic addition. The spacing of the planes described by
the expression (42.2) is characterized by two interplanar spacings dL and ds
and it is easy to see that dLlds = 1 + liT. The presence of two interplanar
distances distinguishes a quasiperiodic spacing of planes from periodic and
378 CHAPTER 10
Apart from the lattice method, one employs also the projection method
[36], which is equivalent to the lattice method [37]. The gist of the projection
method is that a quasi crystalline structure may be viewed as a mapping of a
multidimensional structure, periodic in multidimensional space, onto a space
of lower dimension. For example, a one-dimensional quasiperiodic lattice may
be viewed as the projection of a strip of a two-dimensional periodic lattice,
indicated by a dashed line in Fig. 10.9, onto the AB line.
B
/
;L
\«
'<
'0.':
" /
/
/
/
V /
/ /
/
V,
/
/
JI. /
/ /
/
The quasicrystal phase in AI86Mn14, as stated above, forms from the melt as
a result of a first-order phase transition. To describe such transitions thermo-
dynamically, we may exploit the Landau theory, the major principles of which
are outlined in Chapter 1. All the distinctions of the crystallization-type tran-
sition under consideration from the crystal-crystal transitions treated in the
preceding chapters will tell on the definition of concepts such as the initial or
380 CHAPTER 10
(Oll)o.
(o-n)a
(70-1)Q
(C, L)
Fig. 10.10. Three-dimensional unit cells for constructing three-dimen-
sional quasiperiodic structures.
where the "'L are the star {K,L} arms numbering the IR's of the group Go. The
expression (42.4) for the density function ~p(r) is a model choice. This choice
can readily be understood in terms of group theory. By choosing the density
function in the form (42.4), we allow for part of the IR of the group Go. That
is, of all the possible IR's of the group Go that may participate in the forma-
tion of incommensurate phases, we consider the IR's of the translation group
only. In other words, rather than considering all the low-symmetry phases
possible, we restrict our attention to those that are described by a transla-
tional OP. The orientational symmetry in these phases is a consequence of the
translational symmetry. Phases that are characterized only by orientational
symmetry without long-range order are not considered in such an approach. In
COLOR SYMMETRY IN PHASE TRANSITION THEORY 381
the following, however, we restrict ourselves to the choice of the density func-
tion /1p in the form (42.4) since this function describes the main properties of
quasicrystals.
The thermodynamic potential invariant under the group Go has the form
Here we have not written out the sextic invariants whose coefficients are as-
sumed to be positive. The signs of the coefficients V3 and V5 may be both pos-
itive and negative. The potential (42.5) is written on the assumption that the
phase transition is described by one IR whose index is omitted. The potential
(42.5) does not depend on the direction of the wave-vector "'L (a consequence
of its invariance under the rotation group 0(3)) and depends sensIbly on the
relative position of these vectors. Examples of the various stars, differing in
the relative position of the arms "'L, are given in Fig. 10.11.
4!f
/~-
... \
\
\
-
Fig. 10.11. Examples of wave-vector star describing different types of
structure.
Figure 10.11a shows three arms of a coplanar six-arm star {"'t}, which
satisfy the condition "'L + "'M + "'N = O. The unshown three arms are opposite
382 CHAPTER 10
to the arms K.1, K.2, and K.3. In Fig. 10.llb we present five arms of a coplanar
=
ten-arm star {K.p }, which satisfy the condition K.L + K.M + K.N + K.p + K.s O.
The coplanar structure described by this star is a two-dimensional quasicrystal.
Penrose tilings may serve as a model of such a structure. Indeed, the star {K.p }
coincides with the star shown in Fig. 10.8 which underlies the construction of
tilings in the lattice method.
The non-coplanar twelve-arm star {K.o}, whose arms generate an octahe-
dron (Fig. 1O.llc), describes crystal structures with a b.c.c. lattice [39]. Two
non-coplanar stars, {K.Il and {K.j}, are related to the icosahedron portrayed in
Fig. 10.lld. Thirty arms of the star {K.Il generate the edges ofthe icosahedron
(Fig. 10.11 d). The twelve-arm star {K./} is related to the vectors directed from
the center of the icosahedron to its vertices. Both stars {K.J} and {K.j} are
responsible for the formation of quasicrystalline phases [38,40].
We pass over to the new variables
(42.6)
and minimize the potential (42.5) with respect to the angles OL. Only the terms
V3 and V5 depend on the variables OL in the potential (42.5). On substituting
equation (42.6) into equation (42.5), these terms become
(42.7)
The symbols ~ and <> here denote sUmrilation over the arms satisfying the
= =
conditions K.L + K.M + K.N 0 and K.L + K.M + K.p + K. n + "'5 0 respectively.
In principle, the quartic and sextic summands in the potential ~ include terms
that depend on the angles OL, but we neglect these terms.
We assume that the coefficients V3 and V5 are positive. Then the condi-
tions for the potential ~ to be minimal are satisfied by the following values
of OL: OL = (2n + 1)11' (n= 0,1,2, ... ). Substitution of these values into the
COLOR SYMMETRY IN PHASE TRANSITION THEORY 383
(42.8)
These expressions tell us that in the quartic model the lowest energy belongs to
a b.c.c. crystal phase, whose formation is described by the star {K.o} [13]. The
energetic favorableness of the quasicrystalline phase described by the star {K. I}
may be achieved only in the sextic model, owing to the term V5 [38,39]. Similar
lines of argument show [41] that the quasicrystalline phase may nevertheless
prove to be energetically favorable in the quartic model if one includes into the
consideration two 0 P's {1Jd and {ed corresponding to the stars {K. I} and
{K.~ }.
An analysis of the theoretical diffraction picture for a quasicrystalline
structure described by the 'edge' star has shown [42,43] that for such a struc-
ture the X-ray pattern section perpendicular to the two-fold symmetry axis
does not coincide with experiment on AI86Mn14. In the corresponding section,
experimental diffraction patterns contain reflections that do not arise in the
edge model. Therefore a model [41] with a potential of the form
15 6 30
+u[ L
L,M,N,P=l
12
+ L eLeMeNe p 8( qL + qM + qN + qp)+
L,M,N,P=l
30 12
+ L L 1JL1JMeNe p 8(K.L + K.M + K.N + K.p)] (42.9)
L,M=l N,P=l
may be regarded as the most probable one for describing the quasi crystalline
state of A186Mn14. In the expression (42.9) the symbols K.L and qL denote the
384 CIIAPTER10
arms of the stars {K.Jl and {K.~} respectively. In deriving the potential (42.9),
the following assumptions have been used.
(1) The density function components TJL and eL corresponding to the 'edge'
K.L and 'vertex' qL arms of the stars {K./} and {K.~} are equal partners par-
ticipating in the formation of the quasicrystalline structure in AIs6Mn14' The
necessity of simultaneously allowing for the parameters {TJL} and {eL} was first
noted in [40] in connection with the fact that the arm moduli IK.L I and IqL I
differ by as small an amount as 5%.
(2) The coefficients of the third-degree invariants are equal in magnitude.
The same applies to the coefficients of the quartic invariants, in which case
u > O. Such an approximation implies that we neglect the anisotropy in higher-
order invariants.
= =
We write TJL TJe i9L and eL ee ih and restrict our attention to the most
symmetric quasicrystalline phase (1£ = cPL = O. The potential (42.9) may then
be rewritten in the form
(42.10)
The {j functions contained in the potential (42.9) determine the number of the
=
corresponding terms, giving the values of the coefficients v 0.23 and w 4.39. =
1
i
(zo)
1<-- (00 )
't.z
-+------~ .... - - - - -
The structure described by the function Llp2(r) is shown in Fig. 10.13, which
has been obtained by superposition of five plane waves propagating along the
arms of the star {"'p}. A peak of each wave is indicated by a solid line. The
places where the density function Llp2(r) is at a maximum are marked with
dots.
It is readily verified that the quasicrystal structures described by the func-
tions LlP1 and LlP2 are different from each other. Thus the structure Llp2(r)
is symmetric under the replacement r -+ -r, whereas the structure LlP1 (r)
possesses no such symmetry element.
The symmetry of the quasi crystal structure can be determined in the
usual way, by picking out those elements of the initial-phase symmetry group
Go that leave the corresponding density function invariant. For example, the
quasicrystal structure described by the density function (42.11) possesses no
386 CHAPTER 10
6
~p(r) = EeL cos(K.Lr + tPL) (42.13)
L=1
is invariant under the six translations tPL -+ tPL + 271". We will treat the six
phases tPL as the unit vectors of six-dimensional space. The condition tPL = 0
assigns a point in this space, and the translations tPL -+ tPL + 271" generate
from this point a six-dimensional lattice, which may be called a simple six-
dimensional cubic lattice [38].
Let the quasicrystal phase be described by the thirty-arm star {K./}, whose
arms are parallel to the edges of the icosahedron. Inspection of Fig. 10.11 shows
that each edge of the icosahedron (each arm of the star {K.~}) may be viewed
as a sum of two vectors directed from the center to the vertices (two arms of
the star {K.~}): K.L = qM - qN. The density function may be written in the
form
15 6
~p(r) = E 'TIL cos(K.Lr + OL) = E 'TIM-N cos [(qM - qN)r+ (tPM - tPN)].
L=1 M,N=1
(42.14)
Just as in the foregoing, we may assume that the six tPL phases are unit vectors
of six-dimensional space. The values (h = 0,71" minimizing the potential with
respect to the values of V3 and V5 assign a point in this space. The translations
that generate the lattice consist of the translations tPL -+ tPL + 271" and an extra
translation tP1 -+ tP2 + 71", tP2 -+ tP2 + 71", ... , tP6 -+ tP6 + 71". The resulting set of
translations generates a six-dimensional body-centered cubic lattice [38].
Thus the quasicrystal structures described by the stars {K.J} and {~~}
differ from each other in the same way as, for example, cubic crystals with
b.c.c. and s.c. lattices do. The symmetry groups of the corresponding phases
have identical zero blocks (see §12) and different sets of six-dimensional trans-
lations. In both structures the zero block consists of twelve elements C 5 , twelve
elements Cg, twenty elements C3 , and fifteen elements C2 [38]. Knowledge of
the quasicrystal's symmetry group is needed in investigating elastic properties
of crystals [38], elastic oscillation spectra [44], and a number of other problems.
CONCLUSION
equilibrium properties, and also to explore their dynamic properties near the
ground state [38,48].
A different outlook on quasicrystals as structures corresponding to chaotic
solutions of some non-linear equations is given in [49].
By the definition of the OP, one should choose as the OP a quantity that is
equal to zero in the initial phase and is non-zero in the low-symmetry phase.
In the case of a superconducting phase transition this quantity is a function of
the form
(43.1 )
which has the meaning of the wave-function of a Cooper pair (see, for example,
[50,51]). Here", is the wave-vector of an electron, Q' and j3 are the spin variables
of an electron and C,WI the Fermi annihilation operators of an electron in the
state "'Q'.
The OP's that we treated in all the previous sections were classical quanti-
ties such as displacements of atoms from equilibrium position, atomic magnetic
moments, the probability of finding an atom of a given species in a given site,
etc. In the case of a super conducting phase transition we encounter for the first
time an OP that is of a quantum-mechanical nature. The new level of describing
the phase transition manifests itself primarily in the determination of the sym-
metry group of the initial normal phase. Indeed, the wave-function describing a
particular state of the system is given to within a phase factor. A consequence
of such indeterminacy is the invariance of the Ginzburg-Landau functional un-
der the gauge transformation group U, whose elements describe the phase shift
of the quantum-mechanical OP [51]. Apart from this, the Ginzburg-Landau
functional should be invariant under the transformations of the crystal point
group Go and under the time-reversal operation R. In the strong-spin-orbit-
coupling approximation the full group G of the initial phase may therefore be
represented as G = Go x R x U. The group G and its subgroups GD are color
groups. The necessity of employing color groups to describe ordered phases
was noted by Volovik and Gor'kov [52]. In the following we will base ourselves
chiefly on the results obtained by these authors.
COLOR SYMMETRY IN PHASE TRANSITION THEORY 391
Consider the symmetry of the OP ~O',6 (~) with respect to the substitutions
~ -+ -~ and a -+ (3. We single out two cases, which correspond to the total
spin S = 0 and S = 1, and write the expression (43.1) in matrix form [53]:
S = 0: Lio(~) = 'ljJ(~)i(Ty,
(43.2)
S = 1: Lil(~) = ((Td(~))i(Ty,
where the (TO' are Pauli matrices. The quantity ~o(~) stands for the usual
Cooper pairing of electrons with opposite spins. The parameter ~1 (~) de-
scribes the triplet pairing of electrons with parallel spins. Problems of non-
trivial superconductivity with triplet pairing as well as possible realization of
such superconductivity are discussed in [54,55]. It follows from the definition
of the OP (43.1) that
( 43.3)
Thus the scalar function 'ljJ(~) is even with respect to the replacement ~ -+ -~,
while the vector function d(~) is odd.
spin rotation group, which does not affect the atomic coordinates. Elements
of the group G act only in coordinate space. The problem of enumerating the
magnetic subgroups Mp of the group P is treated in detail in [7,56,57]' and in
the following we use some recipes of these papers to find the superconducting
color groups G D.
All the possible groups G D may be subdivided into two types according
to the way the element R is combined with the base elements. The subgroups
of the first type have the structure GD! = H x R, where H either coincides
with the group Go or is a subgroup of it. We will call ordered phases that
are described by groups of this type usual superconductors [53]. The groups
GD! involve a time reversal in the form g = {hlIRIO} (h l being an identity
element), so the corresponding phases are non-magnetic. The enumeration of
the subgroups of the first type is a trivial one and reduces to the subgroups of
the crystallographic point groups.
The subgroups of the second type G D2 contain combined elements, that
is, along with non-loaded elements g = {hIOIO}, the group GD2 contains color
elements of the form g = {hIRI¢}. According to whether or not the element R
is involved in the group G D2, the non-trivial superconducting phase may either
possess or not possess magnetic properties. In the case of magnetic structures,
the color exchange groups Alp are analogs of the groups G D2. We will make use
of this analogy in enumerating the various GD2 type subgroups by the example
of cubic crystals Go = Oh.
To begin with, consider the action of the non-loaded inversion g25 =
(h 25 1010) on the functions 1j; and d. From the evenness of the function 1j;(K,)
and the oddness of the vector function d(K,) with respect to the substitution
K, ---+ -K, (see equation (43.41)), it follows that
g251j;(K,) = 1j;(K,),
(43.5)
g25d(K,) = -d(K,).
It then follows from the requirement of invariance of the OP Li under the sym-
metry group G D2 that in the case of singlet superconductivity (S =
0) the
inversion involved in the group G D2 is non-loaded, while for triplet supercon-
ductivity (S = 1) the inversion involved has a load ei7r , that is, has the form
g25 = (h Z5 1017T) or g25 = (h 25 IRI7T). This conclusion holds good for all groups of
the second type G DZ. In enumerating these groups in the case of cubic crystals,
we may therefore restrict ourselves to the group Go = 0 (Oh = 0 x T) instead
of the initial-phase group Go = Oh. The inversion may be added at the final
stage, allowing for the peculiarities noted in the cases S = 0 and S = 1.
COLOR SYMMETRY IN PHASE TRANSITION THEORY 393
There are several methods of enumerating color groups (see, for example,
[56,57]). One of them rests on the isomorphism theorem [58] according to which
any color group GD must be isomorphic to a space group. The formulation of
the theorem contains an algorithm of finding color loads for base elements of
the space groups. According to this algorithm, to find groups G D2 isomorphic
to the group Go, a normal subgroup H should be isolated in the group Go.
Then the group may be decomposed into cosets:
(43.6)
Take the group P = (pi, ... , Pn), whose elements are elements of the color
loads. If the group P is isomorphic to the group F, then by the theorem [58],
a color load Pi may be assigned to all elements of the coset Ii = gi H. As a
result, we obtain a color group GD2 isomorphic to the group Go.
For example, letting H =
T, the group Go =
0 may be decomposed into
cosets:
(43.8)
The factor group F = OIT is isomorphic to the group C2. Next, from the
elements of the groups U and R we may construct the group P = (1, ei1f )
isomorphic to the factor group C2 • Then the color group sought, Gm = O(T),
must have the form
(43.9)
Inasmuch as the element R is not involved in the group P, the final result for
the group GD2 may be represented as GD2 O(T) x R. =
In addition to the group T, we may choose as the normal subgroup of the
group 0 also the group D 2 • Then we have the following decomposition of the
group 0 into cosets relative to the subgroup D 2 :
(43.10)
The factor group F = OlD 2 is isomorphic to the group D 3 • The load group P
may be chosen in the form
21ri
P = (1, f.2, f., Rf. 2 , R, Rf.), where f. = exp ("3)'
394 CHAPTER 10
The groups T and D2 exhaust the list of normal subgroups of the group 0, so
the rest of the groups C D2 will be isomorphic to the subgroups of the group
o : T, D4, D 3 , C3 , D 2, and C2. In finding the rest of the groups C D2
we proceed similarly, considering the normal subgroups of each of the groups
enumerated.
We construct color groups CD2 isomorphic to the group T. As the normal
subgroup we may take the group D 2. The factor group F = T / D2 is isomorphic
to the group C3 . The corresponding group of loads has the form P = (1, f, (2).
As a result, we obtain the required color group CD2 = D2 + {h 510If2}D2 +
{h 910k 2JD 2. Note that this group could be obtained as a subgroup of the
color group CD2 = O(T) x R already found. By exhausting the rest of the
possibilities, we obtain a complete list of color groups of the second type. The
maximal groups in this list have the form:
O(T) x R = [T + {h131017r}] x R,
0(D2) = D2 + {h510If2}D2 + {h 9 10If}D2 + {h14IRk 2}D2+
+{h19IRI0}D2 + {h24IRjf2}D2'
D 4 (CI) = C1 + {h41017r}C1 + {h1410IHC1 + {h15 101- t}C1+
+{h2IRI7r}C1 + {h3IRI0}C1 + {hdRlt}C1 + {h16IRI- t}C1, (43.12)
D4(C4) x R = [C4 + {h21017r}C4] x R,
(43.14)
The expressions (43.13) are easy to obtain by the projection operator method
or the invariants method (see §21). The function f(K.) is invariant under the
group Go = O. Evidently, the functions (43.13) and (43.14) are invariant under
the group G D = O(T) x R. Indeed, since what corresponds to the elements
hl - h12 in the representation A2 is the identity matrix 1, these elements as
part of the group GD should have no loads, that is, they should be of the form
g = {hiOiO}. The rest of the group 0 elements h 13 - h24 in the representation
A2 correspond to the matrix -1. In pure form without loads, these elements
do not leave the functions tf;( K.) and d( K.) invariant and may not be part of
the group G D. The minus sign may be compensated for by the color load
e i 1l"; as a result, the elements corresponding to the matrix -1 of the IR A2
are introduced into the group GD • Since the functions (41.13) and (41.14) are
real, each of the group O(T) elements may be taken in combination with the
element R. The final result is that the representation A2 corresponds to the
group G D = O(T) x R.
396 CHAPTER 10
The explicit form of the expressions for 7I"(K.) (43.13) and d(K.) (43.14)
defines the directions (on the Fermi surface) in which the conditions
(43.15)
and
(43.16)
are satisfied. Fulfillment of these conditions isolates on the Fermi surface the
symmetry-induced points and lines corresponding to the zero gap in the Fermi
excitation spectrum [53,59]. The expression (43.13) says that in the singlet
phase (S = 0) described by the representation A 2 , symmetry-induced zeros
in the gap appear at the intercept of the Fermi surface with planes such as
K.II [110]. In the triplet phase (S = 1), as can be seeen from equation (43.14),
the condition (43.16) is fulfilled at the intercept of the Fermi surface with the
cube edges K.II [100] and body diagonals K.II [111].
The occurrence of zeros in the OP excitation spectrum is due to the pres-
ence of color elements in the symmetry group of the ordered phase. This can
be readily verified if, with a specially chosen direction of the vector K., we act
with these elements on the OP. Thus for S = =
1, the element gI6 {h I6 1017l"} is
a loaded element, where h I6 is a rotation by 180 0 about the xii axis. Choose K.
parallel to the [111] axis: K.O = /\:(111). Act on the vector d(K.o) with the ele-
ment g5 = {h 5 1010} involved in the group GD = O(T) x R. From the invariance
condition g5d(K.0) = d(K.o), it follows that the vector d(K.o) should be parallel
to the [111] axis: d(I>.o) II [111]. Then we have d(K.o) =
{h I6 1017l"}d(K.o)=
-h16d(hI61>.0) = -hI6d( -1>.0) =
hI6 d(I>.0) = -d(I>.o), whence it follows that
d(I>.o) = O.
Similarly, it is easy to see that the IR Al is related to the group GD =
o x R. The singlet and triplet basis functions of the representation Al have
the form
7/!(I>.) = 1(1).), (43.17)
d(/',,) = (e",/\:x + ey"'y + ez/\:z)f(K.). (43.18)
The expressions (43.17) and (43.18) tell us that in the phase G D = 0 x R
corresponding to the IR Al there are no isolated Fermi-surface points or lines
where the conditions (43.15) and (43.16) are satisfied.
Consider the two-dimensional IR E. The basis functions that transform
according to the representation E may be chosen in the form:
We now act on the functions (43.21) and (43.22) with the elements of all the
groups CD (43.12) enumerated above and require that the functions .,p( ~) and
d(~) be invariant for each group. From the list (43.12), we thus obtain for
each group CD a system of equations for the coefficients '7i and In most of ei.
the cases the resulting systems of equations have a zero solution: '7i O. = ei =
This means that the corresponding ordered phases CD are not described by
the representation E. The rest of the CD groups allow non-zero solutions:
Recall that the analysis performed above applies to the group O. The tran-
sition to the group Oh is realized by adding the color element 925 = {h 25 iOi1r}
= = =
for S 1 and the element 925 {h25iOiO} for S O. Here the two-dimensional
representation E of the group 0 generates two two-dimensional representations,
Eg and E u , even and odd with respect to the inversion. The even represen-
tation Eg describes the transition to the singlet phase S = 0, while the odd
representation Eu describes the transition to the triplet phase S = 1.
The conditions (43.15) and (43.16) for the gap on the Fermi surface to
vanish depend also on the mixing coefficients. We write down the expressions
for the functions .,p(~) and d(~) (these expressions correspond to the mode
mixing types found (43.23)):
From the expressions obtained and from the expressions (43.21) and (43.22),
we see that the conditions (43.15) and (43.16) for the phase CD O(D2) are =
398 CHAPTER 10
fulfilled at the intercepts of the Fermi surface with the three-fold axes, that is,
= =
K- ,,(111). For the phase CD D4(D~) x R these conditions are fulfilled for
S = 0 on the lines of intersection of the Fermi surface with the diagonal planes
of the cubic cell, while for S = 1 they are fulfilled at the intercepts of the Fermi
surface with the four-fold axes. Finally, the condition (43.15) for the phase
CD = D 4( Ct) x R is fulfilled in the case of ±"x = ±"y = ±"z = ". In all the
phases except the last, the occurrence of zeros in the gap is symmetry induced
since, just as in the case of a one-dimensional IR, the zeros appear as a result of
the presence of loaded symmetry elements in the group G D. There are no such
elements in the phase G D = D 4 ( Cd x R and the fulfillment of the condition
(43.15) for S = 0 is accidental. Inclusion of concomitant parameters will lead
to the resulting expression for 'I/!(K-) not vanishing at ±"x = ±"y = ±"z = "
[53].
The ordered phases corresponding to the three-dimensional IR's FI and
F2 of the group 0 are found similarly. We write out the chief results [53]. The
functions 'I/!(K-) and d(K-) for the representations FI and F2 have the form:
1] = 17(100) FI : G D = D 4 (C4 ) x R, F2 : GD = D 4 (D 2 ) x R,
17 = 1](liO) GD = D 4 (CI ), CD = D 4 (Ct),
(43.28)
1] = 1](111) GD = D 3 (C3 ) x R, GD = D 3 (C3 ) x R
11 = 1](lff2) GD = D 3(Ct), CD = D3(Ct).
The mixing coefficients ~i coincide with the coefficients 1]i and are therefore
not written out in the expressions (43.28). The transition to the group Oh and
its IR's Flu, FIg, F 2g and F 2u is realized in trivial fashion, just as with the
two-dimensional IR E. The information contained in the expressions (43.25)-
(43.28) enables us to find readily the corresponding points and lines of the
Fermi surface where the gap is likely to vanish.
COLOR SYMMETRY IN PHASE TRANSITION THEORY 399
The possibility of the gap going to zero manifests itself in the character
of the temperature dependence of the heat capacity Ce(T). In the case of
the usual superconductivity the exponential dependence Ce(T) .-v exp( -t:J./T)
holds. The vanishing of the gap at isolated points and lines on the Fermi surface
leads to the dependences Ce(T) .-v T2 and Ce(T) .-v T3 respectively [53]. Thus
experimental heat capacity measurements may indicate the type of pairing that
occurs in a given crystal.
THERMODYNAMIC POTENTIALS
We wish to note once again a special feature of the above approach to the
symmetry analysis of ordered phases. Just as was done in §42 in the symmetry
description of commensurate-incommensurate phase transitions whose analysis
calls for color symmetry, the color groups of the low-symmetry phases are
placed in correspondence with the space group IR. It is this factor that is
decisive in the thermodynamic description of such phase transitions in terms
of Landau theory. The thermodynamic potentials for all the IR's considered
above (A 1 , E, F 1 , and F 2 ) have been explored in different sections of this book.
\Ve write them out once again and discuss some of their properties, without
duplicating the complete analysis of these potentials:
By virtue of the special choice of the basis functions (43.26) of the representa-
tion F1 and of the basis functions (43.27) of the representation F2 , the potential
here is the same in both cases.
The potential <P E describes two phases:
For the two-dimensional representation E, the 1]4 model contains only the mod-
uli of the complex OP components 1]1 = =
11711e i 1>1 and 1]2 11]2Iei 1>2, so the phases
</;1 and </;2 remain indeterminate. To lift this indeterminacy, we need to allow
for sextic invariants, for example w( 1]~1]23 + 17t31]~). For w > 0 we then have
phase 2 with </;1 - </;2 = 7r /3, while for w < 0 we have the same phase with
</;1 - </;2 = 27r/3.
400 CHAPTER 10
The potential <PF has the following peculiarities. When U2 < 0 the mini-
mum of the potential corresponds to phases with a real vector ",. Phases with
a complex vector '" = ",' + i"," occur for U2 > 0; from the condition for the
potential <PF to be minimal, we find 7]2 = 0, which implies that ",' 1. "," and
117'1 = 1","1· It is convenient to introduce the vector L = [",'","], the direction of
which determines the symmetry of the supe~conducting phase. Thus the vec-
tor L in the phase", = 7](1((2) is directed along a three-fold axis. The phase
'" = 7](liO) corresponds to the vector L parallel to a four-fold axis.
We conclude by noting that the vector L has the physical meaning of
magnetic moment direction, and all the phases described by this vector should
exhibit magnetic properties [53]. Indeed, it is easy to see from the expressions
(43.28) that in phases with a non-zero vector L the symmetry with respect to
time reversal is absent and the element R is involved only in combination with
some base group elements.
In the description of superconducting phase transitions, as can be seen
from the material outlined above, we have repeatedly employed the techniques
expounded in the previous sections of this chapter. In conclusion, we wish to
emphasize some fundamental points of modern phase transition theory that are
of a general character and do not depend on the nature of the phase transition.
In this chapter we have explored in sufficient detail the symmetry aspect of
various phase transitions: magnetic, structural, liquid-phase - quasicrystal, and
superconducting transitions. With these four examples we have investigated
the role of color symmetry in phase transition theory and shown that in addition
to t11e usual classificational meaning, the color groups have a physical content
that is determined by a particular type of thermodynamic potential. Thus the
choice of a color group type in the case of magnetic transitions is associated
with the choice of the most essential interactions and neglect of the others. In
the case of superconducting transitions it is the color loads that determine the
unusual properties of superconducting phases.
Another facet of the color groups has been demonstrated with the example
of incommensurat~ structures and quasicrystals. Owing to the color groups it
has been possible in both cases to achieve agreement between the experimen-
tally observed properties and the symmetry of these structures. Moreover, it is
the color symmetry that has helped to con1prehend the reality of the recently
discovered new type of crystalline state in quasi crystals.
An important feature that unites all the four sections of this chapter is the
relation of the colour groups to the space group IR's. This relation comes from
COLOR SYMMETRY IN PHASE TRANSITION THEORY 401
the isomorphism theorem [58], according to which the color groups treated here
are isomorphic to the space groups. As a consequence, it has been possible to
describe these transitions in terms of the celebrated Landau thermodynamic
theory of second-order phase transitions.
Finally, in the case of superconducting phase transitions we have encoun-
tered for the first time a quantum mechanical order parameter, the role of
which is played by Cooper-pair wave-functions. It is remarkable that in this
case too, one may employ the usual scheme of group-theoretic analysis of phase
transitions, created for classical OP's. The quantum mechanical nature of the
OP has manifested itself in the necessity of introducing color groups. It is
the presence of color symmetry elements that is responsible for the unusual
properties of the superconducting phases.
References
16. L.M. Corliss, J .M. Hastings, W. Kunnmann et al.: Phys. Rev. B 31, 4337
(1985).
17. Yu.A. Dorofeev, A.Z. Men'shikov, G.L. Budrina, and V.N. Syromyatnikov:
Fiz. Met. Metalloved. 63, 1109 (1987).
18. V.A. Koptsik: Group Theoretic Methods in Physics Proceedings of the
Third Seminar 1, Yurmala, May 22-24, 1985, Nauka p. 710 (1986).
19. T. Janssen, and A. Janner: Physica 126A, 165 (1984).
20. O.V. Gurin, V.N. Syromyatnikov: Fiz. Met. Metalloved. 63, 193 (1987).
21. P.M. de Wolff: Acta Cryst. 30A, 777 (1974).
22. P.M. de Wolff: Acta Cryst. 33A, 493 (1977).
23. A. Janner, and T. Janssen: Phys. Rev. B 15, 643 (1977).
24. A. Janner, and T. Janssen: Acta Cryst. 36A, 408 (1980).
25. P.M. de Wolff, T. Janssen, and A. Janner: Acta Cryst. 37 A, 625 (1981).
26. J.M. Perez-Mato, G. Madarraga, and M.J. Tello: Phys. Rev. B 30, 1534
(1984).
27. D. Schechtman, I. Blech, D. Gratias, and J.W. Cahn: Phys. Rev. Lett. 53,
1951 (1984).
28. D. Levine, and P.J. Steinhardt: Phys. Rev. B 34, 596 (1986).
29. C.H. Chen, and H.s. Chen: Phys. Rev. B 33, 2814 (1986).
30. M. Gardner: Scientific American 236, 120 (1977).
31. A.L. Mackay: Physica 114A, 609 (1982).
32. P.J. Steinhardt: Quasicrystals Preprint. Workshop on Amorphous Metals
and Semiconductors, Coronado, California, May 13-17, 1985.
33. J.E.S. Socolar, P.J. Steinhardt, and D. Levine: Phys. Rev. B 32, 5547
(1985).
34. :t\LV. Jaric: Phys. Rev. B 34,4685 (1986).
35. J .E.S. Socolar, and P.J. Steinhardt: Phys. Rev. B 34, 617 (1986).
36. A. Katz, and M. Duneau: J. Physique 47, 181 (1984).
37. F. Gahler, and J. Rhyner: J. Phys. A Math. and Gen. 19, 267 (1986).
38. P. Bak: Phys. Rev. B 32, 5764 (1985).
39. S. Alexander, and J. McTague: Phys. Rev. Lett. 41, 702 (1978).
40. P.A. Kalugin, A. Kitaev, and L.Levitov: Pis'ma Zh. Eksp. Teor. Fiz.41,
119 (1985).
41. V. Dvorak, and J. Holakovsky: J. Phys. C 19, 5289 (1986).
42. D.R. Nelson, and S. Sachdev: Phys. Rev. B 32, 689 (1985).
43. J. Wolny, and B. Lebech: J. Phys. C 19, L161 (1986).
44. M. Widom: Phys. Rev. B 34, 756 (1986).
COLOR SYMMETRY IN PHASE TRANSITION THEORY 403
45. V.E. Dmitriyenko: Pis'ma Zh. Eksp. Tear. Fiz. 45, 31 (1987).
46. M. Gardner: Mathematical Puzzles and Diversions, Bell, London.
47. N.D. Mermin, and S.N. Troian: Phys. Rev. Lett. 54, 1524 (1985).
48. T.C. Lubensky, S. Sriram Ramaswamy, and J. Toner: Phys. Rev. B 32,
7444 (1985).
49. G.M. Zaslavskii, M.Yu. Zakharov, R.Z. Sagdeev, A. Usikov, and A.A.
Chernikov: Zh. Eksp. Tear. Fiz. 91, 500 (1986).
50. M. Cyrot: Rep. Progr. Phys. 36, 103 (1973).
51. R.M. White, and T.H. Geball: Long-Range Order in Solids, Academic
Press (1979).
52. G.E. Volovik, and L.P. Gor'kov: Pis'ma Zh. Eksp. Tear. Fiz. 39, 550
(1984).
53. G.E. Volovik, and 1.P. Gor'kov: Zh. Eksp. Tear. Fiz. 88, 1412 (1985).
54. P.W. Anderson, and P. Morel: Phys. Rev. 1123, 1911 (1961).
55. P.W. Anderson, and W.F. Brinkman: Phys. Rev. Lett. 30, 1108 (1973).
56. D.B. Litvin, J .N. Kotzev, and J.1. Birman: Phys. Rev. B 26,6947 (1982).
57. V.A. Koptsik, and G.M. Chechin: Group Theoretic Methods in Physics
Proceedings of the Third Seminar, Yurmala, May22-24, 1985, Nauka 1,
695.
58. M. Hall. The Theory of Groups, Macmillan (1959).
59. E.L. Blount: Phys. Rev. B 32, 2935 (1985).
CHAPTER 11
CRITICAL INDICES
(44.1)
The first two indices have been introduced for temperatures above and below
Te, the index f3 has been determined evidently only for T < Te.
Apart from the indices for the thermodynamic quantities, one introduces
indices that characterize the behavior of order parameter correlations in the
disordered phase and at the phase transition point itself. The correlations are
assumed to decrease exponentially at T > Te and to obey the power law at
T=Te:
(1](0)17(r)} "- e- r/€, (T> Te), (44.2)
1
(1](0)17(r)) "- r d - 2 +1I ' (T = Te) (44.3)
e,
(d is the dimensionality of the space). The quantity that is, the correlation
length, increases indefinitely at T = Te and is characterized by a critical index
v:
(44.4)
The relation (44.3) defines one more correlation length index, namely, the
Fisher index 1].
404
FLUCTUATIONS AND SYMMETRY 405
(44.5)
taking into account only the fact that there exist fluctuations, that is, that the
OP distribution is non-uniform, and disregarding the interplay of the order pa-
rameters. Statistical averaging with the Hamiltonian Ho leads to the following
expression for the Fourier component of the correlator (7](O)7](r)):
(44.6)
2(3
d - 2 + TJ = -,
II
dll + 0: = 2, 'Y = (2 - TJ)II, .•• , (44.8)
which state that the critical indices should depend sensibly on space dimension.
It is readily verified that the Landau indices (44.7) for real space d = 3 do not
satisfy these relations, but do so for space d = 4. This immediately points to
the distinct character of space dimension d = 4. Indeed, modern theory [3]
shows that at d = 4 the critical indices of the system are the same as those
of a system with non-interacting fluctuations. This circumstance allows one to
exploit a specific perturbation theory in which the critical behavior of systems
in a four-dimensional fictitious space is used as the zero approximation and then
to make a continuous mathematical transition to a space of lower dimension.
On outlining the fundamentals of the contemporary theory of critical phe-
nomena, we consider in detail the role that the interaction of fluctuations plays
in the description of phase transitions in highly anisotropic systems and discuss
primarily the interaction effects due to their symmetry.
We write the most general form of the Hamiltonian H in the 1]4 model:
H = J dd x { ~ L: [r1]~ +
).,
('V' 17>Y] + L:
)"PJJII
u).,PJJII 17).,17p1]JJ1]1I } (44.9)
the order parameter here is so normalized that the coefficient of the gradient
term is equal to zero. The expression that we have written includes all the
invariants of degree four for the initial-phase symmetry group and a given IR,
so that U)"PJJII is the sum over all these invariants, that is,
In the homogeneous case this expression transforms to the usual Landau free-
energy expansion with which we were concerned when investigating the various
dissymmetric phases (the difference lies in the factor of r, which is convention-
ally introduced in writings on fluctuation theory).
In the momentum representation the expression corresponding to the
Hamiltonian (44.9) is
J
).,
J
function
z = Sp e-{3HGL = D1]e- H , (44.12)
which is a continual integral over all the values of the OP 1])., (q) between -00
(44.14)
in which the integration is carried out with respect to the volume elements
associated with the wave-vector interval qo/b < q < qo (b> 1).
The OP 1]>..(q) for q from this interval is labeled by 1]l>..(q), while for q in
the interval 0 < q < qo/b it is labeled by 1]0>..( q). Thus for any q value in the
complete momentum interval 0 < q < qo the OP may be formally represented
as a sum 1]( q) = 1]0 ( q) + 1]1 (q) of two terms in which either one term or the
other is identically equal to zero.
On integrating with respect to Df/l' the expression (44.14) determines a
Hamiltonian H'[1]o] that involves only long-wave fluctuations 1] and depends on
an arbitrary numerical parameter b. Further transformation of the Hamiltonian
H'[1]o] consists in, first, extending the momentum scale by a factor b, owing to
which the interval of new momenta q' becomes the same as the initial interval,
that is, 0 < q' < qo, and second, varying the scale of the OP 1] by a factor z.
This formal transformation, expressed by the relations
The operators R, R2, R3, ... constitute what is called a renormalization group
[5]. The behavior of a system at the phase transition point is associated with
FLUCTUATIONS AND SYMMETRY 409
H* = RH*. ( 44.18)
(44.19)
(44.20)
J
Here
HO[1]d = ~ dq(r + q2) L 1]1>. (q)1]lA( -q). (44.21)
>.
We assume that the rescaling (44.16) in equation (44.21) is fulfilled. Therefore
(44.22)
The integration in the last expression is carried out over the interval 0 < q' < qo,
while in H O[1]l] it is over the interval qo/b < q < qo. We dispose of the scale
410 CHAPTER 11
factor in such a way that the term with ql2 in the renormalization expression
(44.22) has the same form as that in the initial equation (44.21), that is,
(44.23)
After such gauging, the Hamiltonian H' will depend only on the parameter b.
The average with respect to the exponent involved in the expression (44.19)
can be calculated using perturbation theory. If we expand the exponent in a
series, the problem boils down to a calculation of the ( ... )0 averages with
respect to the quantities 'T]1. It is easy to show that the theorem that reduces
the product average to pairwise averages (Wick's theorem) is available here:
(44.24)
Each term of the series may be represented graphically according to the fol-
lowing rule: The quantity u)../J.vp is represented by a point (vertex) at which
four lines converge; the free-ended lines refer to quantities 'T]o).. (q), and the lines
that connect vertices refer to quantities (44.24), an integration being carried
out with respect to momenta that correspond to such a line.
Thus the series for (exp( -Hint})O is represented by graphs with different
number of free ends - with two, four, six, etc., part of which are connected
and part of which are unconnected. As usual, the connectedness theorem holds,
according to which the resulting series is reduced to the exponent
where S['T]o] contains only connected diagrams of the initial series. The relation
(44.19) now yields the expression for the effective Hamiltonian
The first of the graphs evidently renormalizes the expression for the free en-
ergy of fluctuations, the second describes the bare interaction of fluctuations,
the third gives the short-wave fluctuation correction for this interaction, the
subsequent graphs (with six ends) describe the triple interaction of long-wave
fluctuations which is induced by short-wave fluctuations, etc. If we confine
FLUCTUATIONS AND SYMMETRY 411
(44.28)
(44.29)
For generality, the quantity l' involved in equation (44.28) is supplied with
the index A for the case if need arises to consider a phase transition that goes
over several IR's. In analytic form these equations are
(44.30)
(44.32)
The integration is performed over the range qo/b < q < qo. Such integrals
over the d-dimensional momentum space are calculated with the aid of the
relation
where
(44.34)
is the surface of a d-dimensional sphere of unit radius.
Near T e , where r --+ 0, we may let the arguments in equation (44.33) tend
to zero, whereupon
qo/.
dqq-f-l =
q-fW - 1)
"'d 0
f
• (44.35)
(44.38)
This trinomial for u>'p!-'v assures the required symmetry with respect to all
indices. In this case the fundamental equations (44.30) and (44.31) lead to the
followng equations [9]:
(44.41)
(44.42)
(44.44)
To start with, we determine the fixed point (r*, u*) on the (r, u) plane of
the parameters of the Hamiltonian. By definition (44.18), this point is found
from the system of equations
(44.46)
This fixed point is called the isotropic or Heisenberg point. The smallness of f
implies also the smallness of the effective fluctuation interaction in the vicinity
of the point, thus justifying the perturbation theory described above, that is,
the possibility of truncating the effective Hamiltonian at the terms'" 7]4.
414 CHAPTER 11
(44.47)
where .6.r = r - r*, .6.r' = r' - r*, etc. The quantities Ar and Au are deter-
minable from the relations
( 44.49)
It stands to reason that the index1/ does not depend on the parameter b, which
is chosen arbitrarily.
Using the expression for Ar gives 1/ to first order in f.:
(44.51)
The Fisher critical index 1] is equal to zero in this approximation. Its values
can be found to second order in f.. To this end, the renormalization-group
equations must be supplemented with terms of higher order in the interaction
u. This gives
n+2 2
17= 2(n+8)2€ + .... (44.52)
FLUCTUATIONS AND SYMMETRY 415
It is easy to verify that the Gauss point turns out to be unstable for the
Hamiltonian (44.41); thus the critical behavior of the system is, in actual fact,
described by an isotropic fixed point (44.46).
The formulas (44.51) and (44.52) show that the critical indices depend only
on the number of OP components and on space dimension and are altogether
independent of the magnitude of the bare interaction.
Summarizing, recurrent renormalization-group equations may be repre-
sented as differential equations. Consider, in particular, the equation (44.44)
for the interaction parameter. Assume that the quantity b, on which we have
placed only one constraint, namely, b > 1, differs from unity by a small amount
and choose it in the form b = exp t. Allowing for the smallness of f, transform
equation (44.44) up to terms of order'" f2 to
d(u) 2
dt = -fU - 4(n + 8)K4 U , (44.53)
where the derivative du/dt has arisen from the expression (u' - u)/t with t--+
O. The quantity t, which specifies the interval of integration over short-wave
fluctuations in a single renormalization-group transformation, plays the role
of time. It is the condition for solutions to be stationary, du/dt = 0, that
determines the fixed point u* (44.46). Solving a differential equation gives a
path, originating from a starting point that corresponds to the value of u in
the bare Hamiltonian at t = 0 and to a fixed point at t --+ 00 (for an infinite
number of renormalization-group transformations).
UNIVERSAL CLASSES
with u and u' being P-component columns constituted by quantities up. The
fixed point of the transformations is defined by the equation
u* = f(u*) (45.2)
r
416 CHAPTER 11
where M" is a Px P matrix that depends only on the number ofOP components
n, space dimension ( and number b.
Renormalization-group transformations 'cause' the point {Ul, U2, •.. , Up },
characterizing the initial Hamiltonian, to 'move' in a path in the parametric
space of the Hamiltonian, the path being described by the renormalization-
group equation (45.1). In the general case the path either terminates at
some stable fixed point (if any) or goes beyond the stability boundaries of
the Ginzburg-Landau Hamiltonian. The latter possibility is discussed in the
section that follows.
The isotropic model, as we have seen, has one stable fixed point u* and all
paths originating at any given u lead to u*. In the general case this is not so. A
system may, in principle, have several stable points, each of which is attainable
within some region of the parameters of the Hamiltonian ('domain'). Another
'domain' may have a stable fixed point of its own. Since the critical behavior
of a system is determined by a stable fixed point, it will be the same for all
systems corresponding to the given 'domain'. All systems that are described by
a Hamiltonian whose parameters lie within one 'domain' are said to belong to
one and the same universal class. This is the generalized concept of universality
for second-order phase transitions. Within a given 'domain' the critical indices
do not, in fact, depend on the magnitude of the Hamiltonian's parameters.
The stability of the fixed point is determined by the eigenvalues of the
matrix .!'vI". These should necessarily be less than unity. It is convenient to
represent the eigenvalues as bAi , then the stability condition consists in the
requirement
=
Ai < 0 (i 1, ... , P). (45.4)
The correlation length index 1/ is determined by renormalizing the parame-
ter r. In the general case, if the transition goes over a reducible representation,
there are several parameters ry (1/ is the IR number). So the corresponding
linearized renormalization-group equation is a matrix equation:
(45.5)
where 1", 1'" and 1"* are columns made up of all components ry and Mr is the
corresponding matrix. We express the eigenvalues of the matrix Mr in terms
FLUCTUATIONS AND SYMMETRY 417
of b)..r, setting all Ar > O. The critical index v is determined by the maximal
eigenvalue [10]:
1
v= . (45.6)
maxA r
Vtle conclude the exposition of the general approach by commenting on the
fundamental importance of the concept of the I group in the analysis of critical
phenomena in crystals. As we saw earlier in the text, different phase transi-
tions that go over different IR's (but the same dimension!) may have identical
I groups and be specified by the same Ginzburg-Landau Hamiltonian. Ac-
cording to what has been stated above, all of these transitions should exhibit
similar critical behavior (that is, possess the same critical indices). Along these
lines the body of mathematics concerned with I groups is the mathematical
machinery employed in the theory of universality of phase transitions in crys-
tals.
Later in the text we shall illustrate the universal behavior of phase transi-
tions by several examples, but now we proceed to discuss the role of anisotropy
in simple n-component models.
CUBIC ANISOTROPY
H = JddxH L [r1]~ +
)..
(\71])..)2] + u1(L 1]~)2 + U2 L 1]1}
)..)..
(45.7)
l. (0,0), = A2 = f;
A1
1
2. (0, 36 K4)' A1 = 3"f,
f A2 = -f;
3.
(4(n:8)K4'0),
A1 = -f, A2 = -n-4
-f;
n+8
(45.9)
4-n
4. (f f(n - 4))
12K4n' 36nK4 '
>'1 = -f, >'2 =- -f
3n
418 CHAPTER 11
respectively. For each of these points the eigenvalues Ai of the matrix Mu are
indicated. For the points 1 and 2 the stability conditions (45.4) are not fulfilled,
whereas for the other two points the stability depends on the quantity n. For
n < 4 the point 3 is stable, while for n > 4 the point 4 is stable.
For n = 4 the points 3 and 4 turn out to be degenerate, one of the values
of A going to zero. To solve the stability problem, it is necessary in this case to
go beyond the framework of the first order in L Allowance for terms of order
'" (2 shows that when n = 4 it is the point 4 that is stable.
Thus for n ::; 3 the point 3 is stable, while for n ?:: 4 the point 4 is stable.
The point 3 of the Hamiltonian (45.7) coincides with the point of the
isotropic-model Hamiltonian (44.41). For this reason it is often called a Heisen-
berg fixed point. The point 4 is called a cubic fixed point, since it reflects the
cubic symmetry of an (anisotropic) limiting Hamiltonian.
The critical index v is determined, by use of the technique described in
§44, from the recurrent equation for r:
(45.10)
(45.11)
Using the formula (44.40) now yields the indices v corresponding to the Heisen-
berg (n ::; 3) and cubic (n ?:: 4) fixed points:
(45.12)
n-l
v = ~ + ~(+ ... , n?:: 4. (45.13)
Thus for n ::; 3 the critical behavior of our anisotropic system coincides with
the behavior of an isotropic system. When the cubic anisotropy is included,
the critical behavior in the case n ?:: 4 becomes different.
These results have been generalized in [12], where the isotropic fixed point
is shown to be always stable at n ::; 3 and unstable at n ?:: 4 with respect to
any anisotropic interaction,...., 1]4. Put another way, the critical indices do not
depend on the presence of an anisotropy when n ::; 3 but do for n ?:: 4.
Note that in a cubic-anisotropy system with n ::; 3 the symmetry of the
limiting Hamiltonian at the fixed point, corresponding to the phase transition
point, becomes higher. It coincides with the symmetry of the isotropic-model
FLUCTUATIONS AND SYMMETRY 419
The authors of [13-15] investigate a variety of phase transitions with the num-
ber of OP components n 2:: 4. We quote some of these to illustrate the impact
of various anisotropic interactions on the critical behavior.
To begin with, we consider the structural transition in Nb0 2 with the sym-
metry change p4 2 jmnm ~ [41 ja [16]. The unit cell in this instance increases
in size by a factor of sixteen and the resulting dissymmetric phase is described
by the wave-vector", = (~~~). The latter belongs to a four-arm star {",9}
whose arms are
'" 1 -_ (111)
442' '" 2 -_(111)
442' '"3 -_(111)
442' '" 4 -_(111)
442 .
4 4
H= Jddx{~2:[1'1]~+(V1]-,)2] +u1(2:1]1)+
-'=1 -'=1
+U2 ( 1]i 1]~ + 7J5 7J~) + U3 (7Ji 7J5 + 7J17J~ + 7J~ 7J5 + 7J~ 7J~) }. (45.14)
Magnetic cell doubling occurs along the x axis with which the atomic spins
are aligned, so that the magnetic structure is described by a one-dimensional
representation of the group G". The phase transition, therefore, is specified by
420 CHAPTER 11
H= J dd X { ~L
A=l
6
+ U1 L 171+
A=l
+u2(l7Il7l + 17~17~+ 17517n+
+u3(l7il7~ + l7il7~ + l7~l7l + l7ll7~ + l7I175 + l7ll7~+
+l7Il7~ + l7l175 + 17~175 + 17~17~ + 17~17~ + 17~175}. (45.15)
H= J ~ f.=1
dd x { [1'(17; + 77;) + (vr l7.)2 + (vr 77.)2}+
m m
+U3 '" ( 2 2
~ 17.17.' + 17.17.'
2 -2 + -2 2 + -2 -2 }
17.17.' 17.17." (45.16)
.<.'
with m = 2 and 3 respectively.
The renormalization-group equations for the Hamiltonian (45.16) have the
form [14]
The last three equations determine eight real points in (U1 U2U3) space (Table
11.1). Analysis shows [14] that only the points 5 and 6, as functionals of the
quantity m, may be stable. At the point 5 (u; = Us = 2ui) the limiting value
of the Hamiltonian is isotropic:
(45.19)
FLUCTUATIONS AND SYMMETRY 421
so this point should be called an isotropic one. This isotropic point is stable
when the number of OP components is less than four, 2m < 4. Otherwise, that
is, when 2m > 4, it is the point 6 that is stable. Inspection of Table 11.1 shows
that with m = 2 (that is, n = 4) the points 5 and 6 coalesce to first order in f.
The stability of this point is verified by making allowance for the ""' (2 terms.
Indeed, the point turns out to be stable.
Table 11.1
Fixed points of the renormalization-group transformations for
the Hamiltonian (46.16)
N ui u:; Us
1 0 0 0
2 1/40 2ui 0
3 1/36 0 0
4 1/72 1/12 0
Using equation (45.17), we find the expression for the critical index v:
v=; r
= !+![6ui+u;+2(m-1)u;]K4' (45.20)
For 2m > 4 the critical behavior is determined by the point 6 and the relation
(45.20) yields
1 3 m-1
v = 2" + 4: 5m _ 4 ( + .... (45.21)
We cite also the expression for the critical index 'f/ [14]:
(45.22)
422 CHAPTER 11
1 1
v = 2 + g{+ .... ( 45.23)
Since the atomic spins are parallel to the wave-vector, the structure is described
by a one-dimensional representation of the group G,., therefore the phase tran-
sition is specified by a four-component OP and the Hamiltonian [15]
H = J
dd x {
4
~ 2: [r1]~ + (V'1]>Y] + U1 (2: 1]n 2+
A=l
4
A=l
4
+U2 2: 1]1 + U31]11]21J31]4}' (46.2)
A=l
U~ = bf {U3 - 48u1U3J{4Inb}
with positive Ai, which are therefore unstable. Besides, there is also the
isotropic fixed point 5( 48~4 ' 0, 0), for which two values of A are equal to zero to
first order in t. To solve the problem of its stability, it is necessary to elevate
the accuracy of equations (46.3) by including terms of order,..., t 2 into these
equations. In this approximation the point 5 turns out to split up into four
fixed points, all of which are unstable.
Table 11.2
Magnetic systems for which the renormalization-group equations have no
stable fixed points [21]
CeS,TbSe,NdSe
NdTe
a-MnS,ErP,
ErSb,EuTe
with the Landau theory's criteria, the phase transition in these materials should
be of second order; however, none of these systems has stable fixed points [21]
and it is therefore assumed that the phase transition in the systems involved
should be a first-order transition. In respect of U0 2 , MnO, Cr and Eu it
is known that the magnetic phase transition in each of them is a first-order
transition.
A detailed analysis of the instability of fixed points for the magnetic phase
transitions listed in Table 11.2 has been given and possible types of first-order
phase transitions determined in [19,20].
We have considered examples of a fluctuation-induced break-down of phase
transitions to first order in systems that have no stable fixed points. However,
in anisotropic systems one may encounter situations in which a system does
have a stable fixed point but the region of its attainability is confined in the
426 CHAPTER 11
parameter space of the Hamiltonian. For some initial values of the Hamiltonian,
which lie outside the accessible region, the phase transition breaks down to a
first-order transition [22-24].
It should be noted that the conclusion that in a system where a stable fixed
point is absent, the phase transition should be of first order has been drawn
from an analysis of the renormalization-group equations in (4 - f)-space. In
three-dimensional space the behavior of such a system may turn out to be more
complicated. Thus no stable fixed point has been found in anti-ferromagnetic
N dSn3. At the same time experiment has not shown that a first-order phase
transition takes place: The OP decreases with temperature continuously and
there is no hysteresis, but the critical scattering indispensable for the phase
transition to be viewed as a second-order transition has not been detected
either. The physical nature of the magnetic phase transition observed remains
to be ascertained.
As a concrete analysis [19-21] shows, the absence of stable fixed points
is a phenomenon typical of systems with a large number v of independent
invariants of degree four (and, consequently, with a large number n). However,
as shown in reference [25] of Chapter 1, there are examples of systems with
a large number of invariants v > 3 in which a stable fixed point exists and,
therefore, there (1re no topological constraints on its existence in the general
case. It might be well to point out also that the author of the above-cited paper
has obtained a very strong result: With the number of OP components n being
greater than 4, there is only one stable fixed point (if it exists). It corresponds
to a second-order phase transition, the symmetry of the Hamiltonian at this
point being higher then the symmetry of the system itself and being described
by the group SO(n).
e
In §20 we considered the problem of two interacting OP's 1] and within the
framework of the mean-field approximation. The coefficients rl and r2 of the
terms 1]2 and e in the thermodynamic potential depend on temperature T and
some parameters X, for example pressure, concentration, etc. Two second-
order phase transition lines, determined by the equations
may intersect at some point (Te, Xc), in the vicinity of which the interaction
between the subsystems described by the quantities TJ and is substantial. e
As we saw, as a function of the magnitude of the interaction between the
subsystems, two different phase diagram types arise. These are diagrams with
a bicritical point and diagrams with a tetracritical point, which does not allow
e
the existence of phases with TJ f:. 0 and f:. 0 (Fig. 11.1). The fluctuations
may largely affect the behavior of the phase transition in the vicinity of the
intercept (Te, X e) and give rise to other phase diagram types.
T T
\ //
\ /
/
/
\/
A
,,
x
a) 0)
Fig. 11.1. Phase diagrams with bicritical (a) and tetracritical (b)
points in Landau theory.
Let TJ = {TJ1, ... , TJn} be an n-component parameter and = em} an 171-e {el, ... ,
component parameter and let them belong to two different IR's of dimensions
nand 171. There is no anisotropy in this Hamiltonian, a fact which enables
one to trace the dependence of the critical behavior solely on the numbers n
and 171. A complete analysis of the problem of the phase transition in such a
system has been performed in [26] and (independently, by invoking Wilson's
renormalization-group technique) [27]. Taking into account the fundamental
importance of this example, we reproduce here the major steps of the analysis.
Vve assume the quantities rl and r2 to go to zero in the general case at
different temperatures (which, independently, may coincide in particular cases),
so that the OP's may fluctuate in different ways and should be independently
renormalized with the help of the associated quantities Zl and Z2 (see the second
428 CHAPTER 11
r~ b
= Zrb-d + 4(n + 2)A(rt}u1 + 4mA(r2)u3] '
(47.3)
1'; = Zib- d [1'2 + 4(m + 2)A( 1'2)U2 + 4nA(rdu3] ,
u~ = Ztb-3d(ul-4[(n+8)ui+mu~]"41nb),
u~ = Zib- 3d (u 2 - 4 [(m + 8)u~ + nU5] 1>:4 In b), (47.4)
(47.6)
(47.7)
Table 11.3
Fixed points for the system oftwo coupled order parameters, described
by the Hamiltonian (47.2)
0 0 0 0 { { { -
1
1 n+8 0 0 -{ {
(n!8){ (1,0)
2 0 n+8
1
0 { -{
(m~8){ (-1,0)
32-2(n+m)-nm
3 1
n+8 n+8
1
0 -{ -{
(n+8)(m+8) { (n;~;16'0 )
1
4 n+m+8 1 1 8 n+m-4
n+m+8 n+m+8
-{
-n+m+8{ n+m+8{ (0, ~)
Up in terms of ~.
strong fluctuations of one ofthe components, so that the critical indices depend
only on the number of components n of the strongly fluctuating parameter.
At the intercept r1 = r2 both parameters fluctuate strongly, so it is nec-
essary in equations (47.3) and (47.4) to choose Zl = Z2 = b1+ d / 2 , whereupon
the right-hand side of each of the equations will contain the standard factor bf •
To first order in { the equations (47.4) have six fixed points (Table 11.3).
Examination of Table 11.3 shows that the first three points are unstable.
The stability of the points 3 and 4 is determined by the respective inequalities
For the point 5 all the three values of ui are non-zero [27]. The stability
regions of the points 3, 4, and 5 on the (n, m) plane are shown in Fig. 11.2.
The point 3, for which the interaction parameter of two subsystems is equal
to zero, refers to the critical behavior of uncoupled systems. The point 4 is
isotropic, because the limiting Hamiltonian
(47.9)
Fig. 11.2. Stability regions of fixed points 3, 4, and 5 for the Hamil-
tonian (47.2) [27]. a, line n +m = 4; b, line 2(n + m) + nm =32.
The equations (47.3) for this fixed point degenerate to one equation (44.43) of
the isotropic model with the number of components n + m. The correlation
length indices may be calculated from the formulas (44.52) and (44.53) with
n replaced by n + m. According to the relation (47.8), the isotropic point is
stable only when n + m < 4; in the case of n + m = 4 one of the Ai values
for it is equal to zero, and to solve the stability problem an analysis to second
order in (02 is required. The authors of [26] maintain that the isotropic point is
unstable for n + m = 4.
Now we have to determine the regions in the parametric space ofthe Hamil-
tonian that correspond to each stable fixed point. The positive-definiteness
region for the quartic form of the Hamiltonian (47.1) is determined by the
inequalities [26]
(47.10)
The phase diagrams for the Hamiltonian (47.1) in the mean-field approximation
(when fluctuations are disregarded) were explored in §20, where we found that
for
(47.11)
a tetracritical point occurs in a system with two coupled OP's and, conse-
e
quently, a mixed phase with 7J :I 0 and :I 0 may exist. Otherwise, when
(47.12)
the phase diagram has a bicritical point, and no mixed phase arises.
Let us ascertain which of the relations, (47.11) or (47.12), the parameters
of the Hamiltonian satisfy at the fixed points. It is convenient to introduce the
FLUCTUATIONS AND SYivIMETRY 431
!J
-f
Fig. 11.3. Phase diagram for a system of two coupling order pa-
rameters at n = m = 1. 1-5, fixed points of renormalization-group
transformations.
variables
x = (Ul - U2)(Ul + U2)-1, y = U3(Ul + U2)-1, (47.13)
(47.14)
On the (x, y) plane these inequalities cut out the stability region of the Hamilto-
nian (47.2) (Fig. 11.3). Within the ellipse the tetracriticality condition (47.11)
is fulfilled, while outside the ellipse the bicriticality condition (47.12) is satis-
fied. Thus in the mean-field approximation the system should exhibit at the
intercept of the phase transition lines (rl = r2) a first-order transition outside
the ellipse and a second-order transition inside it.
Let us see what changes these regions undergo when fluctuations are taken
into account. To this end, we need to draw the phase portraits that follow
from the recurrence relations (47.4). Given on the (x, y) plane is a schematic
portrayal of the renormalization-group transformation paths for the specific
case n = m = 1.
A stable fixed point here is the point 4. The region of its accessibility
lies between the separatrices 1-5,5-2, on the one hand, and 1-3, 3-2, on the
other. If the bare constants of the Hamiltonian (47.1) lie outside this region,
the phase pathways go beyond the stability boundaries of the Hamiltonian. In
432 CHAPTER 11
this case the phase transition from a disordered to a condensed phase (11 =F 0
e
or =F 0) will be of first order.
If the bare constants lie within the delineated region, then the situation
becomes complicated: The character of the phase transition depends also on
the locus of the point for the bare Hamiltonian in this region. If the point
lies between the separatrices 1-3, 3-2 and the curve 1-4-2, then the phase
transition will be of second order and have a tetracritical point on the phase
diagram [26]. Beyond this part of the accessibility region of the point 4, in
the vicinity of the intersection of the phase transition lines rl = 0 and r2 = 0,
segments arise on these lines that correspond to first-order phase transitions
[26]. Evidently, the size of these segments is determined by the size of the
critical region, in which the interaction of fluctuations is essential. Outside the
critical region the phase diagram structure is given by predictions of mean-field
theory. Recall that within region bounded by the ellipse there is a tetracritical
point and outside the ellipse a bicritical point.
~, t "-
.... - x
Summarizing, we note that the results presented above apply to the case
n = m = 1 only. In the general case the phase plane will be non-symmetric
and the stable fixed point will not necessarily be isotropic. Each particular
case calls for an individual phase diagram analysis.
The critical behavior of phase transitions in a number of crystals, described
by the Hamiltonian (47.2) allowing for the crystalline anisotropy, was studied
in a similar way in [28].
We now proceed to explore the role of fluctuations in other singular phase
diagram points.
The tricritical point is defined as a point on a phase transition line where
the second-order phase transitions are succeeded by first-order transitions
(Fig. 11.4). At the bicritical and tetracritical points there converge two and
four second-order phase transition lines respectively. At the tricritical point
FLUCTUATIONS AND SYMMETRY 433
three such lines converge [29,30]. This can be seen only in a phase diagram
in the space of three variables, where it is necessary to consider, along with
T and the thermodynamic force X, one more thermodynamic force, Y. An
example of a system with a two-component OP that allows the existence of a
tetracritical point was described in §25. Inspection of Fig. 7.8, which gives the
phase diagram, shows that, indeed, three second-order phase transition lines
converge at the tricritical point.
The simplest system, in which the situation depicted in Fig. 11.4 manifests
itself, is described by the potential (§25)
(47.15)
(47.17)
where the integration is carried out in the momentum interval qo/b < ql, q2 <
qo. To estimate this integral, it is necessary to pass over to the spherical system
of coordinates for the vectors ql and q2 and then, on integrating over the
moduli of ql and q2, to introduce the polar system (ql = qcosB, q2 = qsinB).
In the 2d-dimensional integral (47.17) the integration over the variable q will
then be separated out, so that
(47.18)
where the integration is carried out over the interval b-1qo < q < qo. From this
it is seen that for d = 3 a logarithmic divergence occurs in the long-wave limit
(the integral is proportional to In b), and at d> 3 the integral converges. Thus
for the tricritical point the quantity d = 3 is a critical dimension above which
434 CHAPTER 11
the energy corrections due to fluctuation interaction are finite, and the critical
behavior at this point is described by classical estimates of Landau theory.
The dimension of real space emerges on the upper critical dimension d =
3. In this boundary case the critical indices cannot alter due to fluctuation
interaction, but there arise multiplicative logarithmic additions to all power-
type temperature dependences [31].
LIFSCHITZ POINT
Consider now the critical behavior of a system that has a Lifschitz point on
the phase diagram. This point lies on the disorder-order phase transition line
and is characterized by replacement of the homogeneous phase by a modulated
phase with a zero modulation wave-vector (Fig. 9.2). The Hamiltonian of the
system tolerating the existence of such a point may be chosen in the form [32]
( 47.19)
with rJ = {rJl,' .. ,rJn} being an n-component OP. It is assumed here that the
entire d-dimensional space may be divided into two sectors: One of them,
of dimension m and indexed by I, contains a modulation wave-vector; in the
other, of dimension d - m and specified by the index c, there are no modulation
wave-vectors.
The Hamiltonian (47.19) is a gener alization of Michelson's thermodynamic
potential (35.1), written for a one-component OP and d = 1. (As is customary
in the fluctuation theory of phase transitions, a coefficient 0.5 is placed before
the quadratic terms in the Hamiltonian.)
At the Lifschitz point
(47.20)
so one should allow for the term with the square of the second derivative with
respect to the coordinates of sector I and take f3 > O.
In the momentum space the expression for the Hamiltonian (47.19) is
J
u dd ql ... dd q4c5(ql + .. . q4)(rJ(qdrJ(Q2)) (1J(q3)1J(Q4)) , (47.21)
where
(47.22)
FLUCTUATIONS AND SYMMETRY 435
Here q1 and qc are the moduli of the vector q, if it belongs to the sector I and
c respectively, so that
d
q; = E q;. (47.23)
Q=m+l
We dispose of the parameters a and b in such a way that the terms with
q~ and qJ remain invariant. This leads to the equations
(47.26)
The equations (47.24) and (47.25) lead to equations that give the renor-
malization of rand u:
(44.30)
in which the integration is carried out over the momenta: qo/b < qc < qo and
qo/a < g1 < qo; also we have denoted
m
(= 4+ - - d. (47.31)
2
436 CHAPTER 11
B(O, 0) ~ Kdm l qO
qo/b
q-f- 1dq ~ Kdm In b, (47.33)
r
*
= -f n
n+2
+8 q5
2' *
u = f
[4}' (n + 8)]-1 ,
idm (47.34)
which coincides in form with the Heisenberg fixed point (44.46) for d = 4 - f.
In the vicinity of the fixed point, the equation (47.28) may be represented in
the standard form
(47.35)
(47.36)
2(
Ar = b 1 - f
n+2 )
n + 8 In b . (47.37)
The equations (47.35) and (47.36) yield expressions for the critical indices
ec and eI
of the correlation length:
1 1 In+2
Vc = Ar = 2 + 4' n + 8 f + ... , (47.38)
1 1 In+2
VI = 2Ar = 4' + 8n + 8 f + .... (47.39)
FLUCTUATIONS AND SYMMETRY 437
References
17. Yu.A. Izyumov, V.E. Naish, and R.P. Ozerov: Neitronografiya Magnetikov
(Neutron Diffraction of Magnetics) 2, Atomizdat (1981).
18. D. Mukamel, and S. Krinsky: J. Phys. C 8, L496 (1975).
19. I.E. Dzyaloshinsky: Zh. Eksp. Teor. Fiz. 72, 1930 (1977).
20. S.A. Brazovskii, I.E. Dzyaloshinsky, and B.G. Kukharenko: Zh. Eksp. Teor.
Fiz. 70, 2257 (1976).
21. P. Bak, S. Krinsky, and D. Mukamel: Phys. Rev. Lett. 36, 52 (1976).
22. I.F. Lyuksyutov, V.L. Pokrovskii: Pis'ma Zh. Eksp. Teor. Fiz. 25, 22
(1975).
23. E. Domany, D. Mukamel, and M.E. Fisher: Phys. Rev. B 15, 5432 (1977).
24. H.H. Iacobson, and D.J. Amit: Ann. Physics 133, 57 (1981).
25. Zvi Barak, and M.B. Walker: Phys. Rev. B 25, 1969 (1982).
26. I.F. Lyuksyutov, V.L. Pokrovskii, and D.E. Khmel'nitskii: Zh. Eksp. Teor.
Fiz. 69, 1817 (1975).
27. J .M. Kosterlitz, D.R. Nelson, and M.E. Fisher: Phys. Rev. B 13, 412
(1976).
28. Yu.A. Izyumov, V.E. Naish, Yu.N. Skryabin, and V.N. Syromyatnikov:
Fiz. Tverd. Tela 23, 1101 (1981).
29. R.B. Griffiths: PhI/S. Rev. Lett. 24, 715 (1970).
30. R.B. Griffiths: Phys. Rev. B 7,545 (1973).
31. F.J. Wegner, and E.K. Reidel: Phys. Rev. B 7,248 (1973).
32. R.M. Hornreich, M. Luban, and S. Shtrikman: Phys. Rev. Lett. 35, 1678
(1975).
33. J. Sak, and G.S. Grest: Phys. Rev. B 17, 3602 (1978).
34. G.S. Grest, and J. Sak: Phys. Rev. B 17, 3667 (1978).
35. M.A. Anisimov, E.E. Gorodetskii, and V.M. Zaprudskii: Uspekhi Fiz. Nauk
133, 103 (1981).
36. J .N!. Hastings, L.M. Corliss, W. Kunnmann, and D. Mukamel: Phys. Rev.
B 24, 1388 (1981).
37. J .M. Hastings, 1.M. Corliss, W. Kunnmann, R. Thomas, R.J. Begum, and
P. Bak: Phys. Rev. B 22, 1327 (1980).
Appendix
hI-I, h2-2x, h3-2y, h4-2z, h5-3 111 , h6-3 111 , h7-3 111 , hS-3 111 ,
l
h9-3 1 1, hlO-3~ll' hll-3~/I' hI2-3 I /1 , hI3-2xy , hI4-4z, hI5-4;1,
hI6-2 xy , hI7-2gx, hlS-2 yz , hI9-4x, h20-4;1, h21-2xz, h22-4- 1
Y ,
h23-2 xz , h24-4y.
hI h2 h3 h4
1 0 0 1 0 0 -1 0 0 -1 0 0
0 1 0 0 -1 0 0 1 0 0 -1 0
0 0 1 0 0 -1 0 0 -1 0 0 1
h5 h6 h7 hS
0 1 0 0 1 0 0 -1 0 0 -1 0
0 0 1 0 0 -1 0 0 1 0 0 -1
1 0 0 -1 0 0 -1 0 0 1 0 0
h9 hl0 hll h12
0 0 1 0 0 1 0 0 -1 0 0 -1
1 0 0 -1 0 0 1 0 0 -1 0 0
0 1 0 0 -1 0 0 -1 0 0 1 0
439
440 APPENDIX
HI H2 H3 H4
1 0 0 1 -1 0 0 -1 0 -1 0 0
0 1 0 1 0 0 1 -1 0 0 -1 0
0 0 1 0 0 1 0 0 1 0 0 1
H5 H6 H7 H8
-1 1 0 0 1 0 -1 0 0 0 -1 0
-1 0 0 -1 1 0 -1 1 0 -1 0 0
0 0 1 0 0 1 0 0 -1 0 0 -1
H9 HI0 Hll H12
1 -1 0 1 0 0 0 1 0 -1 0 0
0 -1 0 1 -1 0 1 0 0 0 1 0
0 0 -1 0 0 -1 0 0 -1 0 0 -1
Index
441
442 INDEX