110
ADVASiCES IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University of Edinburgh, Mayfield Road, Edinburgh, Scotland [Received 5 April 1979] ABSTRACT Structural phase transitions can be studied by a variety of techniques; macroscopic fields, X-ray, light and neutron diffraction, and resonance methods. The information which can be obtained by these different techniques about the order parameter and susceptibility at structural phase transitions is described and compared. Landau theory coupled with the 'sot~ mode' concept provides a simple picture of many structural phase transitions in terms of relatively few phenomenological constants. A review is given of the application of this type of theory to the wide variety of different phenomena which can occur at structural phase transitions, in particular, emphasis is placed on the nature of the fluctuations within Landau-Ginzburg theory and on the role of secondary order parameters. The nature of incommensurate phases and of lock-in phase trans- itions are discussed using both the conventional lattice dynamical approach and the soft soliton theories. The parameters of the Landau theoi.~ can, in principle, be obtained from a microscopic model. We critically discuss the attempts that have been made to develop models for anharmonic crystals, order-disorder systems and various electronic instabilities. CONTENTS PAGE § I.l. INTRODUCTION. 2 § 1.2. PRELIMINARIES. (5 1.2.1. The hamiltonian. 6 1.2.2. Susceptibility. 9 1.2.3. Scattering cross-sections, ll §I.3. ORDER PARAMETERS AND FLUCTUATIONS. 13 1.3.1. Primary order parameters. 13 1.3.2. Secondary order parameters. 15, 1.3.3. Experimental determination of the order parameter. 16 1.3.3.1. Macroscopic measurements. 16 1.3.3.2. Scattering measurements. 18 1.3.3.3. Resonance methods. 20 1.3.4. Static susceptibility. 22 1.3.5. Soft modes and the dynamic susceptibility. 24 1.3.6. Measurement of susceptibilities. 29 1.3.6.1. Macroscopic measurements. 29 1.3.6.2. Scattering measurements. 30 1.3.6.3. Local susceptibilities. 37 §I.4. PHENOMENOLOGICAL LANDAU THEORY. 40 1.4.1. Single-component order parameters. 40 1.4.1.1. The Landau theory. 40 1.4.1.2. Coupling to strains. 43 1.4.1.3. Fluctuations in systems with short-range forces. 44 0001 8732/80/29(110001 $02-00© 1980 Taylor & Francis Ltd

Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

  • Upload
    dinhthu

  • View
    216

  • Download
    2

Embed Size (px)

Citation preview

Page 1: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

ADVASiCES IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110

Structural phase transitions

I. Landau theory

By l~. A. COWLEY

Department of Physics, University of Edinburgh, Mayfield Road, Edinburgh, Scotland

[Received 5 April 1979]

ABSTRACT Structural phase transitions can be studied by a variety of techniques;

macroscopic fields, X-ray, light and neutron diffraction, and resonance methods. The information which can be obtained by these different techniques about the order parameter and susceptibility at structural phase transitions is described and compared. Landau theory coupled with the 'sot~ mode' concept provides a simple picture of many structural phase transitions in terms of relatively few phenomenological constants. A review is given of the application of this type of theory to the wide variety of different phenomena which can occur at structural phase transitions, i n particular, emphasis is placed on the nature of the fluctuations within Landau-Ginzburg theory and on the role of secondary order parameters. The nature of incommensurate phases and of lock-in phase trans- itions are discussed using both the conventional lattice dynamical approach and the soft soliton theories. The parameters of the Landau theoi.~ can, in principle, be obtained from a microscopic model. We critically discuss the a t tempts that have been made to develop models for anharmonic crystals, order-disorder systems and various electronic instabilities.

C O N T E N T S

PAGE § I . l . INTRODUCTION. 2

§ 1.2. PRELIMINARIES. (5

1.2.1. The hamiltonian. 6 1.2.2. Susceptibility. 9 1.2.3. Scattering cross-sections, l l

§I.3. ORDER PARAMETERS AND FLUCTUATIONS. 13 1.3.1. Pr imary order parameters. 13 1.3.2. Secondary order parameters. 15, 1.3.3. Experimental determination of the order parameter. 16

1.3.3.1. Macroscopic measurements. 16 1.3.3.2. Scattering measurements. 18 1.3.3.3. Resonance methods. 20

1.3.4. Static susceptibility. 22 1.3.5. Soft modes and the dynamic susceptibility. 24 1.3.6. Measurement of susceptibilities. 29

1.3.6.1. Macroscopic measurements. 29 1.3.6.2. Scattering measurements. 30 1.3.6.3. Local susceptibilities. 37

§I.4. PHENOMENOLOGICAL LANDAU THEORY. 40

1.4.1. Single-component order parameters. 40 1.4.1.1. The Landau theory. 40 1.4.1.2. Coupling to strains. 43 1.4.1.3. Fluctuations in systems with short-range forces. 44

0001 8732/80/29(11 0001 $02-00 © 1980 Taylor & Francis Ltd

Page 2: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

R. A. Cowley

PAGE

1.4.1.4. Uniaxial ferroeleetrics. 46 1.4.1.5. Acoustic mode instabilities. 1.4.1.6. Interaction between acoustic modes and ferroelectrie 47

fluctuations. 48 1.4.2. Multiple-component order parameters. 52

1.4.2.1. The Landau theory. 52 1.4.2.2. Coupling to strains. 55 1.4.2.3. Improper ferroelectries. 56 1.4.2.4. Fluctuations and short-range interactions. 58 1.4.2.5. Ferroelectrics. 58 1.4.2.6. Acoustic instabilities. 59

1.4.3. Phase diagrams. 59 1.4.3.1. Tricritical points. 59 1.4.3.2. Phase transitions without symnmtry breaking. 60 1.4.3.3. Competing interactions. 61

1.4.4. Incommensurate phase transitions. 64 1.4.4.1. The simplest incommensurate phase. 64 1.4.4.2. More complex incommensurate phases. 66 1.4.4.3. Fluctuations of' incommensurate phases. 68 1.4.4.4. Secondary order parameters. 74

1.4.5. Lock-in phase transitions. 75 1.4.5.1. Phase-locked phases. 75 1.4.5.2. Soft solitons. 77 1.4.5.3. Lock-in without soft solitons. 84 1.4.5.4. Improper lock-in transitions. 84

§I.5. MicroscoPic THEORIES. 85 1.5.1. Weakly anharmonic crystal models. 86

1.5.1.1. The soft mode. 86 1.5.1.2. Models of SrTiO3. 88

1.5.2. Order-disorder systems. 91 1.5.2.1. Time-dependent Ising model. 9l 1.5.2.2. Tunnelling model. 93 1.5.2.3. Spin-phonon interaction. 94 1.5.2.4. Applications to real systems. 96

1.5.3. Electron-phonon models. 98 1.5.3.1. The formalism. 98 1.5.3.2. Jahn-Tel ler systems. 100 t.5.3.3. Charge density waves. 101 1.5.3.4. Other charge density waves. 103 1.5.3.5. Dynamic effects and temperature dependent effects. 105

L5.4. Conclusions. 106

REFEt~ENCES. 107

§ 1.1. INTRODUCTION S t r u c t u r a l phase t r ans i t i ons occur when a m a t e r i a l changes i ts c ry s t a l l og raph i c

s t ruc ture . T h e first s tudies of these t r a n s i t i o n s were m e a s u r e m e n t s of macroscop ic proper t ies ; specific hea t , d ie lec t r ic su scep t ib i l i t y or t h e r m a l expans ion , and i t was t h rough these m e a s u r e m e n t s t h a t t he t r ans i t i ons were d i scovered a n d ~nitially cha rac te r i zed . Because the macroscop ic p rope r t i e s change a t the t r ans i t i on , the i r s t u d y is of i m p o r t a n c e for the va r ious d i f ferent a p p l i c a t i o n s of these ma te r i a l s : an e x a m p l e is t he use of fe r roe lec t r ics as p iezoelec t r ic c o m p o n e n t s a n d as pyroe lec t r i c de tec tors . The re has also been a g r e a t dea l of effort d e v o t e d to e luc ida t ing the n a t u r e of s t r u c t u r a l phase t r ans i t i ons a t a more microscopic level. This w o r k began wi th the

Page 3: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions .~

careful determinat ion of the crystal s t ructure of the phases on both sides of the transition. Some of the earliest measurements were performed on BaTiO3 by Megaw (1947), Kanzig (1951) and by Shirane et al. (1955) and most recently by H a r a d a et al. (1970). The results of their investigations are illustrated in fig. I. 1, which shows the low-temperature phase of BaTiO 3, T < 393 K, in which the atoms are displaced away from the centre-symmetr ic positions they have in the high-temperature cubic phase. In ferroeleetric BaTi03 the static displacements of the a toms in every unit cell of the distorted phase are the same.

Pig. I. 1

z

DISPLACEMENT (.~) Ba 0

0 0 112 005 009

Ti -005 ;3 The distorted perovskite structure of ferroelectrie BaTi03.

At other structural phase transit ions the unit cell in the distorted phase consists of two, four or maybe more unit cells of the high- temperature phase. As an example of this type of phase transit ion we show, in fig. 1.2, the structure of SrTiO3 below the transit ion a t 110 K. In this ease the high- temperature perovskite phase is distorted by an anti-phase rotat ion of neighbouring oxygen octahedra (Unoki and Sakudo 1967).

Fig. 1.2 0

. . . . I / ; - t

@ Sr

© 0

qo

A projection down a cube axis of the distorted phase of SrTiOa as deduced by Unoki and Sakudo (1967), after Shirane and Yamada (1969).

Page 4: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

4 R . A . Cowley

These measurements showed that at least at some structural phase transitions the atoms in the distorted phase are slightly displaced away from the equilibrium positions of the high-temperature phase. This then led to the suggestion by Coehran (1960) and Anderson (1960), and earlier but largely unnoticed comments by others (Raman and Nedungadi 1940 and Saksena 1940, for example) tha t these phase transitions might be the result of an instability of the crystal against a particular normal mode of vibration of the high-temperature phase. Since this suggestion, there have been numerous measurements of the so-called 'soft modes' associated with structural phase transitions as reviewed for example by Scott (1974) and by Shirane (1974), and at least for these 'displaeive' phase transitions their existence is well established.

There are other structural phase transitions where the phase transition is associated with an ordering of some variable which is disordered in the high- temperature phase. One example of this 'order-disorder ' type of behaviour is the ferroelectric phase transition in NaNO2. In the high-temperature (paraelectric) phase the triangular NO2 groups have the N ions along either the positive or negative b axis with equal probability, but below 436 K the N ions tend to align in the same direction, as shown in fig. 1.3 (Hoshino and Motegi 1967). Since the reorientation of the NO2 groups involves very large displacements, particularly of the N ions, these motions cannot be described in terms of small displacements about the equilibrium positions. Consequently in these orde~disorder systems 'soft mode' ideas are less useful than in the displacive materials. Finally, we mention another type of ordeI~ disorder phase transition, namely that occurring in alloys. In fig. 1.4 we show the structure of ordered CuZn (fi-brass); above the ordering temperature, 741 K, the Cu and Zn atoms are randomly arranged on the body-centred cubic lattice, but below the transition they take up the regular arrangement shown.

Fig. 1.3

STRUCTURE OF NoN 0 2

0 OXYGEN

0 SODIUM

The structure of ferroelectric NaNO2 (Hish~no and Motegi 1967). Diagram taken from Sakurai et al. (1970).

Page 5: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 5

Fig. 1.4

t o l , 0 ® ..... ® ® . . . . ® ®

The st~'ueture of ordered fl-brass.

During the 1960s and 1970s our understanding of structural phase transitions progressed rapidly-- in a large part owing to the development of experimental techniques which enabled the microscopic properties to be probed more accurately. In particular, the measurements by neutron and Raman spectroscopy and by magnetic resonance techniques have proved to be especially powerful in exposing the l~chness and variety of phenomena displayed by systems undergoing structural phase transitions.

I t is our intention in this, the first of three articles, to describe these developments within the unified framework afforded by the phenomenological theo .ry of Landau (1937); we defer to parts I I and I I I a discussion of those areas where Landau theory has proved to be inadequate. In this review we shall use as examples the phase transitions in BaTiO3, SrTi03, NaNOa and/?-brass to illustrate the wide range of the phenomena occurring at structural phase transitions. We shall concentrate on providing a unified framework and approach rather than to provide an exhaustive account of the experimental data and theoretical treatments of any particular material. Excellent books using this lat ter approach have been writ ten by Jona and Shirane (1962) and more recently by Lines and Glass (1977).

Section [.2 of this article is a review of the notat ion and the basic lattice dynamical results tbr the susceptibility and scattering cross-sections of a crystal. In § 1.3 we describe and compare the different experimental techniques which can be used to measure the order parameter and susceptibilities close to a phase transition. We also discuss the advantages of different techniques and illustrate the results with particular regard to the phase transitions discussed above.

The simplest theory of structural phase transitions is Landau 's (1937) phenome- nologieal theory, which was developed independently in great detail for BaTiO3 by Devonshire (1949). We believe that the first stage in the understanding of a structural phase transition is the development of the appropriate Landau theory in terms of various phenomenologieal parameters. In § 1.4, we describe the application of Landau theory to a very wide range of different structural phase transitions. In view of the recent interest in incommensurate phases, we discuss these in considerable detail.

Landau theory involves a considerable number of parameters which, at the phenomenological level, must be obtained from experimental measurements. In

Page 6: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

(i g . A . Cowley

principle they can also be obtained fronl a microscopic theoi T which will then answer such questions as why the undistorted phase is unstable, and what is the transit ion temperature, To. In § 1.5 we review the different microscopic models which have been developed to describe structural phase transit ions and to obtain the parameters of a Landau expansion.

§ 1.2. PRELIMINARIES

1.2. I. The hamiltonian Below the critical t empera ture for a s tructural phase transit ion the mean

position of an a tom is displaced from its position in the h igh-symmetry phase. ( 'onsequently the theory of structural phase transitions makes use of the results and notation developed to describe displacements of the a toms from their equilibrium positions. We adopt the notat ion which is now commonly used in lattice dynamics (Cochran 1973, Cowley 1968). In the high-temperature, h igh-symmetry phase the mean equilibrium positions of the a toms are R(hc), where I denotes the unit cell and K the lcth type ofs different a toms within the unit cell, so tha t R(/~) = R(l) + R(~). The displacements of the a toms from these equi li br ium positions are wri t ten u (l~), so tha t the position of the a toms at any instant is r(/K) = R(l~) + u(/~).

The potential energy of the crystal, V, may then be expanded, tbrmally, as a power series in the displacements, u(hc). In the harmonic approximat ion all of the terms of higher order than second in this expansion are neglected. Within the harmonic approximat ion the equations of motion for an a tom can be diagonalized by introducing a plane-wave representation, when the equations dccouple for different wave-vectors, and the resultant dynamical matrices can be diagonalized, to give the frequencies ~o(qj) and normalized eigenvectors e(K, q j) for the normal modes. In terms of these normM modes the displacements are wri t ten

u(/~) = ~ (NM~) l/2e(K, q j ) exp (iq. R(hc))Q(qj), (i.2.1) qJ

where M~ is the mass of a toms of type K, N is the number of unit cells in the crystal and Q(qj) is the ampli tude of the normal mode, which, since u (lye) is real, mus t satisfy the constraint,

Q(qj) = Q*( - q j). (I.2.2)

In terms of these normal modes the harmonic par t of tile hamiltonian becomes

H2=1-2 2 qj [Q(qJ)Q( - q j) + c°(qJ)2O(qJ)O( - q J)]' (I.2.3)

where

Q(qj) _ dQ(qj) dt

the momen tum conjugate to Q(qj).

- - - p ( q j ) ,

In the usual development of lattice dynamics the normal mode coordinates are wri t ten in terms of phonon creation and destruction operators;

• / / h "~1/2 Q ( q 3 ) = t ~ ) (a(qj)+a*(-qj)) (I.2.4)

Page 7: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 7

from which we can obtain the thermodynamic average

(Q(qj)Q(-qj )}= ~ (2n(qj)+ 1), (I.2.5)

where n(qj) is the Bose occupation number, n(qj)=l/(exp(flhco(qj))-l), and = (kBT)- 1. In most applications classical statistics are adequate when eqn. (I.2.5)

becomes

( Q ( q j ) Q ( -- q j ) ) = (]~(A)(qj)2)- 1 (1 .2 .6)

Structural phase transit ions occur only in crystals tha t are significantly anharmonic and for which the strictly harmonic frequencies, co(q j), may be imaginary. The results of eqns. (I.2.4)-(I.2.6) can, however, be used provided tha t co(q j) is interpreted as a sui tably renormalized frequency as we shall discuss in § 1.5.1 (Cowley 1964). We shall need to consider the terms in the expansion of the potential energy beyond second order which m a y be writ ten as

v/'q 1q2q3"~ - . / q t\ -./'q2"~ ~/q3\ . . . q . ? . q . + HA -- q.xi~2.qs t., 132.) 3 ) t~ 1 ) t.)2 ) t.,3 ) .... (I.2.7)

7132J3

where the detailed tbrm of the coefficients is given in terms of the atomic potential functions and the normal mode eigenvectors by Born and Huang (1954). Since in a pm'e system wave-vector is conserved q3 = - q l - q2 +'c, where • is any reciprocal lattice vector of the h igh-symmetry phase.

At many structural phase transitions the size and shape of the crystal alters. These changes are described by introducing infinitesimal strain parameters , r/~. The displacement of an a tom due to an infinitesimal strain is

u g l l c ) = ~ ~=~R~( l l c ) . ( I . 2 . 8 )

A comparison of eqns. (I.2.8) and (I.2.1) then suggests tha t

q~n = lira - i% Q(q~, c~), (I.2.9)

q.+o 4(NEsvs )

where Q(q~, c~) is the acoustic mode ampli tude with eigenvector polarized in the direction and wave-vector in the fi direction. Alas, however, it is not possible to represent any -particular combinat ion of strain parameters in terms of the acoustic mode amplitudes: consequently, the terms involving the strains (or long wavelength acoustic modes) mus t be excluded from the s tmlmation over (q j) in eqns. (I.2.3) and (I.2.7) (Born and Huang 1954) and wri t ten as fro'thor contributions to the hamiltonian, with the form

V 0

+v (:Ls.10)

where we have writ ten only the three most impor tan t terms in the double power series expansion about the equilibrium positions of the h igh-symmetry phase. The

Page 8: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

8 R . A . Cowley

first term in eqn. (1.2.10) describes the coupling of a strain to a long wavelength optic mode of vibration, while in the second te rm use has been made of wave-vector conservation. The third te rm contains the usual elastic constant multiplied by the volume of a unit cell.

At orde~disorder phase transitions, as discussed in the introduction (figs. 1.3 and 1.4), the different possible positions of the a toms may be a considerable distance apart . Although it is always possible to express the hamil tonian in a power series in the dispiacements, the series will be, a t best, only very slowly convergent and consequently very difficult to employ in practice. An al ternat ive approach is to introduce a new Ising spin variable which describes the large-scale motions. In the ease of NaNO2 the spin variable locates the N a tom of each NO2 group in either the positive, S(l~c) =½, or negative, S(l~c) = -½, b direction. In the ease of NaNO2 there is only one NO2 group in each unit cell so tha t we may then drop the suffix, K, and have a single spin variable S(l) describing the orientation of each NO2 group.

A similar notat ion may be developed for a binary, AB, ordeI~disorder alloy. I f the crystal has only one type of site (i.e., one value of 1c) in the h igh-symmetry phase then we may introduce a spin variable S(l) =½ if the site l is occupied by an A atom and - ½ if it is occupied by a B atom. This model m a y also be extended to include more complex alloys and more complex high-temperature phases.

In the case of the A B alloy the potential energy between a toms at R(1) and R(l') may have different functional forms according to whether both a toms are of type A, both are of type B, or if there is one of each. Denoting these interaction energies b y 4AA, 4BB and 4AB = 4BA the energy of interaction between the two sites may then be wri t ten as

4(ll') = 4o(ll') + 24 ~ (ll') (S(l) + S(l') ) + 442(ll')S(1)S(l'), (1.2.11)

where

1 4o = ~ ( 4AA + 4B. + 24AB),

1 41 -~7(4AA-- 4BB)I

1

This is then a general representation of the potential between l and 1' irrespective of the occupation of the sites.

The third term of eqn. (I.2.11) gives rise to an interaction between the spin variables a t different sites. I t is frequently useful to diagonalize this te rm with the a toms at their mean equilibrium positions to obtain eigenvectors E(K, qj) and eigenvalues J (q j ) such tha t

S(l~) = ~ N ~ E(~, q j ) e x p (iq. R(1K) )S(qj),

where S(qj) is the ampli tude of the spin normal mode.

(1.2.12)

Page 9: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 9

1.2.2. Susceptibility The change in the proper ty of a system when an external field is applied is

described by the susceptibility: for example, the dielectric susceptibility gives the change in the dielectric polarization on application of an electric field. More generally, we consider a property of the system described by an operator, B(r), which is dependent upon position, and apply to the system some time-dependent 'field' E(r) whose interaction with the system we write as H ' = - A E with A(r) some other operator. We may then introduce a position-dependent susceptibility

XBA(r--r' = - - B r (I.2.13) ) dE(r ')( ())[E(,')=0.

Writing the response in the form

(B(r))=Tr{exp(-fl(H+H'))B}/Tr{exp(-fi(H+H'))}, (I.2.14)

where H is the hamiltonian of the unperturbed system, the susceptibility follows from eqn. (I.2.13) in the classical limit, as

ZBA(r-- r') =fl(B(r)A(r')). (I.2.15)

Since most of our interest is directed towards pure materials (where wave-vector conservation plays an impor tant role), it proves useful to introduce the Fourier transform of this susceptibility,

Z,A(K) =fl(B(K)A(- K)), (I.2.16)

where the operators B(K) are the Fourier transforms of B(r)

B(K) = S B(r) exp ( - i K . r) d3r. (I.2.17)

The operators B(K) may be expanded formally in powers of normal mode coordinates Q(~) (or, in the appropriate cases, the 'spin' coordinates S(qj)): thus we write

( l ; 2 B(K) =B°(K)A(K) + ~ y'B(K,qj)Q(Rj)A(K+R), (I.2.18)

qJ

where B°(K) is the operator for the undistorted crystal. Wave-vector conservation is ensured by the delta function which is defined by

(2g) 3 A(K) = ~ exp (iK. R(I)) = ~ 5(3)(K-v), (I.2.19)

l V

where the summation is over the reciprocal lattice vectors, ~, of the undistorted lattice, while 5(3)(K) is the normal three-dimensionM delta function.

The expressions for the coefficients in the expansion, eqn. (I.2.18), may be found once an expression is known for the operator B(r). For example, if B(r) is the dipole moment operator

B(r) = M (r) = ~ Z~u (IK)5(r -- R(IK)),

where Z~ is the apparent charge, then

while B°(K)=0 .

M (K, q j) = ~ Z~M[ 1/Ze(~, q j) exp (i(K + q). R(~)), (i.2.20) K

Page 10: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

10 R . A . Cowley

Now, returning to the susceptibility function, eqn. (I.2.16), we make use of the expansion (I.2.18) for B(K) and A(K), picking out, in each case, the term in the expansion tha t is linear in the normal coordinate. The resulting contribution to the susceptibility is

Z,A(K) = ~ A(K + q)B(K, qj)A(- K, q j ' ) z ( q j j ' ) . (I.2.21) qjj"

The importance of this result is tha t it shows that any one-phonon susceptibility can be obtained in terms of' the susceptibility for the normal mode operators,

z(qjj ') = f i<Q(qj)Q( - qj')>. (I.2.22)

Using the results quoted above, eqn. (I.2.6), for the harmonic crystal with classical statistics, this becomes,

l .u(qj j ' ) = co(q j ) - 25jj,. (I.2.23)

A similar analysis may be carried through for an order-disorder system. The operator, B(r), is expanded in a power series in the spin operators (or, if necessa W a double power series in the spin and displacement operators) with

/ l \1/2

\ / qJ

where the coefficients B , ( K , q j) are given in terms of the spin eigenvectors. If" the dipole moment associated with the lcth type of' molecular group is p~, then the coefficients in the expansion of the dipole moment operator are given by

M,(K, q j) = 2 ~ p,~ E(,< 'tJ) exp (i(K + q). R0,:)). (I.2.25) / (

The susceptibility is then given by eqn. (1.2.21) but with the normal mode susceptibility replaced by

z,(q.~i') = f i < a ( q j ) S ( - qj')>. (I.2.26)

The susceptibility can be generalized to describe the response of a system to a t ime-dependent field, E = ½ E o (exp (icoz) + e x p (-ic~v)). The susceptibility is then dependent upon the frequency (o as well as the wave-vector K and written ZBa(K, (,J). The calculation of this fl 'equency-dependent susceptibility is more difficult than that of the static susceptibility described above but is detailed in many texts such as Landau and Lifshitz (1959). Since a t ime-dependent field may lose energy to the system, the frequency-dependent susceptibility has both real and imagina W parts which are connected by a Kramers-Kronig relation

xRA(K,(o) = 1 r/~A, (K'c°') dec', (1.2.27) J (o~ -~)~

and which satisfy the symmetry properties

zRA(K,o))=Z~B(--K--CO) and Z~A(K,(o)=--Z~B(-K.--co ).

Since energy must be lost to the system, the imaginary part of the diagonal part of the susceptibility, ;g~a(K co), is positive for co>0.

Page 11: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Struc'tural phase transitions l 1

Finally the dynamic susceptibility can be obtained for one-phonon processes as described above for the static susceptibility. The dynamic one-phonon susceptibility x(qJJ', 0)) can then be calculated explicitly for harmonic normM modes as

xR(qj~,, 0 ) ) = (0)(q j )2 __0)2) - 15jj, ' (I.2.28 a)

and

X ~ (q2.)'", o) ) - 20)(qj) [(5(co-0)(qj))-5(0)+0)(qj))]Sjj,. (I.2.28 b)

1.2.3. Scattering cross-sections The calculation of the cross-section for the scattering of an incident plane-wave

beam of particles with wave-vector k 1 and energy E o into a plane-wave beam with wave-vector k2 and energy E is described by Van Hove (1954). Within the first Born approximation the differentiM cross section can be written as

d2a k 2 , - - /~S(K, 0)), (I.2.29)

dF~d0)

where K is the wave-vector transfer N = k ~ - k 2 , and h0)=Eo--E is the energy transfer. The Van Hove scattering function, S(K, 0)), is given by

S(K, 0)) - - 2 piF(K)ifF ( - K)fib(~ - 0)fi), (I.2.30) if

where F(K)i f is the expectat ion value of an operator describing the interaction between the initial and final states of the system. I t is sometimes more convenient to rewrite eqn. (I.2.30), making use of the Heisenberg representation of the operators:

S ( K , 0 ) ) = , l ~ I ~ {F(K, OF(-K,O)}exp( io)~)d~. (I.2.31) ZT~ j - c o

In the case of X-ray or neutron scattering, the interaction between the particles and the specimen can be expressed in terms of a scattering length bt~ as

F ( K ~) = ~ bl~ exp (iN. r(b(, r)), (I.2.32) IK

where r(hc, r) is the instantaneous position of the atoms (he). For the scattering of X- rays by rigid atoms of a particular type Ic, the scattering length bz~ is the same for each site l, being equal to the scattering power of each single electron multiplied by the atomic form factor, which is dependent upon the wave-vector, N. For neutrons the scattering length may vary from site to site, owing to the randomness in the isotopic distribution and the spin alignment of the nuclei. This randomness results in incoherent scattering which usually gives rise to a troublesome background: we shall largely neglect it. The coherent scattering reflects the average scattering length b~ = 1/N E b~ and hence includes the interference effects between the different ions. The

t

X-ray or coherent neutron scattering is then given by

F(K, ~) = ~ b~ ~ exp (iK. r(/K~ z)). (I.2.33) l

Page 12: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

12 R . A . Cowley

The operator F(K, r) can then be obtained by expanding the position coordinate r(bc, ~)in powers of the displacements. When these displacements are transformed into the normal mode coordinates one finds

/ 1 \1/2 F ( K , r ) = F ° ( K ) A ( K ) + I | ~F(K,qj)A(K+q)Q(qj.T)+ .... (I.2.34)

\ N / .~

where Q(q)', ~) is the instantaneous amplitude of the normM mode coordinate, and the coefficients are given by

while

F°(K) = ~ b~ exp (iK. R(lc)) exp ( - W~), K

(1.2.35)

F(K,qj)=~b~M;~/2(K.e(~c, qj))exp(i(K+q).R(~c))exp(-W~), (I.2.36)

where W~ is the Debye-Waller factor. Using these results S(K, co) can be divided into a Bragg part, S°(K, co), and a one-phonon part, SI(K, co), where

S°(K, co)= N]F°(K)]ia(CO)A(K) (1.2.37)

and

Sl(K,co)= ~ A(K+q)F(K,qj)F(-K, -qj')&~(qjj',co), (I.2.38) qjj '

where ~90(qjj ', co) is the displacement-displacement correlation function

5P(qjj', co)---- 2~ (Q(qj, r)Q(-qf,O)}exp(icor)dr. (I.2.39) --cG

This displacement-displacement correlation function can be related to the imagina W part of the dynamic susceptibility introduced in the previous section as shown for example by Cowley (1963). The result is tha t

(,~. (CO) _~_ I . . . . . 1)Z (q33, co) = ~ ( q . 7 3 , co), (I.2.40)

where n(co) is the Bose-Einstein factor. Fm'thermore the static displacement- displacement correlation function

f ~o /f (qjj', ~(qJi ' ) = co) dco (1.2.41)

is within the classical approximation, ha)<< ]CBT, related to the static susceptibility, by use of eqn. (1.2.27), to give

fl~(qj?") = )/(qjj'). (I.2.42)

Since all X-ray and many neutron scattering experiments measure the integrated scattering, this relationship is of considerable importance.

These results for the susceptibility and scattering show the one-phonon parts are determined by the normM mode susceptibility multiplied by the appropriate coefficients. Provided, then, tha t these coefficients can be treated as wave-vector-, frequency- and temperature-independent constants, the behaviour of the suscept-

Page 13: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Str'a, ctura/ phase transitions 13

ibility or scattering cross-section close to a phase transition will be very similar, irrespective of the probe used to perform the measurements.

A similar analysis can be carried through for an order disorder system. For example in an A B alloy the coherent scattering lengths may be bA and bB, which are not usually the same. The structure factor, eqn. (I.2.33), then becomes

F(K, z) = ~ ~ (b~ + S(l~c, r)b'~) exp (iK. r(b~, ~)), (I.2.43) l

where b~ = ½(bA + bB) and b'~ = b A - b,. The expansion of F(K, z)is now given by eqn. (I.2.34) with b-~ replacing b~. The leading term in the spin variables is given by

(1)U22Fs(K qj)A(K+q)S(qj , z), (I.2.44)

where

F~(K,q j )=~b '~exp( -W~)E(~c ,q j )exp( i (K+q) .R(~) ) . (I.2.45)

Substitution of eqn. (I.2.44) into eqn. (I.2.31) and repeating the development given above then shows that the cross-section mirrors the behaviour of the spin correlation flmction given by eqn. (I.2.26).

Finally to determine the Raman scattering of laser light requires an expansion of the polarizability of the materiM in powers of the displacements. Formally the expansion can be written

/ 1 \1/2 P~p(K) =P°~(K)A(K)+ ( ~ ) ~, [P~e (qj)Q(qj)A(K + q)

\ / qJ

+ P(~(qj)S(qj)A(K + q)] + . . . . (I.2.46)

Substituting this expansion into eqn. (1.2.31) yields an expression for the Raman scattering tensor I ~ , which determines the scattering into a polarization (fiS) fl'om an incident beam polarized in the (~) direction. The cross-section is thus dependent upon the displacement displacement or spin-spin correlation functions, as in the case of X-ray and neutron scattering. Unlike the neutron or X-ray scattering case, however, in g a m a n scattering the wave-vector transfer K, is restricted to be much smMler than a reciprocM lattice vector.

§ 1.3. O R D E R P A g A M E T E R S A N D F L U C T U A T I O N S

1.3.1. Primary order parameters At a structural phase transition the symmetry of the crystal changes: some of the

symmetry operations of the high-symmetry phase, which is usually but not always the high temperature phase (see §I.4.5), are not symmetry operations of the low symmetry phase. Above the ferroelectric phase transition ofBaTi03, fig. I. 1, each Ba and Ti ion is at a site having the full cubic symmetry but in the distorted phase the symmetry at each site is only tetragonal. This breaking, in the low-symmetry phase, of some of the symmetry operations of' the high-symmetry phase is the essential characteristic of most phase transitions. (However, cf. §I.4.3.2.)

The displacements of the atoms which are associated with the breaking of the symmetry are characteristic of the distorted phase, and, furthermore, at a

Page 14: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

14 R. A~ Cowley

continuous phase transit ion decrease continuously to zero as the h igh-symmetry phase is approached. In a single domain of the distorted phase, the displacements in any one unit cell have a definite relationship to those occurring in any other unit cell. In BaTiOs the displacements are equal in every unit cell, while in SrTiO3, fig. 1.2, they are of equal magni tude but are systematical ly in opposite directions in neighbouring unit cells. This long-range order is characteristic of distorted phases, and suggests tha t the displacements can be described in terms of the normal modes, or spin variables, introduced in §I.2. This is indeed the case a t the major i ty of structural phase transitions one exception being the lock-in phase transitions discussed in § 1.4.5.

The simplest example of a structural phase transit ion occurs when the distortions may be described by a single normal mode; the ampli tude of the distortion is then given by the order paramete r

Qo = N - 1/2(Q(qsj)) , ([.3.1)

where (q.~j) is the part icular normal mode which describes the atomic displacements in the distorted phase, and the factor of N U2 is introduced so tha t Q0 is independent of the crystal size.

I f the phase transit ion is continuous the order pa ramete r will approach zero as T-~To and is assumed to be given by a power law;

Q0 ~ (T¢ - T) ~, (I.3.2)

where fl is known as the critical exponent for the order parameter . In classical theories of phase transitions, such as those discussed in §§ 1.4, and 1.5 the exponent fl =0-5, while in practice fi varies between 0"25 and 0-5 as described in par t II .

Frequent ly the possible atomic displacements of the distorted phase cannot be described solely in terms of a single normal mode, but only in terms of linear combinations of n normal mode coordinates; n is then known as the number of components of the order pa ramete r (or the order paramete r dimensionality) attd in the h igh-symmetry phase all these n normal modes have the same frequency. We shall write the ampli tudes of these normal modes

Qi = N - 1/2 ( Q(qsiji) )

and the order parameter in the distorted phase as

Qi = q0~i, (I.3.3)

where

F, ;~ = ~, (~.3.4) i

and the parameters ~ determine the structure of the distorted phase, while Qo varies with tempera ture as described by eqt~. (I.3.2).

Both BaTi03 and SrTi03 are examples of systems where n = 3. [n the ease of BaTi03 the ordered phase may be characterized by displacements along the x, y or z axes, each of which corresponds to a different normal mode, while in SrTi03 the oxygen oetahedra can rota te about any of three cube axes. In both of these cases the wave-vectors qsi are the same for all three values of i; q~=0 for BaTiO3.

1 1 1 q.~ = (3, 3, ~)2~c/a for SrTiO 3. For this type of phase transition n cannot be larger than 3. Recently, however, there has been considerable interest in incommensurate phase

Page 15: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 15

transitions, where (in cubic materials) the value o f n may, in principle, be as large as 48, as described in § 1.4.4.

Although we have introduced the concept of an order parameter through the displacive phase transit ions shown in figs. 1.1 and 1.2, the same approach m a y be used to specify an order pa ramete r in the ordet~disorder ease when

Qo = N - 1/2(S(q~)}. (I.3.5 a)

Thus, for example in the case of NaNO2, the order paramete r gives a measure of the difference between the numbers of NO: groups displaced along the positive and negative b axis.

A more format description of the order pa ramete r at a phase transit ion may be given in the language of group theory (Landau and Lifshitz 1959). The density distribution of a crystal p(r) can be expressed as a linear combination of normalized functions, ~b, which t ransform into one another under the symmet ry operations of the group, G, of the h igh-symmetry phase. I t is convenient to choose the functions ~b so tha t they divide into a sum of sets of functions which t ransform only within the set under the symmet ry operations. These functions, ~bf, then form the bases of the irreducible representations of G, where p labels the irreducible representat ion and i the number of the functions in tha t representation. Then amongst the functions ~b~ there is one which is invariant under G, and this describes the s t ructure p0(r) of the h igh-symmetry phase. Any distortion 5p(r) can then be wri t ten as

~p = Z' Z clef, (I.3.5 b) p i

where Z' omits the term describing the structure of the h igh-symmetry phase. Now at a continuous phase transit ion the distortion in s tructure can be described in terms of only one irreducible representation, p = s, say; the coefficients C~ then describe the structure of the distorted phase.

The displacements of the a toms about their equilibrium positions arc a set of flmctions which t ransform under the s ym m et ry operations of the group, G. In setting up the normal modes in § 1.2 the symmet ry of the crystal ensures tha t each normal mode transforms as a part icular irreducible representat ion of the space group. The functions Q(qj) are then proport ional to the functions ~b~ ~ introduced above and the coefficients C~ are proportional to the order parameters Qi. The number of components of the order parameter , n, is then the dimensionMity of the sth irreducible representation of the full space group.

1.3.2. Secondary order parameters The te rm secondary order parameters is used, alas, in two different ways. In the

first instance, it is used to describe distortions (in the low-symmetry phase), which arise at the phase transition, but which do not break as many of the symmet ry elements of the h igh-symmetry phase as are broken by the pr imary ordering. This concept may , perhaps, be clarified by considering BaTi03 and SrTiO3, tbr both of which there is a macroscopic strain in the distorted phase in addition to the pr imary distortion. In the case of BaTiO3 the point group of the high- temperature phase has the full cubic symmetry , m3m, and the distorted phase has the tetragonal point group 4 mm. If, however, the distortion to a tctragonal s tructure arose purely from the tetragonal strain, ~-(~7~+~y/2), then the distorted phase would have the symmet ry 4 /mmm. The point group 4 / m m m can be obtained from 4 m m by adding a mirror plane perpendicular to the tetragonal axis. The symmet ry of this mirror plane

Page 16: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

16 R . A . Cowley

is broken by the distortions described by the pr imary order parameter but is not broken by the secondary order parameter , the tetragonal strain.

In the ease of SrTi03 the symmet ry of the low-temperature phase is 4 /mmm but contains two formula units in each unit cell, whereas if the only deformation was the strain the point group symmetD, would be 4 /mmm but there would be only one formula unit in each unit cell. Consequently the prima~2v order paramete r breaks the translational symmet ry between the two formula units in each unit cell of the distorted phase whereas the secondary order paramete r (the tetragonal strain) does not.

Since the distortions produced by a secondary order paramete r break the symmet ry of the h igh-symmetry phase they must decrease to zero as T-+T¢. We therefore write the tempera ture dependence of a secondary order parameter , (~, whether it be the strain or another normal mode, as

~)~ (T¢ - T) ~. (I.3.6)

Since the seeondaLT order pa ramete r does not break all the symmetI:y elements tha t are broken a t T¢, it must, when close to T¢ have a vanishingly small ampli tude compared with the pr imary order parameter ; thus one expects t h a t ~ > ft.

The other use of the te rm secondary order paramete r arises when there is more than one normal mode (q j) which has the symmet ry of the pr imary order parameter . For example, in BaTiO3 there are three different frequencies of the transverse optic modes of long wavelength. Although we can choose the mode eigenveetors so tha t the displacements close to T~ are given by only one of these modes, as the distortions become larger the pa t te rn of the atomic displacements may alter and this can be represented as a coupling to the other normal modes.

1.3.3. Experimental determination of the order parameter

1.3.3.1. Macroscopic measurements I f the order pa ramete r has wavevector qs = 0, a ferrodistortive transition, then its

tempera ture dependence may be determined by macroscopic measurements. For example, in ferroelectric materials }he spontaneous polarization, Ps, is proportional to the ampli tude of the distortion of a long wavelength infra-red active mode. In the notat ion developed above

1 M (Q?')(Q(0j)}, (I.3.7) PS--vN1/2 j

where v is the volume of the unit cell-and the dipole moment operator of the mode is given by eqn. (I.2.20).

I f there is only one mode j, eqn. (I.3.1) shows tha t

1 Ps = - M(0j)Qo, (I.3.8)

V

showing tha t a measurement of Ps determines the order pa ramete r Q0. In fig. 1.5 we show the measured spontaneous polarization for BaTiO3 (Merz 1953); note tha t the phase transit ion in BaTi03 is of first order.

The spontaneous polarization gives a measure of the pr imary order paramete r only if the phase transit ion is to a ferroeleetric phase. I f the order paramete r

Page 17: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 17

28

24

E o 16

o u ~ 12-- o

EL

4

0 20

Fig. 1.5

1 I t I I I I

_ - - ~

I I I 1 I I I l 50 40 50 60 70 80 90 I00

Tempero ure, °C HO 120

The spontaneous polarization of BaTi03 as a function of temperature (Merz (1953) aider Jona and Shirane (1962)).

corresponds to a q~ =0 mode of even symmetry, then there will be a change in the refractive index or optical polarizability of the crystal, proportional to the order parameter Q0. For example, the optical polarizability is given by eqn. (1.2.46) as an expansion in powers of the normal mode coordinates. If the leading term is non-zero the change in polarizability is given by

& P=f = P~f (Oj)Qo, (I .3.9)

if there is only one mode j. Har ley and MacFarlane (1975) have measured the order parameter in DyVO¢ and TbV04 using the birefringence of the crystal as illustrated in fig. II.15 of par t II.

Finally if the primary order parameter is the macroscopic strain, it can be measured macroscopically, or possibly more reliably by using neutron or X-ray diffraction.

Macroscopic techniques have the disadvantage that they can only measure the primary order parameter at ferrodistortive, qs = 0, phase transitions, and further- more require single domain samples to be available. Unfbrtunately single domain samples can usually be produced only by applying a field to the specimen which may then influence the results. Also close to To, the susceptibility becomes large, see § 1.3.4, and the measurement of a small spontaneous polarization, in the presence of a large dielectric constant is far from being a technically easy experiment.

Secondary order parameters can be measured using macroscopic techniques, but the results for the spontaneous polarizatioa or refractive index may be difficult to interpret. This is because the two-phonon term in the expansion of the dipole moment or polarizability, eqn. (I.2.46), is of the form,

p(2)(K 1 /o / q i q2~Q(ql~Q(qE~A(K+ql+q2). aft, ) : N qlq2jJj2 ~ afl~kjl j2 , ] \ ~ 2 / I \ j l , ]

I f we approximate (Q(qlJ~)Q(q2J2)) by (Q(qsj))(Q(-q,j)), then this term will

Page 18: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

18 g . A . Cowley

contribute a term to the change in polarizability

a P~ ~ l°~a(qj - q~ q2 j jo .

This term may well be of the same order of magnitude as the term arising from the one-phonon contribution of the secondary order parameter ~), which is given, eqn. (I.2.46), by

Without a detailed knowledge of the polarizability parameters it is impossible to distinguish between these two processes experimentally.

1.3.3.2. Scat ter ing m e a s u r e m e n t s The theory of the scattering of X-rays and neutrons from crystals was briefly

reviewed in §I.2.3. The structure of the crystal is determined by the Bragg reflections, which are dependent upon the infinitely long time part of the correlation function, eqn. (I.2.31). In the long time limit

(F(K, ~)F(K, 0)} = KF(K, 0)512, (I.3. ]())

so that the Van Hove scattering function has the form

S°(K, co)= KF(K, 0)}12c5(co). (I.3.11 )

In the high-symmetry phase these results have already been given by eqns. (I.2.35) and (I.2.37) and enable the structure of the phase to be determined experimentally.

In the distorted phase the scattering can be obtained if the positions R(~') are replaced by the positions in the distorted structure. I t is, however, mm'e instructive to expand the structure factor (F(K,0)} in powers of the displacement as in eqn. (1.2.34). When the expectation value of eqn. (2.34) is taken one obtains

( F ( K , 0 ) } = F ° ( K ) A ( K ) + ~ F(K,qsij~)(Q~)A(K+q~) (I.3.12) i = 1

where we have neglected, for the present, higher powers in the primary order parameter and all secondary order parameters.

Now if q~ is non-zero, eqn. (I.3.12) shows that there will be new Bragg reflections present in the distorted phase with an intensity given by NIF(K,q , i j~ )12(Qi} 2 A(K + q~i). The intensity of these Bragg reflections enables the relative sizes of the different components of the order parameters, the (Q~}, to be found while their temperature dependence gives a measure of the exponent fl describing the temperature dependence of the primary order parameter. The temperature depen- dence of the new Bragg reflection in the antiferrodistortive phase of SrTi03 is shown in fig. 1.6 as determined with neutron scattering techniques by Riste et al. (1971).

The behaviour is less straightforward for ferrodistortive phase transitions, qs = 0. The intensity is given by

S°(K, co) = Nh(co)IF°(K) + ~ F(K, Oj,) <Q,}I2A(K). (I.3.13)

The change in the intensity oI" these reflections close to T¢ is proportional t o (Qi) 2 because the cross terms cancel, provided that F°(K) is not zero or strongly

Page 19: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 19

500

c c ° 40(3

30(3

c 20C

E o:mC

0 8'0"

Fig. 1.6 i i

IXQ

X Q

90

S~T, %(½ ~ ~> EO:13 5meV • • thick crystal . . t h i n crystal

.

x .

%

I n

100" 110" ( 'K)

27~ The temperature dependence of the superlattice reflection K = - - (g,1 g,t g)3 1 n ~r~l" "]O 3 after glste'

a, et al. (1971). The dotted line indicates that To tbr the specimen was 106"1 +0 '3K.

tempera ture dependent. (We discuss the tempera ture dependence of the Debye Waller factors in §§ 1.3.5 and 1.4.)

Tile use of scattering techniques to determine secondary order parameters is subject to the same difficulties as those discussed for macroscopic properties. In the expansion of the s tructure f~ctor ( F ( K , 0)>, eqn. (I.2.38), there are not only linear terms in Q(qj) but second-, third-, and higher-order terms. The second-order terms will give rise to additional contributions to the scattering with the wave-vector conservation conditions given by A(K) and A(K + 2q~). The intensity arising from these diffraction effects caused by the p r imary order parameters , must be included when the ampli tude of the secondary order parameters are being deduced from the intensities measured at these wave-vector transfers. Note that , whereas the relative sizes of the one- and two-phonon terms in the expansion of the polarizability are largely unknown, in the ease ofdisplaeive phase transitions these relative amplitudes tbr scattering experiments are given exact ly in terms of the supposedly known eigenveetors and scattering lengths. I t is thereibre possible to separate these two effects at least in principle. This advantage of diffraction is frequently rather illusory in practice, because the effects of the two-phonon terms in a diffraction experiment are usually relatively larger than in the polarizability experiment.

At strictly orde>disorder phase transitions, such as the AB alloy, there are only linear terms in S(hc) in the expansion of F(K, ~), eqn. (I.2.43). Consequently the interpretat ion of the scattering is simpler in these systems than in the displacive materials. In the molecular systems, such as NAN02, however, the si tuation is more complex because of a coupling between the spin and the displacement variables, so tha t it again becomes difficult to ext rac t secondary order parameters reliably.

There are two experimental problems in the application of diffraction techniques to the determinat ion of order parameters . First ly Bragg reflections from real crystals frequently suffer from extinction effects. At antiferrodistort ive phase transitions close to T~, the new Bragg reflections of the distorted phase are weak and probably not greatly influenced by extinction, as shown (fig. 1.6) by the agreement between results of experiments on different sizes of crystal; however, as the system is cooled extinction will become more and more important . Ext inct ion provides a particularly difficult problem at ferrodistortive transit ions because the crystal will usually break

Page 20: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

20 R . A . Cowley

up into domains changing the extinction effects on the normal Bragg reflections. Changes in these Bragg intensities will not then reflect the order pa ramete r in the way suggested af ter eqn. (I.3.13).

A second problem (to be discussed in detail in the next section) is that , close to T¢, the diffuse scattering is large for small frequency transfers and at those wave-vector transfers at which the Bragg reflections of the distorted phase appear. This is the cause (at least in part) of the high tempera ture ' tail ' to the Bragg reflection intensity shown in fig. 1.6 (to which we shall devote much at tent ion in I I I ) .

The exponent fi, can be accurately obtained only to the extent tha t the diffuse scattering can be accurately subtracted out. Since the Bragg scattering is a delta function in both energy and wave-vector transfer whereas the diffuse scattering is broad the subtract ion is made easier if the experiment is performed with very good resolution in either K or co. In the case of conventional X- ray scattering it is impossible to energy discriminate against the diffuse scattering, but in principle it is possible to use very good K space resolution. I t is possible to energy discriminate using MSssbauer 7-rays as a source for the scattering and then to perform energy analysis (Darlington and O'Connor 1976) using a similar MSssbauer absorber. I t is surprising tha t this technique has not been applied to measure ft.

By using neutron scattering techniques, the energy analysis is more easily performed. Conventional inelastic scattering techniques give an energy resolution of about 0,-1 meV which is not normally good enough to permit the neglect of the diffuse scattering. The development of new techniques with a resolution of 1 #eV, such as the backscattering spectrometer (Alefeld et al. 1969) and the spin-echo spectrometer (Mezei 1972)holds out the possibility of eliminating the diffuse scattering more completely. The wave-vector resolution is, however, inferior to t ha t which can be obtained with the best X ray spectrometers.

1.3.3.3. Resonance methods The order pa ramete r a t a structural phase transit ion may be determined by using

electron paramagnet ic resonance, E.P.R. , or nuclear quadrupole resonance, N.Q.R. (Miiller 1971, Blinc 1971). The E.P. IL method depends upon the presence in the crystal of a magnetic ion. The most suitable ions are those with a half-filled closed shell, Mn 2 +, Fe 3 +, Gd 3 + or Tb 4 + because; (i) they are spherically symmetr ic and so do not introduce local distortions in symmet ry and (ii) they have very long E.P.R. relaxation times even a t room temperature . One of these ions must then be present in the crystal a t a low concentration, say, 101 s cm-3; Fe3 + is usually present in 'pure ' SrTiO3 at about this concentration.

The crystal field hamil tonian for a 3d transition metal ion can be expressed as a spin hamil tonian in terms of the normalized magnetic spin operators, 0N (Hutchings 1964), which t ransform in the same way as the corresponding cartesian polynomials;

2 4

m = - 2 m - - --4-

The coefficients in this expansion B~, B~ may then be obtained a t least ap- proximate ly in terms of the ionic charges, bonding and nearest neighbour distances (Hutchings 1964). Typical ly the B-coefficients are about 10-3 meV.

In the h igh-symmetry phase the site symmet ry is such tha t some of the B7 coefficients are zero; for example in SrTiOa with an Fe 3 + ion replacing the Ti 4+ ion

0 + 4 only the coefficients B 4 and Bg are non-zero. In the distorted phase, one (or more) of

Page 21: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Str,t~ctural phase transitions 21

the B-coefficients becomes non-zero and increases in a manner dependent upon the primary order parameter, Q0, and the secondary order parameter

B = A ,Qo + A2Q 2 + . . . + A 1(~ + A 2 ~ )2 + . . . . (I.3.14)

In the case of Fe 3 + in SrTi03, B ° is allowed in the distorted phase. However, since the F e 3 + ion is situated at a centre of symmetry even in the distorted phase, the coefficient A 1 in eqn. (l.3.14) is zero. The secondary order parameter for SrTiO3 is the strain and so 41 is non-zero. Consequently B ° for SrTi03 has two contributions proportional to Q2 and to Q, and as with the macroscopic polarizability it is difficult to distinguish the effects of these two different processes.

In contrast Gd 3 + ions in SrTi03 are situated at the Sr 2 + site, which is not a centre of symmetry in the distorted phase, and so A t of eqn. (I.3.14) is non-zero (Unoki and Sakudo 1967).

Miiller and Berlinger (1971) were, however, able to directly measure the primary order parameter in SrTi03 using the Fe 3 + centre. In a single domain specimen of distorted 8rTi03, fig. 1.2, the two Ti 4+ sites have their environments twisted in opposite senses. Hence, by determining the orientation of the principal axes of the spin hamiltonian for these two types of sites, the rotation angle of the oxygen octahedra could be determined. Using this approach with the Fe 3 +-vacancy centre in SrTi03 Miiller and Berlinger were able to determine the rotation angle in SrTi03 to _+0"01 °, fig. II.17, but more typically the accuracy of such experiments is 0"1 °

In an E.P.R. experiment the line width of the resonance is typieMly abou~ 10 .4 T, and so there will be a splitting once the local distortion of the structure persists for times longer than 10 6 s. Untbrtunately, the E.P.R. method cannot be used when the material is magnetic and it depends on being able to produce material with the necessary impurities situated at known sites in the host material. I t Mso suffers "from the disadvantage that there is always a niggling uncertainty as to whether the behaviour near the E.P.R. impurity site is the same as that of the bulk material.

Nuclear quadrupole resonance arises from the transitions between the different nuclear energy levels eaused by the interaetion between the nuclear quadrupole moment and the electric field gradient. The eleetrie field gradient may be expanded in a similar manner to the eleetrostatie potential which determines the E.P.R. spectrum, only the selection rules for the coefficients are different. In particular, if the ion is situated at a centre of symmeti T in the high-symmetry phase, the linear term, A 1, in eqn. ([.3.14) is zero even in the distorted phase and the B coefficient is proportional to Qo 2. Examples of this behaviour are provided by the work of Borsa (1971), using 139La and 27A1 resonances in LaA103, which undergoes a similar phase transition to SrTi03.

In materials of lower symmetry the linear term in eqn. (I.3.14) is non-zero and the B coefficients are proportional to the order parameter. This is illustrated by the 13 3Cs resonance in CsH2 As04 as shown in fig. 1.7 (Blinc and Mali 1969).

N.M.R. has the advantage over E.S.R. that it does not necessitate the introduction of an impurity, but the selection rules are such that the B parameters are more commonly proportional to Qg than to Q0 and so cannot be used to determine the exponent ft. The frequency resolution of the spectrometers is such that the local distortion must persist for 10 - 38 to give a sharp line characteristic of the distorted phase.

Finally the MSssbauer effect may be used to determine the order parameter at a

Page 22: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

22 R . A . Cowley

structural phase transit ion (Shenoy 1973). At a phase transit ion the MSssbauer line may split because of a change in the nuclear eneigy levels caused by the interaction of the nuclear quadrupole momen t with the electric field gradient. The selection rules and dependence of the splitting on the order pa ramete r are then the same as those described for N.M.R. experiments. The accuracy of M~issbauer experiments is usually less than tha t of N.M.R. experiments and furthermore the experiments frequently require tha t the MSssbauer nucleus be introduced into the materiM as an impuri ty.

Fig. [.7

2 4 -

22 0,= 4

20

,6 ',

I

I 8

4

2

o ]1 1 2 0 150

Tc

Temperature dependence

(0, H o) = 20" VL = 6 0 0 0 Mc/s

£ J._HO

O ~) "O

i i [ 200 250 5 0 0

r ['K]

of the nuclear quadrupole 133Cs magnetic resonance in CsH2As04 after Blinc and MMi (1969).

1.3.4. Static susceptibility In § 1.2, the susceptibility of a one phonon process was shown to be dependent

upon the coupling between the probe and the phonon (described by the operators such as B(K, q j) or F(K, q j)) and the one-phonon response fmmtions, z(qjj'), eqn. (I.2.22). h the distorted phase (Q(qsJ)} is non-zero, so that , at a continuous phase transition, it is reasonable to expect tha t (Q(qsJ)Q(-q~j)}, a measure of the fluctuations of the normM mode coordinate Q(q~j), will diverge at To giving rise, eqn. (I.2.22), to a divergence of tile susceptibility x(qsjj), as il lustrated for BaTi03 in fig. 1.8. We shall therefore s tudy the behaviour of the susceptibility x(qjj) as the tempera ture T ~ T o and wave-vector q-~q~, tbr the part icular modes whose expectation values are non-zero in the distorted phase. I t is convenient to introduce a notat ion to describe the fluctuations for the normal modes with wave-vectors close to q~. We define

(~/(k) =(2(k+qsi j i)-((~(k+qsi ii)} }

and 1 (1.3.15a)

Qi(k ) = Q(k + q~iJi).

Note tha t in the fluctuating par ts of the order pa ramete r we have not introduced the factor of N 1/2 included in eqn. (I.3.1); the Q~(k) are proportional to the phonon amplitude.

We introduce a corresponding notat ion for the susceptibility

z~(k) = )/(k + q,i3"~J~), (I.3.15 b)

= f i ( 0 ~ ( k ) ( ~ i ( - k ) } .

Page 23: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 23

90oo

8090

7000 ¢:

6000

5000

:~ 4000

3O0O x x

2000

1000 x

o 20

Fig. 1.8

I I I I I I

x T rising • T falling

x x • x

. 2

x

x •

x x I:a o ~ ] x x

x x x x x x X e 7 x x~Cx x x x e x e ~

1 I I I 1

4 0 6 0 8 0 100 120 140

Temperature T (°C)

.x ¢

e x x x

I

1/~ 4 x 10 4

] 1

=.

2x 10 -4

3

l x l 0 4

.G-

I 0 160 180

Static dielectric constant in BaTiO3, both above and below To. Below To there are two dielectric constants, parallel to (q) and perpendicular to (G) the ferroetectric c axis. (After Johnson 1965.)

In the h igh - symmet ry phase Xi(k) is independent of i and its divergence is given by sealing theory (of. S tanley 1971, §II .4.3.5) for T--*Tc, k--*0, as

- y 2 2 z i ( k ) ~ C + t D(k~{=), (1.3.16)

where

t = ( T - T o ) / T ¢ , ( I . 3 . 1 7 )

and 7 is the susceptibi l i ty exponent which is un i ty for classical theories, hut can be much larger in real systems. T h e funct ion D(x 2) is a scaling funct ion, (normalized such t h a t D (0) = 1 ) describing the wave-vec to r dependence of the susceptibil i ty. ~ is the correlat ion length in the ~ directibn. The correlat ion length diverges at a cont inuous phase transi t ion, and the divergence is described by the exponen t v

~ = f + ( a ) t -v, (I.3.18)

where classical theories yield v = 0"5, bu t more exact theories again give ra ther larger values.

There are two special cases of eqn. (1.3.16) of par t icular impor tance

Zd0) = C + t - ' , Ik ]=0 (I.3.19)

and

Zi(k~)=Do/k~ -~, T=To, (1.3.20)

where the exponent , ~/= 2 - (y/v), is zero according to classical theory , and remains small in more refined theory . This exponent is par t icular ly hard to measure and we shall not discuss it further.

Page 24: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

24 R.A. Cowley

In the distorted phase, the symmetry is broken so that the different susceptibility functions are neither independent nor equal. I t is, however, possible to choose basis functions from the Q,(k) in terms of which the susceptibilities are independent. In BaTiO3 with a ferroeleetric c axis the dielectric susceptibility in the ferroelectric phase is different for the a and c axes as illustrated in fig. 1.8.

In the distorted phase each of the different susceptibilities is divergent as T-+T~, and the nature of tha t divmgence may be described by eqns. (I.3.16)--(I.3.20) except that the amplitudes C+ and f+(~) must be replaced by C i_ and fi_(a). In all but exceptional circumstances (cf. footnote to Table II.1 in par t II) these divergences will be described by the same exponents (~ and v) as characterize the divergences when T o T + . Note, however, tha t if a continuous symmetry is broken at the phase transition, a more careful t rea tment of the low-temperature phase susceptibilities is required (§§ 1.4.2, 1.4.4).

[.3.5. Soft modes and the dynamic susceptibility The imaginary part of the dynamic susceptibility is related to the static

susceptibility by the Kramers-Kronig relation, eqn. 0.2.27), according to which

x~(O) = Zl(oy)co, 1 d~' - - c o

The imaginary part of X'(co) gives the power dissipation of tile system, and depends upon the matr ix elements coupling the initial and final states of the system and on the density of initial and final states whose energies differ by hco. The static susceptibility diverges at T c for k=O. Since it is unreasonable to assume that the matr ix elements are very temperature dependent, this divergence must reflect a movement in the weight of xl(co) to smaller co--the 'critical slowing down' discussed by Van Hove (1954).

Dynamic phenomena have played an especially important role in the develop- ment of our understanding of structural phase transitions. In particular the appealing concept of a soft mode, introduced independently by Anderson (1960) and Cochran (1960), has frequently proved helpful in describing struetural phase transitions, and in illuminating their essential characteristics.

The idea was prompted by the form of the static susceptibility of a harmonic oscillator, given in eqn. (I.2.23) by Xn(qjj). Comparing this result with the behaviour of the static susceptibility in the classical theory (eqn. (I.3.19) with ? = 1) suggests that

co(q~j)2=a(T--Tc) , T > T¢. (1.3.21)

Thus, in the soft mode approach, the phase transition is viewed as an instability of the crystal against a particular normal mode, the frequency of which decreases to zero at T~. In the low-symmetry phase the crystal is distorted so as to stabilize this mode, and the distortion is described by a 'frozen in' amplitude of this normal mode.

As an illustration of the soft-mode concept the result of inelastic neutron scattering measurements probing the soft mode in SrTiOa are shown in fig. 1.9. There is clearly a peak in the scattering whose frequency decreases towards zero as T--+T¢. In fig. I. 10 we show the square of these frequencies plotted against temperature and the agreement with eqn. (I.3.21) is clearly excellent. Below T¢ the triplet soft mode splits into a singlet and a doublet and rises in frequency as described in § 1.4.2. Similar results have now been obtained in many materials undergoing structural phase transitions (Scott 1974, Shirane 1974).

Page 25: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 25

Fig. 1.9

300 L ' , , , , -

2 0 0 ~ / / o ~ 200°K

~,ook- '%~ \

o] A : oor/ /Z ___ \

o II L ~ o "-"2oc II I \ 92 K

H I°° ,>.~ ,oc II ! '\ \

0 0.5 1.0 1.5 2.0 2.5 FREQUENCY,v (1012 Hz)

Neutron scattering observed at three temperatures from the sot~ mode in grTi03. The Bragg reflection below 110 K is shown on a reduced scale. Below 110 K the triplet soft mode splits into a singlet and a doublet. (From Cowley et al. 1969.)

2.5

Fig. I. 10

TEMPERATURE DEPENDENCE OF F25( ½ ' ' ~-~ ) MODE IN STRONTIUM TITANATE

u ] I I J

1.5 %

212

C~

_o i.o

%

0.5

I00 150 200 250 300

TEMPERATURE °K The tempera ture dependence of the frequencies of the soft mode in SrTiO3 both above and

below T¢. (From Cowley et al. 1969.)

Page 26: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

26 1%. A. Cowley

Despite these successes, the soft-mode theory is in a strict sense inconsistent. The frequencies of the normal modes in a harmonie crystal are tempera ture independent. Fur thermore soft modes are found to be damped and sometimes even overdamped showing also tha t the harmonic theory is inadequate. A soft-mode frequency must therefore be interpreted as some type of effective or renormalized frequency as we shall discuss further in § 1.5.1.1.

Phenomenologically the damping has been incorporated by approximat ing the susceptibility by the susceptibility of a classieMly damped oscillator.

Z l(~J, co)=co(~)2-co2-2icoTo, (I..3.22)

where 7o is the damping constant. The poles of the susceptibility in the complex co plane are given by

ca +-= +_(co(qj)2-Y~)l/2 +i7o, ca(qj)>7o,

2 • 2 1 / 2 (I.3.23) = i ( 7 0 ± ( 7 0 - c a ( q 3 ) ) ), ca(qj)<7o,

showing tha t the soft mode is underdamped when co(q j) > 7o but overdamped when 7o > co(q j). I f 7o is t empera ture independent the mode always becomes overdamped as T-+ T¢, and the pole of the complex frequency susceptibility approaehes the origin of the frequency plane along the imaginary co axis, as shown in fig. 1.11.

Complex co Plane

Fig. I.ll

Im[~l

a ] Undamped Oso//ator

J i i

) )

b] Damped Oscff/ator

( <

I C] Oebge Oso//a~or

Real [co] The complex frequency plane co, shewing the motion of the poles as T-~T c for a classical

oseillator, both (a) undamped and (b) damped, and (e) for a Debye relaxation response.

When the damping becomes very large the mode is overdamped and o) 2 may be neglected compared with 202.o in eqn. (I.3.22): the response is then given by

X- 1 (qjj, co) = co(~')2(1 -icazD), (I.3.24)

Page 27: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 27

where ~D is known as the Debye relaxation time and is given by l; D = 2?0/¢o(qj) 2, and so diverges as T-+To. In this case the pole of the susceptibility in the complex co plane is on the imaginary frequency axis, fig. I. 11, and the measured response is concentrated at very low frequencies.

A continuous phase transition involves a divergence of the static susceptibility at To. This is always associated with the approach of a pole of the susceptibility to the origin of the complex frequency plane, as T approaches T¢. This pole can be called a soft mode even when it is always heavily overdamped, so that a continuous phase transition always has a soft mode. Alternatively, the use of the term soft mode can be restricted to those cases where the mode shows at least some vestiges of resonant behaviour, the frequency of which decreases as T--~ T¢. This restriction would confine the use of the term to the displacive class of materials such as SrTi03 as illustrated by figs. 1.9 and I. 10. In the order~tisorder materials, the modes describing the small oscillations of the atoms or molecules are not strongly temperature dependent as illustrated in fig. 1.12 for the optically active modes in NaNO2. The ferroeleetric

A N "1-

O

T E M P E R A T U R E

IN NoNO 2 BY

>- (.) Z LL~ Z) 0 ILl

b-

Fig. I. 12

DEPENDENCE OF MODES BARNOSKI AND BALLENTINE

i I I

ITc

EIIc

Ellb

Ella

I ,;o ,;o ' 200

T E M P E R A T U R E (°C)

The frequencies of the infra-red active modes in NaNO2 as a function of temperature (Barnoski and Ballentyne 1968).

fluctuations in NaNO2 are associated with the relatively slow reorientation of the NO2 groups and, as shown in fig. I. 13 by the dielectric susceptibility measurements, the anomalous behaviour of the susceptibility is entirely absent for probe frequencies in excess of 109 Hz. The soft mode in this ease is not one of the normal modes of

Page 28: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

2 8

E'

7OO

R. A. Cowley

Fig. 1.13

o 32 MC/sec q

600

500

aC=784 k£~

4 0 0

• 64iviC/sec .

300

200 , i s o M c / ~ E"-'x,,.~,

0 900OMc/secl- ~ . . . . ~ °, T

f60 ,70 ,80 ,90 200 2,0 °c

The dielectric susceptibility of NaN02 as a function of frequency as measured by Hatta (1968) after Yamada et al. (1968).

vibration describing the small oscillations of the a toms and molecules, but one of the modes described by the spin variables introduced in § 1.2. I. Since this hopping is not t reated in the harmonic theory of lattice dynamics we prefer to restrict the use of the term 'soft mode ' to displacive materials. Alas, since there is not a clear distinction between displacive and ordei~disorder materials, this inevitably leaves a number of materials, such as K D P , where it is uncertain if a soft-mode description is

appropriate. In this section we have discussed the dynamical susceptibility in terms of

phenomenologieal models which have one or perhaps two simple poles in the complex 09 plane. In § 1.5.1, and in pale I I I , we shall discuss a more general approach to this problem. Nevertheless, a characteristic feature of the phase transit ion is tha t the dynamical response becomes slower as T---~T c. I t is often convenient to describe the fluctuations of the order pa ramete r by a characteristic frequency coo where

1 ('~oQ 1 1 - J - - I m [g(qJJ, co)] d~o =~z(qJJ) .

On comparing this definition with eqn. 0.2.27), it is clear tha t ~oe-*0 as T--*To no ma t t e r how complicated the spectral response.

Page 29: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 29

1.3.6. Measurement of ,susceptibilities

1.3.6.1. Macroscopic measurements Macroscopic measurements probe the susceptibilities tbr 14 = 0, and consequently

measure the susceptibility for the pr imary order parameters only if q~ =0. The dielectric susceptibility can be measured by dielectric measurement or by infra red absorption techniques. The susceptibility can be obtained from eqn. (I.2.21) if both operators A and B are taken to be the dipole momen t operator M. ] f only one mode is tempera ture dependent then

l z~p(~o) = z $ + 2%(Oj)M~(qi)z(Ojj, ~), (1.3 25)

Y

where X~°~ is the dielectric susceptibility of the other modes and the electronic excitations. Static dielectric measurements then give a measure of the soft-mode static susceptibility as illustrated in fig. [.8 for BaTiO3. Using electrical techniques it is possible to measure the dynamical susceptibility for frequencies up to 5 x l 0 i° Hz as illustrated in fig. 1.13. In principle, a measurement of the infra-red reflectivity gives the dielectric response at higher frequencies and in fig. 1.14, we show the dielectric response of BaTiO3, as measured using these techniques. In practice, however, these techniques are difficult to apply because the range of frequency of most interest in displacive systems is frequently between l0 l~ and l0 lz Hz, and this is a vel3; difficult range experimentally. Secondly, the measure- ments yield the reflectivity, and when the susceptibility is large, the reflectivity is close to unity and large differences in the susceptibility give only very small changes in the reflectivity.

Fig. 1.14

1 0 0 I I 1 I I 2 4 o C 1 i I

- - - - - - IO0°C E i t o £ a x i s .............. 127°C - - . - - 145°C

95 ....... 200°C

o~ 72.. ........... O points from millimeter and >2 ~ e...... ,, microwave measurements _~ " ~ ' " - . ,

F- 9 0

b - b ]

,-: . "7 . , , . . "~

8 ( i I i I I I I [ I 0 I0 20 50 40 50 60 ?0 80 90 I00

FREQUENCY, cm -'1

The refleetivity of BaTi03 as ~ function of temperature. (After BMlentyne 1964.)

Ultrasonic measurements a t s tructural phase transit ions have been reviewed by Rehwald (1973). When the strain is a pr imary order parameter , the ultrasonic velocity plays the role of e)(q~j) 2, and hence the velocity decreases to zero as T ~ T c as shown for K D P (KH2P04) in fig. 1.15. More frequently the strain is a secondary order parameter and the ultrasonic velocity measures the inverse of their correlation function whose 1)ehaviour (to be discussed in more detail in I 1.5.1.2) is typified by the

results shown in fig. 1.16.

Page 30: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

30 R. A. Cowley

~J

2 o

I,,- Z < 4

7

0

~J

7 -

b -

$ -

Fig. I. 15

q-2 , = 0

(,,,322,)0 (]- 2. (.. E :o

" %

. 4 3 O

-2 o

- I J , : J J J ,

121 1 2 2 " 0 122'~1 I '22 2 " "I

i I I | ' I 7 0 130 14O

Io i O T¢ IEMpERA|URE(OK)

The elastic constants ofKDP, C ~ 6 , and C~6 = 6 ' 6 6 a s a function of tempei'ature as measured by Brody and Cummins (1968) using Brillouin scattering techniques.

Ultrasonic experiments also measure the a t tenuat ion of the elastic waves; the critical behaviour of the a t tenuat ion is discussed briefly in III .3.2.2.

The specific heat may also exhibit singular behaviour near a structural phase transition. In practice the change in the specific heat is frequently very small and difficult to measure accurately as shown in fig. 1.17 tbr SrTiOB.

1.3.6.2. Scattering measurements' By far the most detailed information about susceptibilities can be obtained using

the scattering of X-rays, neutrons or light. In the case of X- ray scattering, it is not possible to obtain sufficient energy resolution to measure the dynamic susceptibility; consequently the cross-section for a wave-vector transfer K is given in the one- phonon approximat ion by eqns. (I.2.38) and (I.2.42) as

da(K~) = k B T ~ F(K, qj)F(-K,-qj')x(qjj')A(K+q). (I.3.26) d~ q z '

Page 31: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

S t r u c t u r a l p h a s e t r a n s i t i o n s 31

Fig. 1.16

I I I i ~ I g O 1XOl X X

8.0 xOO

t 7.8

I o ×

7/-, x o • + o ~ 5.1 dpo4 o o l o ÷ o e

x" i R,,[ool] 7.2 x t,.9 c;

X xg ~ .

+ + + ~ o 4.7""

+ . ~ o p °o

J i = A i ~ ~.5 loo 110 120 T[*K]

Velocities of longitudinal and shear waves propagating in a [!00] direction in SrTi03 from Liithi and Moran (1970). Key: (S) shows results at 50 MHz; • at 30 MHz; × at 16 MHz; and + at 10MHz.

I 73

Fig. 1.17

O.13cal/mol K . ~

I I I I I I I I ~8 7O0 702 70~ 106 7O8 770 772 17~

T(K) - - The specific heat of SrTiO3 near the 110 K phase transition. (After Franke and Hegenbarth

1974.)

I f we separate ou t the term involving the order pa rame te r susceptibilities zi(k)

da(K)

d ~ = ( ~ B ( K ) + k B T ~ l F ( K , q s , + k j l ) 1 2 z i ( k ) A ( K + q s i + k ) , (I.3.27)

i k

where aB(K) represents the o ther terms which produce a background scattering. The

Page 32: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

32 R . A . Cowley

experimental results are frequently interpreted by neglecting the K-dependence of the structure factors F(K, q j), when the observed intensity is, after subtraction of the background, proportional to the susceptibility. This approximation is probably reasonably accurate for small k and can be tested by measuring zi(k) for several different wave-vector transfers K giving equivalent k.

There are two different techniques which have been used to determine the X-ray scattering cross-section. In both cases it is necessaJ T to use monochromatic radiation to reduce the background. The technique, developed at Orsay by Com6s and his collaborators (Com6s et al. 1970) employs photographic detection and hence is invaluable for a rapid qualitative survey of the nature of the scattering. In fig. I. 18

Fig. 1.18

(a) (b)

(c) (d)

The diffuse X-ray scattering in KNbO3: ,(a) cubic phase at 770 K; (b) tetragonal phase at 520 K; (c) orthorhombic phase at 220 K; and (d) rhombohedral phase at 220 K. Note the disappearance of the streaks of diffuse scattering as the material changes phase. (After Combs et al. 1970.)

Page 33: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 33

we show a typical resnlt obtained for KNbO3 which undergoes very similar transitions to BaTiO3. Notice the diffuse streaks joining the Bragg reflections. These show tha t xi(k) is anisotropic, its value being considerably larger for the wave- vectors, k, along the [100] directions than for k along other directions. The other technique involves the use of counters and so in principle provides more quant i ta t ive information but is slower. In fig. I. 19 we show the diffuse scattering obtained using counter techniques in NAN02 by Yam ada and ¥ a m a d a (1966). This technique is capable of very high resolution in wave-vectors ~ 10 4 A - 1 and so can be used to give very accurate measurements of the exponents, 7 and v. I t is ra ther surprising that there has not been more effort made to exploit this technique in the field of structural phase transitions. (There is, however, some difficulty in constructing accurately controlled tempera ture environments for specimens for X- ray scattering.)

Fig. 1.19

NaNO2 TN ÷ 3K

(A) ~

C

r"

o o.~ o12 05 o~ o~s o 0.05 010 0 005 040 015

-o5

c_

C3 -0 6

/ / 0 01 0.2 0 3 0.4 0.5 OOS 010

q, CL

[ o o~s o',o o~s

q~

The diffuse X-ray scattering observed in NAN02 at T c + 3 K for scattering for K = v + ~aa*, K=v+l/8a*+~2b* and K=~+l/8a*+~3c*. The lower part of the figure shows J(k)=-D(q) calculated using the electrostatic approximation. (From ¥amada and Yamada 1966.)

I t is frequently possible to measure the differential cross-section using neutron scattering, since neutron energies are comparable with the energies of the phonons unlike X- ray photon energies. However , this feature leads to difficulties in the determinat ion of the total scattering cross-section using neutron scattering. In a two- axis spectrometer the neutrons scattered in a part icular direction are recorded irrespective of their energy. As shown in fig. 1.20 neutrons with different energy transfers have different wave-vector transfers. The integrated intensity is not then an integral over all frequencies for a particular wave-vector transfer, but an integral along a part icular pa th in (N, co) space. This will correspond to the true integrated intensity only if the var ia t ion in wave-vector transfer N, is negligible for all the

Page 34: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

34 R . A . Cowley

Fig. 1.20

• - k 2 •

;o

Reczproca/ Space

/ Path m

A ([-£],o~)Spaee

Sketch in reciprocal spuee of a two-~xis neutron diffraction experiment. Note that the length of k: is not fixed. In the lower part of the figure the path of integration in Ikl, co space is sketched.

frequency transfers occurring in the experiment, &. This can be achieved when the energy of the incident neutron beam is very much larger than h59. In principle, the accuracy of the integration can be tested by performing a differential scattering experiment in which the spectrometer is controlled so as to obtain a frequency distribution tbr constant wave-vector transfer, 14, but this requires more complex experimental equipment.

One of the cases where there is no difficulty with the wave-vector dependence in the integration is in o rde~disorder systems where the energy transfers from the diffuse scattering are negligible. In fig. 1.21 we show the diffuse scattering of neutrons, due to the ordering of Cu and Zn a toms in brass as measured by Als-Nielsen and Dietrich (1967). The scattering becomes more intense as ToT~ and also more concentrated in k-space.

The differential scattering cross-section is given in the one-phonon approxim ation by eqn. (I.2.35). I f the scattering from the fluctuations in the order paramete r are well separated in frequency from the other modes of the same symmetry , then

N t (K, co) = ~, IF(K, k + q~)I2A(K + qs + k)Sf( K, co), (I.3.28) g

where from cqn. (I.2.4)

h ,~(K, co) = -(n(co) + l)z~(k, co). (I.3.29)

u

I f F(K, k + q~j) is independent of k, then the scattering will have the same form at different wave-vectors K corresponding to the same k, and will give a direct measure

Page 35: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transi t ions

Fig. 1.21

35

1 p_

I m ~ y ) -2 VlmlY ) 0 0 -1 0 0 0 0

0 0 0 o 0 0 0 0 o 0 0 o 0 O 0 --1 0 0

--2 0 0 0

ot Counts

-3000 ~ 0 Measured Intensities ImlY)

/ "Oxo~u|t line I~'I(Y' with XI= 6"1'5"10"2~- 1

=00 / 000

- - I I I , I

-20 -15 -iO -~i 0 ~i 10 15 20-10"2 ll~-I Missetting from Brogg Peak y ,,-

The neutrons scattered from fl-brass 8"9 K above T¢, and a fit to a Lorentzian cross-section as a function of ]c, or misset from the Bragg peak. (From Als-Nielsen and Dietrich 1967.)

of z{(k, (9) as shown in fig. 1.9. If, however, the frequency of the fluctuations in the order parameter is not well separated from the frequencies of the other modes of the same symmetry, the interference terms in eqn. (I.2.38) become important and the shape of the observed spectrum for different but equivalent K become different. We shall examine this point in detail in § 1.4.1.

Neutron scattering then provides a very powerful way of studying the critical fluctuations at strueturM phase transitions. Normal techniques enable an energy resolution of about 5 x 10 I° Hz and a wave-vector resolution of about 10 -2 A 1 A novel technique employed by Collins and Teh (1973) to measure the differential cross-section for orde>disorder alloys was to measure the decay of the fluctuations in real time. This technique is applicable for phase transitions where the charac- teristic frequency is about 10 -2 Hz.

Raman scattering is restricted to very small wave-vectors K (effectively K = 0 in most eases) but enables detailed measurements to be made of the frequency dependence of Z(K, co). There are two distinct eases which must be discussed. Firstly there are cases, such as KDP, for which there is a linear coupling between the light and the order parameter fluctuations (the first term in eqn. (I.2.46)). I f this is the only mode of importance in the sum over j then the g a m a n scattering tensor is given by

I ~ = P~e(Kj)P~.~( - Kj)~(K, o)). (I.3.30)

The results of the P~aman scattering in KI )P are shown in fig. 1.22, and show that the spectral width becomes progressively narrower, as T o T e , showing the phenomenon of etqtieal slowing down.

This type of result can olfly be obtained in materials for which the soft mode is g a m a n active in the high-symmetry phase.

I f q~=0 but the mode is inactive it is possible to break the Raman scattering selection rules by applying an external field to the material and measuring the field

Page 36: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

3 6

I 0 0 0

6 0 0

R. A. Cowley

Fig. 1.22

3 0 0

2 0 0

- - T > T c

- - - T < T c

I 0 0

g 60

~ 5 0

~ 2o ?

I0

125°K

2 9 6 ° K

1/11 I \ I \ \ / \ \

6 / to3oK\ \ / \

/ \ / I \

3 -" I I \

2 /I I \ / \ \

/ 82oK\ \ / \ \

- ] 40 - I 0 0 - 6 0 -20 0 20 60 I 0 0 140

co ( c m - I )

gaman scattering from the ferroeleetric fluctuations of KDP. (After Kaminow and Damen 1968.)

induced scattering. This technique has been used to make the infra-red-active (but Raman-inact ive) modes visible in the R a m a n scattering from SrTiO3 and KTaO3 (Fleury and Worloek 1968).

The two-phonon g a m a n scattering has a contr ibution arising from the depen- dence of the polarizability on

Close to the phase transit ion we consider the part icular par t of this te rm for which the normal mode coordinates are those of the order parameters to obtain

k~k~ii" \ 3i'

q~-, k2 -q/i' |2n|(O,(kl,r)0g_kl,r) x P ~ a 3r x Or(k/ , O)Qv( - k2~ 0))~ exp (i~o~) d~. (I.3.31)

I f we neglect the wave-vector dependence of the coefficients, this cross-section gives a measure of the frequency dependent four-point correlation function. In the

Page 37: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Struc tura l phase t rans i t ions 37

interpretat ion of R a m a n scattering results this correlation function has almost invariahly been approx imated by writing

(Qgk, . t)Qi( - kl. t )Oi.(k 2 . O )Qi. ( - - k 2. 0)}

~ 4 ( Q i ( O ) } ( Q , , ( O ) } (Q,(O.T)Q,,(O.O)}. (I.3.32)

The approximat ion (I.3.32) should be reasonable when Qi is large, but will fail when it is comparable in size to the fluctuations Qi(k), i.e. close to T¢. With this approximat ion the Raman scattering becomes i(2) * = f l y a

1<2> _ N X ~ o fq*i -q~i'~p fq,i - .qJ~ ~a>,a~ / _ , l ~ e I . 9~i((Lo))Q 2. (1.3.33) , ,,9, J, J " \ J , a, /

In fig. 1.23 we illustrate R a m a n scattering results from SrTi03 for T-+T¢; there is clearly a soft mode whose frequency agrees with tha t obtained from neutron scattering measurements below T c (fig. 1.10) showing tha t R a m a n scattering provides a valuable way of measuring ~<ai(O,(o ) in the distorted phase. The approximation of cqn. (I.3.32) should be reasonable when (Qi} is large, but will fail when the fluctuations in Qi become comparable with (Qi}--i .e . close to T¢ (see § III.3.2.2). The analysis of R a m a n scattering results close to T¢, in terms of eqn. (1.3.33) is therefore inappropriate. We suspect tha t some of the difficulties experienced in the analysis of results close to T¢ derive from this erroneous use ofeqn. (I.3.33).

Fig. 1.23 5O

4 5

"7

¢9

z

3c Ld

O W fE

U_ 2 0

x:

I 0

0 o Zo 4'o 6b 8'0 ~6o do

TEMPERATURE IN "K

Frequencies of the two soft modes in SrTiO3 below T e as observed with Raman scattering techniques by Fleury et al. (1968). A is the transverse mode and D the longitudinal mode.

1.3.6.3. Local suscept ibi l i t ies In § 1.3.3.3 we discussed the measurement of order parameters using electron spin

resonance and nuclear magnetic resonance techniques. In both cases the method involved determining the resonance frequency of either a part icular electron or nucleus in the electric field produced by its environment. This environment is not static but fluctuating, and hence there will be a line width of the resonance which is

Page 38: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

3 8 R. A. Cowley

determined by the rate a t which the spin (eiectronic or nuclear) can lose its energy. This rate can be calculated and within second-order perturbation theory will depend upon the density of states of the same frequency as the resonance frequency, and the coupling between the spin and these states.

If we extend the expansion eqn. (1.3.14) of the coefficients of the electrostatic potential l3 to include the fluctuations in the order parameter then, remembering that B is a local quantity,

where A , (O) is the A, of eqn. (1.3.14) and we have assumed that there is only one component of the order parameter. The line width for a resonance frequency o is proportional to (Blinc 1971)

where co is the resonance frequency and B(z) is the operator B a t time z. Substituting the first term of eqn. (1.3.34) for B(z) yields

where Y(k , o) is the one-phonon Van Hove scattering function for the fluctuations. This expression has a sum over k owing to the local nature of the resonance technique. Nevertheless, an increase in the line width is found close to T, as illustrated in fig. 1.24 for the 133Cs N.M.R. resonance in CsD,AsO, (Blinc 1971). Unfortunately, since these experiments determine a sum over k (for one particular frequency, w ) i t is frequently difficult to interpret the results without relying on a particular model fbr Y (k, w ) . Furthermore, below T , the resonance line is split by the distortion and it may be experimentally very difficult to distinguish between the splitting and a line width effect (cf. 9 111.4.3.2). Finally if d (0) of eqn. (1.3.34) is zero, the line width results partly from the two-phonon terms in eqn. (1.3.34) and partly from the terms involving A , (k) z k2. The line width incwase a t T , is then small and in practice almost impossible to interpret quantitatively.

The Debye-Waller factor (eqn. 1.2.35) is also a measure of single site properties. To the extent that the motions of the atoms can be described by lion-interacting normal modes (i.e. in the displacjve limit) then

where the displacement-displacement correlation function is

The normal-mode correlation function is given a t high temperatures in terms of ~ ( q j j ' ) by eqn. (1.2.22). Since the susceptibility diverges as T-T, for q+q,, W , will increase as T-+T,. This increase will, however, be less marked than that of the susceptibility ~ ( q , j j ) due to the sum over q j and j' in eqn. (1.3.36).

Page 39: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions

Fig. 1.24

39

+,

o o o o

o o ° ° ° o o

Cs0~As0~ T~--P C$ m v = 6,/, Nliz

o o o o

~ o

$ ° J "

• o o

0

0

0

0

0 0

I t I

a l

I I I I

o i

o l

i ° l

° I I" I I I,

I 1 I I

2.5 ] 15 ~ 63 T:

1T~['K ]

Temperature dependence of the 13 3Cs spin-lattice relaxation r~te in CsD2AsO 4. (After Bhnt 1971)

The behaviour of the Debye Waller factor in an order-disorder system is different. Consider as an example the N ion in NO2 which has two positions _+g for the equilibrium position. In the high-temperature phase, both sites are equally populated so tha t the Debye-Waller factor is given by cos 1~.5. In the low-temperature phase the amplitude of the scattering is propo~¢ional to cos K. ~+2i<S} sin N. 5, so tha t the Debye-Waller factor due to the hopping is given by

exp ( - W~)= (cos 2 (K. g)+ 4{S> 2 sin 2 (K. 5)) 1/2. (I.3.37)

This shows that the Debye Waller factor is constant above To, and increases below T~ reaching unity when (S> =½. Since most real materials are neither strictly order- disorder nor displaeive in character, the interpretat ion of the temperature depen- dence of Debye-WMler faetors is far from unambiguous.

The Debye-WMler factor can be measured, for a particular 14, from the amplitude of the M5ssbauer 7-rays emitted from a suitable source. In fig. 1.25 we show the Debye-Waller factor of 57Fe in PbTi03 as measured using MSssbauer

Page 40: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

40 I~. A. Cowley

techniques by Bhide and Hegde (1972). The Debye-Waller factor can also be measured using the incoherent scattering of thermal neutrons or from the temperature variation of the intensity of Bragg reflections.

Fig. 1.25

5 0 "

2 5

o

z 20

o

1 5

I o | i , i i i i i ~ i i i i i

0 150 500 450 600

TEMPERATURE °C

Temperature variation of the Debye~Waller factor in PbTiO3 as measured from the resonance area of the STFe MSssbauer line by Bhide and Hegde (1972).

§I.4. PHENOMENOLOGICAL LANDAU THEORY

1.4.1. Single-component order parameters

1.4.1.1. The Landau theory In 1937, Landau (1937) developed a theory of phase transitions which he applied

to structural phase transitions. In this section this theory is applied to a wide variety of different systems, because the first step in the understanding of any new material is the development of the appropriate Landau theory. We begin with the simplest familiar (Landau and Lifshitz 1959) example when the order parameter is a singlet so that there is only one component ofQi = Q of§ 1.3 which can then always be chosen to be real. Landau assumes that the free energy of one unit cell of the crystal may be expanded in a power series in this order parameter;

G = Go + lrQ2 + dQ 3 + uQ a, (I.4. l)

where the coefficients may be functions of temperature. A phase is stable only if

~ e o 02G 0G =0 , ~Q:Qo>0. (I.4.2)

The high-symmetry phase <Q} = Q0 = 0 is then stable only if the linear term in Q in eqn. (I.4.1) is absent and if r is positive. The system is distorted if r is negative.

Page 41: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 41

Fur thermore at a continuous phase transit ion G must increase wi th [QI a t T¢, r = 0, so tha t d = 0 and u > 0. Landau ' s theory is then completed by the fur ther assumption tha t since r changes sign a t To, it can be wri t ten as an analyt ic function of T near T¢

r = a (T - T¢),

while the coefficient u is assumed to be independent of tempera ture for temperatures sufficiently close to T¢. The free energy of the simplest model then follows as

l 2 4 G=Go +~a(T- T¢)Q +uQ . (I.4.3)

The stabili ty conditions for the low-temperature phase then give for T < T¢ tha t the order pa ramete r will have the value

a Q~= ~u(Tc - T), (1.4.4)

while the free energy for the equilibrium displacement is given by

a 2

G s = { G } = G o - ~ ( T ~ - T ) 2, T<T~ (I.4.5a)

and consequently the change in specific heat is given by

82G a2T C = - T s T 2 - 8 u ' T < T ¢ ;

= 0 T > T¢. (I.4.5 b)

/ 8 2 G \ - 1 The susceptibility x ( R = O ) = | ~ I , is then

1 x(O)=a(T_T¢), T>Tc;

1 -2a(T~-T)" T<Tc" (I.4.6)

These results give, by assumption, tha t the exponents describing the critical behaviour are fi =½ and ? = 1. In fig. 1.26 we show the free energy for this model both above and below T c, and in fig. 1.27 the various properties as a function of temperature .

In eqn. (I.4.3) we assumed tha t d = 0 . This is always the case when there is only one component of the order parameter , but is not necessarily the ease for multi- component systems, for example at some elastic instabilities (Cowley 1976 a).

When d is non-zero the transit ion is of first order and so the expansion of eqn. (I.4.1) is suspect. Nevertheless, in many systems the transit ion is nearly continuous and the phase transit ion can be analysed using the Landau expansion. The transit ion occurs when T = Tl where

d 2 - - (1.4.7) Tx = T¢ + 2an'

Page 42: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

42 R.A. C o w l e y

Fig. 1.26

5

W

-1 0 1 - I

"r~--A = To/

1 2 -1 0 1

AMPLITUDE Q

The free energy G-Go as a function of the coordinate Q for three models: Model A is the simplest g iven by eqn. (I.4.3) w i th a = 2 , u = 4 and A is one in these units. Model B incorporates the cubic term d which was chosen to be 6 when the first-order transition is at T¢ + A. Model C also has a first-order transition but in eqn. (1.4.9) a = 2, u = - 8 and h = 6 when T I = T c + A .

Fig. 1.27

© o

E£ LU

LLI

<

LU r-i O:z ©

ORDER SUSCEPTIBILITY

- - © o - - - ~(o~ ~

I ~ ~ --

10

0 Tc-A T c Tc+A

TEMPERATURE

PARAMETER AND iNVERSE

5

- - 0 >-

2O O3

15 I~ n LLJ

10 (.9 CO

5 09

0 ILl CO

LU

Z

5

The order parameter ($o and inverse susceptibi l i ty as functions of temperature for the three models described in fig. 1.26. Note that the inverse susceptibil i ty increases very rapidly for T < ~ I for model C.

Page 43: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 43

while the order pa r ame te r takes on a value a t f/'i given by

d Q 0 - 2u '

and below T I is given by

- 3d - (9d 2 - 16ua(T - T¢)) 1/2 Qo = (I.4.8)

8u

In figs. 1.26 and 1.27 we i l lustrate the propert ies of this model in more detail. I n this case the first-order t rans i t ion occurs whenever the s y m m e t r y allows a non-zero cubic term, d. I t is therefore a sym m e t ry -d r i ve n first-order transit ion.

A first-order t rans i t ion also occurs in the simplest model i fu < 0, bu t in this case it is a consequence no t of the s y m m e t r y bu t of the par t icular sign of the u coefficient in the L a n d a u expansion. I f u is negat ive, G is unbounded and the expansion m u s t be extended beyond quart ic terms. We therefbre add an ext ra term to the fi'ec energy, when it is given by

G = Go + l a ( T - To)Q2 + uQ4 + hQ6 + . . . . (I.4.9)

where u is now negat ive and h is assumed to be positive. The transi t ion is of first order and occurs a t T1 where

~t 2

T~=T~+~ah, u < 0 . (I.4.10)

The step in the order pa rame te r at T r is

Q /a(T~ - T ¢ ) \ 1/z

while below T I the order pa rame te r is given by

- 4u + (16u 2 - 24ah(T - T~)) 1/2 (I.4.11 ) Q2 = 12h

The propert ies of this t ype of f irst-order transi t ion, of which K D P ( K H 2 P 0 4 ) is an example, are i l lustrated in figs. 1.26 and 1.27.

1.4.1.2. Coupling to 6'trains At all s t ruc tura l phase t ransi t ions the order pa ramete r is coupled to the elastic

strains or possibly other secondary order parameters by terms in the free energy which are linear in the s t ra in and quadra t ic in the order parameter . I n the simplest case we extend the model of the previous section by including the coupl ing to a single s train parameter , r/. The free energy, given by eqn. (I.4.3), is then ex tended to include the strains as

1 2 4 1 2 G = G o + ~ a ( T - T c ) q +uq +eq2q+~Cq + . . . , 0.4.~2)

where e is the coupling a nd C is the appropr ia te elastic constant . Close to Tc there is no reason to expect t h a t C will be par t icu lar ly t empera tu re dependent , except by an

Page 44: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

44 R . A . Cowley

unlikely accident. The condition fbr the material to be stress-free is tha t OG/3q =0 , which gives

eQ 2 q = - ~ - , (I.4.13)

which can be used to eliminate t / f rom eqn. (I.4.12) with the result

where

1 2 G=Go + ~ a ( T _ T¢)Q +~Q4, (I.4.14)

e 2

~=u-2~. (L4.15)

Since eqn. (I.4.14) for G has exactly the same form as eqn. (I.4.3) except tha t u is replaced by ~, the development goes through unchanged except for the replacement of u by ,~7. Equat ion (I.4.15) shows tha t ~ is less than u: consequently transitions under conditions of constant stress are more likely to be of first order' than ones a t constant strain.

The elastic strain is proport ional to the square of the order parameter , eqn. (I.4.13), and so using eqn. (I.4.4), is expected to be proport ional to T - T ~ . The susceptibility can be measured in two different ways: a t low frequencies the strain will be able to follow the order parameter , Q and the susceptibility is wri t ten Z °, but a t high frequencies the strain will be unable to follow and the susceptibility is wri t ten Z ~. For T>T¢ both susceptibilities are equal to (a(T-T¢)) -1, hut for T<T~

Z°(0) = [2a(T¢ - T)] - ~, (I.4.16)

while Z ~ (0) is given by ~20/~Q2 with the derivatives evaluated with q held constant when the result is

;/~(0) - ~ . ( I . 4 . 1 7 ) 2ua(Tc "1')

Since ~ is less than u, )/~ < Z °, al though they both diverge a t To. I t is also necessary to distinguish between the high- and low-frequency response

for the elastic constants below T c. The elastic constant under conditions of constant order parameter , Q, is ~ZG/Oq2 = C, whereas under conditions where Q can follow the changes in t I it is given by d2G/d~ 2 = C - e2/4u. Hence, a t low frequencies when the order pa~umeter can follow the elastic fluctuations, the elastic constant is predicted to decrease by e2/4u at the transition.

In the low-temperature phase there is also a non-zero susceptibility describing the response of the strain to a field which couples to the order parameter , Q. This is given by [~ZG/O~OQ]- r = 1/2eQo, and hence diverges a t To.

1.4.1.3. Fluctuations in systems with short-range forces The expansion of the free energy in powers of the order paramete r Q is not

sufficient to give a description of the wave-vector dependence of the susceptibility. We now extend the discussion to incorporate the effects of the fluctuations Q(k).

Page 45: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 45

Formally eqn. (1.4.3) can be extended to give

G =Go + ~--~ ~ R(k)Q(k)Q( -k) + ~ kl~k3

u(kl, k2, k3)Q(kl)Q(k2)Q(k3)Q(- kl - k2 - k3) + . . . .

where the coefficients in the expansion are wave-vector dependent and use has been made of wave-vector conservation. This expression for the free energy (effective hamiltonian) includes the interactions between the excitations with different wave- vectors and is difficult to solve as discussed in detail in par t II . Here the effect of the fluctuations is neglected except for the leading terms. This is then the L a n d a ~ Ginzburg theory and is equivalent to neglecting the interactions, the u term, except when at least two of the Qs are replaced by the static order parameter , Q0.

For the simplest model of a transit ion u(0, 0, 0) is non-zero a t To, and so we neglect the k dependence of the anharmonic coefficient u. R(0) is, however, small a t Tc and so we include in R(k) the leading terms in an expansion in k. Since if the order parameter has a single component qs is a 'special point ' within the Brillouin zone, there are no linear terms in k. Consequently for an isotropic system with short-range interactions the free 'energy becomes

G=G, + ~--~ ~ [a(T- T¢)+fk 2 + 12Q~u]t~(k)(~(-it),

where G s is the static par t of the free energy, eqn. (I.4.5 a). The static susceptibility for wave-vector, It, is given by;

Z(k)-l=[a(T-T~)+fk 2] for T>T¢,

and by

X(k)-~=[2a(T¢-T)+fk 2] fbr T < T ¢ . (I.4.18)

It is sometimes useful to define the inverse of these static susceptibilities as the squares of frequencies,

z(k) 1 =~)(k)2,

by analogy with the result for a harmonic crystal, eqn. (I.2.23). These frequencies may or may not correspond to the frequencies of resonant normal modes or 'soft modes' as discussed in § 1.3.5. In terms of these 'frequencies' the free energy may be rewritten as

1 G = Gs + 2N ~ t°(k)2Q(k)Q( - k), (I.4.19) -V

where

~o(k): =a(T- T¢) + fk 2 + 12Q2u.

The expansion coefficient f must be positive a t T¢ for the wave-vector of the distorted phase qs to be stable, so tha t to a first approximat ion it is independent of temperature . These assumptions then give the critical exponents for the correlation function, eqn. (I.3.18), v--0-5, and r /=0, and the line shape of the scattering as a Lorentzian in wave-vector , k, a t fixed temperature . This lat ter result is consistent with experimental results on, for example, fl-brass as shown in fig. 1.21.

Page 46: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

46 R . A . Cowley

The tempera ture dependence of the Debye-Wal ler factor can be obtained from eqns. (I.3.36) and (I.4.19) for a displacive system as, (Borsa and Rigamont i 1972);

z(k) = K~ - K 2 IT - T d ~/2 (I.4.20) k

where K 1 and K : are constants which depend upon a andf . This result shows tha t the Debye Waller factor does not diverge at T¢ but exhibits a cusp-like behaviour (see, however, §II.5.4.1). A similar behaviour is predicted a t other structural phase transitions except when the dispersion relation a t small k cannot be expanded in powers of k as we discuss below.

Only in cubic materials is the coefficient of k in eqn. (I.4.19) independent of the direction of k. More generally eqn. (I.4.19) becomes

~o(k) 2 = a(T - T¢) + 12Q2ou + ~ f~k~k~. ap

The critical scattering is then Lorentzian in wave-vector , but the width of the Lorentzian depends on the direction of the wave-vector as observed close to the a-fi transition in quartz by Dorner et al. (1974). The exponent is, however, independent of the direction of k, as tacit ly assumed in eqn. (I.3.18).

A Lifshitz point (Hornreich et al. 1975) is the point in the phase diagram such tha t for one or more directions of k, the t e rm in k 2 in eqn. (I.4.18) has a coefficient f = 0. This point then separates phase transitions to a phase described by q~ being a °special wave-vector ' from phases where q~ is incommensura te with the underlying lattice. These special points are discussed in more detail in § II.5.3.2.

1.4.1.4. Uniaxial ferroelectrics The formal expansion of the R(k) term in powers of/c 2 in § 1.4.1.3 is valid only if

there are no long-range ibrces in the material . The commonest example of a long- range force is the long-range electrostatic ibrce in ferroeleetric materials. In these materials the macroscopic electric field associated with the ferroelectrie fluctuations makes the response dependent on the direction of the wave-vector; there is a splitting between the longitudinally and transversely polarized modes. Taking acount of this electrostatic interaction, and taking the z direction as the unique ferroelectric axis, the free energy is given by eqn. (I.4.19) but with the frequencies given by

w(k): = a ( T - T¢) + ~f~pk~kp + g ~ + 12uQ~, (I.4.21)

where the coefficient g is given in terms of the dipole moment operator as

2, V~ o

and is independent of k if the wave-vector dependence of Ms(k) is neglected. The susceptibility or scattering cross-section, for wave-vectors k in the x and y

direction, is similar to tha t of the isotropic systems discussed above. For other directions of k the limit as k-+0 is not divergent even at T¢. This gives rise to a characteristic doughnut shape to the critical scattering, a slice through which is shown in fig. 1.28 as measured in D K D P (KD2PO4) by Skalyo et al. (1970). An analysis of this da ta enables the parameters f a n d g to be obtained as shown by Paul et al. (1970).

Page 47: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 47

Fig. 1.28

3"3 / ~ I [ ~ I T

3 .2

L 30 o~ 4~0 ~ ~ 4-

2.9 150

a , 8 KD 2 P 0 4

2.T (505) T = 225 °K

I I I , I t [ I 2.5 2.6 2.7 2.8 2.9 5.0 3.I

H

I P I 1 I

/ !

I I I I 3.2 3 . 5 5.4 3.5

The critical scattering intensity contours in the (h01) zone of KD2P04 from the ferroelectric fluctuations polarized along the z axis. Tc for the ciTstal was 221 K. (After Skalyo et al. 1970.)

The singularity in z(k) as k-*0 cannot occur in real systems: it is prevented by a more detailed t r ea tmen t of the electrostatic interactions including retardat ion or polariton effects, and by the effect of the crystal surfaces (Born and Huang 1954).

The Debye-Wal le r factor of a uniaxial ferroelectric has an even weaker singularity a t T¢ than systems with only short-range forces, eqn. (I.4.20). The result was obtained by Meyer and Cowley (1976) as

z(k) = K 1 + K21 T - Tel log (K31T-- Tel), k

where the Ks are constants.

(I.4.22)

1.4.1.5. Acoustic mode instabilities The stabil i ty of a crystal against macroscopic strains is most conveniently

discussed in the Voigt nota t ion for the strains:

t~x x = t~ , I~yy = 112, ~]zz = 113, ~]x r + ~r x = ~]6, ~x z -[- ~ zx = ~] 5, ~]rz -~- ~]zy = ~]4"

The elastic matr ix , eqn. 0.2.10), then has the form

H v

The crystal is stable (Born and Huang 1954) only if all the eigenvalues of Cp~ are positive, leading to stabil i ty conditions which are dependent upon the part icular crystal class (Cowley 1976 a).

The Landau theory of these transit ions then follows §I.4.1.1 but with the appropr ia te eigenvalue of the elastic constant playing the role of the paramete r r.

Since the point groups of the crystal classes have a t most triply degenerate irreducible representations, the order pa ramete r may have one, two or three components.

Page 48: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

48 R . A . Cowley

An example of a one component order pa ramete r occurs for an/']6, strain in the point groups 42 m, 422 or 4/mmm. The eigenvalue of the elastic constant matr ix is C66 and the symmet ry of the distorted phase is 2 ram, 222, or m m m respectively. The high-symmetry phase will be stable if C66 > 0 and a phase transit ion to the distorted phase will occur if C66 < 0. We therefore assume tha t C66/P = a ( T - T ¢ ) .

The fluctuations associated with the strains are the acoustic modes of vibration. The velocity of any part icular acoustic mode is given by a linear combination of elastic constants, but tha t combination is dependent upon the polarization and the direction of the wave-vector. In the case of our example, the largest fluctuations close to T¢ will arise from the modes with velocities determined solely by (C66/P) 1/2. Since C66 is the smallest eigenvalue of the matr ix Cp~ every other acoustic mode must have a larger velocity. Now in our example the only acoustic modes with velocities given only by C66 are those transverse modes with k along the [100] or [010] directions and with polarization vectors along [010] or [100] respectively.

The frequencies of the acoustic modes with k close to the [100] direction are given by

~ ( k ) 2 = a ( T _ T¢)k2 2 2 2 2 2 4 -t- kz) +f2kx + . . . . +glky +g2kz + f lkx(k r (1.4.23)

and for k nearly parallel to [010] by a similar expression with k r and kx interchanged. This expression shows tha t the critical fluctuations are confined to part icular lines in k space, whereas they are restricted to planes for uniaxial ferroelectrics.

At some phase transitions there are no acoustic modes with velocities given by the eigenvalue of the elastic constant matr ix. In these cases, for example a uniform compression, there are no fluctuations, but they appear in practice to always be of the first order.

1.4.1.6. Interaction between acoustic modes and ferroelectric fluctuations The macroscopic effects of the coupling between the strains and the electric

polarization in ferroelectric materials are described, for example, by Jona and Shirane (1962). In this section we shall emphasize the effect of the coupling on the fluctuations. In materials, like KH2P 04 , which are piezoelectric in the high- symmet ry phase there is a linear coupling between the strain and the electric polarization. Consequently the order pa ramete r is a linear combinat ion of the strain and polarization. I t is convenient, however, to extend the analysis of §I.4.1.1 by explicitly including in the theory both strain coordinates and the polarization. For the case of K H : P 0 4 we label the polarization along the z axis as Q1 and this is coupled to the strains ~xy. I f for simplicity we restrict k to the x direction then the dielectric fluctuations are coupled only to the t ransvers acoustic modes polarized along [010] and we label them Q2- The quadratic term in the expansion of the free energy is for these modes given by

1 G2 =2N ij 2-=1,2 Rij(kx)Qi(kx)Qj(]Cx)'

where

R i i(k) =a(T - T~) + f k z, R22 (]c) ---- C66k2/p,

R12 (k) = R21 (k) = kD.

Here D is given in terms of the piezoelectric constant h36 as D=h36 M z ( O ) (19v) - 1/2 where Mz(0 ) is the dipole moment operator of the ferroelectric mode, eqn. (I.2.20).

Page 49: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 49

The effect of the coupling is most readily seen by calculating the dynamical susceptibility within the harmonic approximation. The response to a field which couples only to mode 1 is given by

R22(k)_0)2 ZI( k' 0)) ~--~ (R1 l (k ) _0)2) (R22 (]c) _0)2) _ R12(k)2 • (I.4.24)

Now, for small co, this expression has the form

Z,l(k, O)= [ a ( T - T ¢ ) +fk 2] - 1

where the free Curie temperature is given by

_~ p D 2 T~ = T¢ + a~66 . (1.4.25)

The static k = 0 dielectric susceptibility diverges at the Curie temperature of the fi'ee crystal which is a higher temperature than the Curie temperature of the clamped crystal, T~. At high frequencies, 0)Z>>R2:(k ), eqn. (1.4.24) reduces to the suscept- ibility of the clamped c rys ta l - - tha t obtained by neglecting the coupling to the acoustic modes.

Similar results are obtained for the acoustic mode response. At low frequencies the response is given by

P ;~2 (k, 0)) - C~6 _p0)2,

where

p D 2 C6E6 =C66 a ( T - T c ) ' (I.4.26)

and the change in the velocity arises because the dielectric fluctuations follow the acoustic mode fluctuations and the crystal responds so as to keep the electric field, I:, constant. At high frequencies where the ferroelectric fluctuations cannot follow the acoustic response, the lat ter reduces to C66 and the electric polarization is constant in the crystal. This behaviour is illustrated in KD2PO 4 where ultrasonic measure- ments yield the velocity, C~6, which decreases to zero at To, fig. 1.15, while neutron scattering measurements at frequencies above the dielectric response give the relatively temperature independent, C66 (Paul et al. 1970).

Throughout this section we have assumed that the dielectric fluctuations are intrinsically temperature dependent. The same analysis can be made if we assume in that C66 is temperature dependent, while Rll(k ) is relatively temperature in- dependent. This behaviour then corresponds to an elastically driven ferroelectric transition of which lithium ammonium tar t ra te is an example (Sawada et al. 1977).

In a neutron or X-ray scattering experiment there may be a coupling to both the ferroelectric and acoustic modes, so tha t the cross-section is given by

~(K, 0))=NkBT Im [{IFI(K)I2(R2, (k) _0)2)

- ( F I ( K ) F 2 ( - K ) + F z ( K ) F I ( - K))Rlz(k)

+ IFz(K)I2(RlI ( k ) - 0)2)}/{(R11 (k) - 0)2)(R22 (k) - co 2)

-R12(k)Z}], (I.4.27)

Page 50: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

50 R . A . Cowley

where the wave-vector transfer K = k + v, with v a reciprocal lattice vector and F 1 (K) and F2 (K) are the structure factors of the modes. Now the unusual feature about this result is tha t since R12(k ) is proportional to k, the interference term between the scattering by modes I and 2 will change signwith k, provided that FI(K) and F z ( K ) do not also change sign. The scattering is then markedly asymmetric in k. This effect has been observed in KD2PO 4 by Skalyo et al. (1970). In KD2PO 4 the ferroelectric fluctuations are overdamped so that inelastic neutron scattering measurements recording the quasi-elastic response gives the scattering by the ferroelectric

1.3

1.2

I.I

1,0

0.9

0 . 8

0 .7

Fig. 1.29

I I i I I i I

KD z PO4 (510} T • 22,5 "K

w

\ ~ . 200 / /

I r I F I I

i i I I I I l I I

0.7

1 1 I f r I I I 4.7' 5.0 5.3

H

The critical scattering contours in the (hk0) zone of KI)2P04 from the ferroeleetrie fluctuations polarized along the z axis. T c for the crystal was 221 K and the upper part shows measurements by Skalyo et al. (1970) and the lower part ealeulations including the coupling of the ferroeleetrie and acoustic modes by Cowley (1976 b).

Page 51: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Str,actural phase transitions 51

fluctuations. In fig. 1.29 the highly asymmetric critical scattering observed around the (5, l, 0) lattice point is shown, together with parameter-free calculations of the pattern, performed using the theory illustrated above (Cowley 1976 b).

In non-piezoelectric materials the leading coupling between acoustic modes and the ferroelectric fluctuations is proportional to k2--there is then no difference above T~ between the clamped and free dielectric behaviour as shown in § 1.4.1.2, and the scattering is almost symmetric in /c. There are, however, interesting interference effects between the acoustic and optic modes in the differential scattering cross- section as a function of frequency co. These effects are particularly marked if the ferroelectric mode 1 is heavily damped so tha t its spectral response overlaps the frequency of the acoustic mode 2. In this ease the term in eqn. (I.4.27) proportional to [F2(K)[ 2 gives a Lorentzian, centred approximately on co2=R22(k); the inter- ference terms, proportional to F I(K)F2( - K), give two contributions--a Lorentzian centred at the same position and a term, asymmetric about that frequency, due to the imaginary part of R1 t(k). This asymmetric term gives rise to a characteristic line shape and a shift in the apparent frequency of the peaks in the scattering.

This type of behaviour, which is common to all anharmonic materials, is illustrated particularly strikingly in the uniaxial ferroeleetrie lead germanate, as shown in fig. I. 30. Near the lattice point (006) the ferroeleetric fluctuations have low intensity and only a well-defined transverse acoustic mode is observed. Near the lattice points (005) and (004) both modes scatter strongly and the interference terms give rise to a characteristic asymmetric behaviour at frequencies close to the acoustic

Fig. 1.30

/ ' ' ' b Ge 6,, , Fol-o.1,-O.l,OlA oo f 4 0 0

• o 30o ~ ~

+~ 200

0 100 u

0

600 ~ ~ ~ , ~ • o , O

4 0 O

2 0 O

0 i ~ i i i i

1 0 -1 -2 -3 -4

ENERGY TRANSFER [MeV] The neutron scattering observed from lead germanate by Cowley et al. (1976). The results

show the interference between the acoustic mode and a heavily damped optic mode resulting in the different speetrM shapes for the different measurements corresponding to the same wave-vector, k.

Page 52: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

52 R . A . Cowley

mode. Near (005) the interference enhances the high-frequency part of the scattering, and near (004) it has the opposite sign and enhances the tow-frequency part. Clearly as with the scattering shown in fig. 1.29 the detailed shape of the scattering is dependent upon the ratio F 1 (K)/F2(K) and so is qualitatively different around different lattice points, ~.

1.4.2. Multiple-component order parameters

1.4.2.1. The Landau theory The order parameter at a structural phase transition may have more than one

component (§ 1.3.1): for example, in BaTi03 and SrTi03 there are n = 3 components. Throughout this section we shall consider the case where the Qi are real. A necessary condition for this is tha t the soft-mode wave-vector qs, must satisfy qs = ~/2 or 0. A sufficient condition can be obtained with the aid of group theory: it is that the antisymmetrie square of the order parameters be zero. This condition appears to be the same as the requirement that

Vk(o(q~ + kj) k=0=0

which is satisfied by modes with qs=0 and for qs=~/2 ibr the symmorphic space groups, but for the non-symmorphic groups it is not necessarily true. For example, in the diamond structure the longitudinal modes at the [100] zone boundary do not have zero gradient. We shall discuss systems in which the order parameter is complex in §§ 1.4.4 and 1.4.5.

We begin by discussing a system with n = 2 components for which the free energy, analogous to eqn. (I.4.3) is

G=Go+;a(T-Tcl(Q~+Q~)+u(Q~+Q~12+v(Q~+Q~)+ .... (I.4.28)

The symmetry of the order parameter imposes the form of the quadratic term and the third-order term in the Q-coordinates must be absent if the transition is continuous. Equation (I.4.28) differs from that of the singlet model eqn. (I.4.3) in that there are two coefficients u and v. Equation (I.4.28) reduces to the form of eqn.

(I.4.3) if we write Q/= ~/Q (eqn. (I.3.3)), where the eigenvector is normalized ~ ~/2 = 1, i

when ~t = u + v(~ + ~) , and h replaces u in eqn. (I.4.3). The free energy of the ordered phase is a minimum if the constant ~ is a minimum,

eqn. (I.4.5 a), and this determines the nature of the ordered phase. There are three cases of continuous phase transitions:

I. v < 0 but v + ~ > 0. In this case ~t = u + v > 0 and the ordered phase has one of four domains (~1 = -+ l, ~2 =0 , or ~2 = -+ 1, ~1 =0).

II . v = 0 and u > 0. In this case h = u and the ordered phase may take up any orientation in the space of Q1 and Q2.

I I ] . v > 0 and ~ = u + ½ v > 0 . In this case there are four ordered domains with

These features, a multiplicity of domains and a structure which is dependent upon the particular values of certain of the interactions, are characteristic of all n > 1 systems.

Page 53: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 53

The analysis of the model foll6ws exact ly as in the n ---- 1 system for the order parameter , Qo, specific heat, free energy and susceptibility above T c. Some additional analysis is, however, necessary to obtain the behaviour of the suscept- ibility below To. The static susceptibilities are given by the inverse of the second derivatives of the free energy with respect to Q1 and Q2, evaluated a t the equilibrium

state. For a distorted phase of type I , with ~ 1 = 1 and ~2 = 0 the second derivatives of the

free energy yield the susceptibilities,

] and A

Below T¢ the ' longitudinal ' susceptibility is given by

)(~ (0) = [2a(T¢ - T) ] - ~, (I.4.29)

the same expression as for the n = 1 system eqn. (I.4.6), while the 'transverse" susceptibility is given by

u + v Z2(0)= av(T¢ - T ) ' (1.4.30)

which since this phase is only stable if v < 0 is always positive. The terms ' longitudinal ' and ' t ransverse ' , are used by analogy with the Heisenberg model of magnetism: Zl describes the fluctuations parallel to the order parameter~ and Z2 the fluctuations perpendicular to tha t direction'. For case I I , v = 0 and X2(0) is infinite for all f ~ T¢. Physically this is because the hamil tonian has continuous rotational symmet~sz in Q1 and Q: so tha t it costs no energy to rotate the order parameter . The existence of the infinite susceptibility is then an example of Goldstone's theorem (see. e.g., Wagner 1966).

In phase I I I , with ~1 = ~2 = 1/~/2, the derivatives of the free energy are as follows:

a2G 02G u + v OQ~ - OQ~ = 2a(T¢ - T)2u + v ' (I.4.31)

and

~2G u ~QlaQ2 = 2a(Tc - T)2u + v"

The susceptibility can be brought into a diagonal form as discussed in § 1.3.3, by considering the susceptibility for fields along the ordering direction, ZL(0)~ and transverse to tha t direction, )(T(0). Then ZL(0) is given by eqn. (I.4.6) or (I.4.29) while

2u+v ZT(0) = 2av(Tc -- T)' (I.4.32)

which is necessarily positive when phase I I I is stable. This analysis can be repeated for an n = 3 component system (Slonczewski and

Thomas 1970) when the free energy is

1 G = G o + ~ a ( T - T¢)(Q~ + Q~ + Q~)+ u(Q~ + Q~ + Q~)2

+v(Q~+Q~+Q~). (I.4.33)

Page 54: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

54 R . A . Cowley

There are again three possible phases:

I. v < 0; the ordered phase is tetragonal (for example ~1 = -~ l, ~2 = ~3 = 0) and ~= u+v. An ordered phase of this type occurs in BaTi03 and SrTi03.

II . v = 0; the orientation of the distortion is undefined. I I I . v>0 ; the ordered phase has trigonal symmetry with ~1= 3-~2 = 3-~3 =

3-1/x/3 and ~ = u + l / 3 v . This phase occurs in LaA103 for which the transition is otherwise similar to that occurring in SrTi03 (Miiller et al. 1968).

In all of these phases there is a longitudinal susceptibility given by

ZL(0) = [2a(T~ -- T)]- ~, (1.4.34)

and a doubly degenerate transverse susceptibility given, for phase i, by

u+v ;iT(0) = -- av(T¢ - T) ' (I.4.35 a)

and for phase I I I by

3u+v ZT(0) = 2av(Tc _ T)" (1.4.35 b)

Finally, we briefly consider the more complicated case of an n = 4 system. The most general form of the free energy is then

= Go + l a ( T - T¢) (Q2 + Q~ + Q~ + Q2) G

+ u ( Q 2 + Q 2 + 2 22 4 Q3 + Q 4 ) + v ( Q I + Q ~ + Q ~ + Q ~ )

2 2 2 2 W 2 2 2 2 +wl(Q1Q2+Q3Q4)+ 2(Q1Q3+Q2Q4)

+w3Q1Q2Q3Q4. (I.4.36)

In real systems not all the fourth-order invariants may be allowed by the symmetry of the problem as we shall discuss later for BaMnF 4 and Nb02.

There are five different possible phases of the most general n = 4 model and we list in table I. 1 a typical ~ vector, for each phase. The most stable phase is that with the smallest value of

i

Table 1.1. The ordered phases of the n= 4 model.

I (1000) u+v - - 1 l 1 1

II (1111)~ u+~v+8(wl+w2)+~w 3

1 1 1 I I I (1 u +--v +--w 1 100)~/2~ 2 4

1 1 1 IV u +--v +--w 2 (1010)x/2'- 2 4

1 1 V (1001)~- u+ v

,/2 2

Page 55: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Struct~tral phase transitions 55

The susceptibilities in the ordered phase may be calculated as discussed, above. In each case one finds a longitudinal susceptibility given by eqn. (I.4.34), and three transverse susceptibilities. These transverse susceptibilities are dependent on the anisotropic coefficients, but unlike the n = 3 systems, they m a y not be all equivalent.

1.4.2.2. Co,aplinq to strains In mul t i -component systems the coupling to secondal T order parameters , such as

the strain is more complicated than for n = l systems described in §f.4.1.2. I t is necessary to include the different strain components so tha t for BaTi03 and SrTi03, for example, there are six components of the strain. The free energy of both of these materials is then given by eqn. (I.4.33) with the addition of an elastic contribution G¢, which is given by (Devonshire 1954, Slonczewski and Thomas 1970)

G~--el(71 +72+ 73) z 2 2 2 2 2 (q, + Q2 + Q3) q- ez[~l(2Ql - qz - q3)

2 2 2 2 2 +72( Q2-Q1-Qa)+rla(2Qa-Q~-Q2)]

1 +e3(Q1Q276+Q1Qarls+Q2Qa74)+~C~l~Vlp, (I.4.37)

where we used the Voigt notat ion for the strains. Minimizing G + G~ with respect to the strains gives

el e 2 7~ - (Q~ + Q~ + Q~) (2Q~ - Q~ - Q~),

Cll -{-2C12 6~11 -C12

and

e3 74 = - ~ Q 2 Q 3 .

When these results are subst i tu ted back into eqns. (IA.33) and (I.4.37)~ the free energy has the same form as eqn. (I.4.33) for the strain-free case except tha t u and v are replaced by

4 4 g=u-- Cll +2C12 C l 1 - C 1 2

0

9 4 el ~ = v + - -

2Cl1+C12 C44"

This result shows tha t the coupling to the strain modifies the parameters of the Landau theo .ry, as found in § 1.4.1.2, and so m a y drive the transit ion to be of first order, or change the s t ructure of the distorted phase. As discussed for the n - - 1 system, it is necessary to distinguish between the high- and low-frequency susceptibilities below T c (Slonczewski and Thomas 1970). Exper imenta l confirm- ation of the necessity to distinguish between the high- and low-frequency sus- ceptibilities is provided by R a m a n scattering and neutron scattering measurements in SrTi03. The ratios of the squa reso f the ' longitudinal ' frequencies below and above T c is 2-4_+0-2 (Worlock and Olsen 1971) in excess of the factor 2 which would be expected in the absence of the coupling to strain.

As described in §I.4.1.2, the theory-can also be used to predict a jump in the

Page 56: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

56 R . A . Cowley

elastic constant a t T¢. This is quali tat ively the correct behaviour as shown by the measm'ements on SrTiO 3 illustrated in fig. I. 16. In detail, however, the behaviour is more complex as described in § II.5.2.2.

In this section the tempera ture dependence of the strain and elastic constants is found to be independent of its symmetry . This is correct only within Landau theory. In a more complete theory discussed in §§II.5.1.2 and II.5.2.2. it is essential to distinguish between strains which are secondary order parameters and break the symmet ry of the h igh-symmetry phase, and strains which have the complete symmet ry of tha t phase. In SrTiO3 an isotropic strain t/1 =t/2 = v/3 does not break any symmet ry operation of the h igh-symmetry phase, whereas a tetragonal strain, ~1 = r/2 = r/1/2 does. The la t ter strains must necessarily be zero above T¢. The former strains need not be zero above T¢ and in a bet ter theory than Landau theory the behaviour of these two types of strains is different.

1.4.2.3. Improper ferroelectrics Improper ferroelectrics are examples of systems for which the secondary order

pa ramete r is a ferroelectric distortion. Their properties are reviewed by Dvorak (1974). Since in a ferroeleetrie material there are different domains corresponding to the different directions of the polarization, there cannot be terms in the free energy of the form QQ2, where Q is the p r imary order paramete r of an n = 1 system. Imprope r ferroeleetrics can occur only when the pr imary order pa ramete r has a t least two components. The simplest model (Cochran 1971 ) of an improper ferroelectie is tha t of an n = 2 system with a coupling to the ferroelectric polarization of the form Q1Q2C2; when the free energy is

1 G=Go + ~ a ( T - Tc) (Q~ + Q~I + u(Q~ + Q~) 2

4 4 - 1 - 2 +v(QI +Q2)+eQiQ2Q+~b Q , (I.4.38 a)

where b is the inverse dielectric susceptibility which is assumed to be only weakly tempera ture dependent. Minimization of the free energy with respect to O yields

e ~) = -~Q1Q2, (I.4.38 b)

which subst i tuted back into eqn. (I.4.38) yields an equation identical to eqn. (].4.38) but with u and v replaced by

e 2 e 2 ~fe = u - and ~ = v -

4b 4b"

I f ~, < 0, the ordered structure is ease I of § 1.4.2.1, which then with eqn. (1.4.38) gives ((~) = 0 and the phase is not ferroelectrie. On the other hand if '~> 0 the ordered structure is case I I I and Q = _+ (e/2b)Qo 2 with the plus sign if ~'1 = - ~ a and the minus sign if ~1 : ~2"

In practice the phase transitions of improper ferroelectries are of first order (Dvorak 1974), possibly because the interaction 7~ is always less than u, but possibly also because of the effects discussed in § II.5.2.2. There is, however, evidence tha t ~) is proportional to Qg as shown in fig. 1.31 for Tbz (Mo04)3 (Dorner et al. 1972).

Page 57: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 57

Fig. 1.31

8 ANGLE (.S.E. {.'JMMINS) T O 158.5°C FIXED T O 155.9°C

OOo

~.,

40 80 120 160

I BRAGG INTENSITY(520)+(250) Tb2(Mo04)5 TO; 158.5°C FIXED

I TC=/54.4 tl.5°C

o ~ ~ , , , , , i 40 80 120 160

TEMPERATURE (°C)

The temperature dependence of the strain distortion 0, the spontaneous polarization and of the square of the primary order parameter as measured by the Bragg reflections in the improper ferroeleetrie Tb2(Mo04) 3. (After Dorner et al. 1972.)

Fig. 1.32

"12

"1,/

"1C

9

8

CAS

i

90

CoIB

i i I I i

95 200 230 0

T (°K)

GMO

&

I

"160 "180 T ('C)

The temperature dependence of the dielectric constant in three improper ferroeleetrics. (After Dvorak 1974.)

The development described in § 1.4.1.2 can be repeated to show tha t the dielectric susceptibility in an improper ferroelectric should, within Landau theory, have a step at To. In fig. 1.32, we show measurements of the susceptibility in three improper ferroelectries. This simple theory describes the results in (NH4)2Cd2(S04)3, but is less sa t i s fac to ry /b r CoI boracite or Gd2(Mo04)3. In par t this undoubtedly reflects the fact tha t the theory given above is very much too simple to describe real improper ferroelectrios. In particular, there is a coupling between the strain and the pr imary order paramete r and between the strain and the spontaneous polarization, as illustrated in fig. [.3 l. I t is then necessary to distinguish between the clamped and free dielectric constant as described in detail by Dvorak (1974).

Page 58: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

58 t~. A. Cowley

1.4.2.4. Fluctuations and short-rawe interactions The wave-vector dependence of the fluctuations in systems with more than one

component for the order paramete r can be obtained in much the same manner as was described in detail in § 1.4.1.3 for n = 1 component systems.

At a transit ion for which the order paramete r is neither an electrical polarization nor an elastic strain, the quadratic te rm in the expansion of.the free energy is

where

1 G2 = 2 N ~ Ru(k)Q~(k)Q:( - k),

Rij(k ) = a(T - Tc)6ij + f:~(ij)k~k~. (I.4.39)

The number of coefficients f~( i j ) may be reduced by symmet ry so that , for example, in SrTiO 3 there are three components of the order paramete r correspond- ing to rotations of the oxygen octahedra about the x, y, z directions, respectively, and three independent f~( i j ) coefficient. In this case eqn. (I.4.39) can conveniently be rewrit ten as

Rij(k ) = [a(T - To) + 2(k 2 +fk2)]~ij + 2hlcikj(1 - 3~j) (I.4.40)

and the coefficients 2 , f and h have been measured for SrTiO3 by Stirling (1972) from the anisotropy in the phonon dispersion relations about the R point, with the results

2 = 2 1 6 + 2 0 ( T H z A ) 2, f= -0"97_+0-01 , h=0"19+0-04

showing a very large degree of anisotropy. The anisotropy and coupling between the modes in these systems means tha t the

eigenvectors of the matr ix in eqn. (I.4.39), are very dependent upon the direction of k. A detailed analysis of the tbrm of the susceptibility is then difficult. An unexpected result of this behaviour occurs in the scattering cross-section around the incipient ordering wave-vector, q~, which may not exhibit the full symmet ry of the crystal. This is illustrated by the form of the scattering for wave-vector transfers K, close to 140 = 2~/a (3/2, 1/2, 1/2) in SrTiOa. At this position the structure factors of the three modes are F (K, 1) =0 , F (K, 2) = F ( K , 3) = F . Because f is close to - l the suscept ibility and scattering is large for the i th component for wave-vectors along the i th direction. Consequently there will be a large scattering obserced for wave-vectors 14 = 140 + (0, k, 0) and 14 = 14o + (0, 0, lc) but no corresponding scattering along K = 14o + (k,0,0) because F(K, 1)=0.

1.4.2.5. Ferroelectrics Tlle long-range electrostatic forces modify the form of the fluctuations for small k

in ionic materials as explained for uniaxial ferroelectrics in § 1.4.1.4. In the case of (:ubic t~rroelectrics, such as BaTi03, we choose the i th component of the order parameter to give the electrical polarization along the i th direction. The expansion of the second term in the free energy is then given by eqn. (I.4.40) but with the additional electrostatic term;

kikj Rij(k)=[a(T-T~)+2(k2+fk{)]b~j+)~hklkj(l--(~ij)-]-g ~ . (1.4.41)

The matr ix Rij(k ) can then be diagonalized for k o 0 to give the 'fi'equencies' of a

Page 59: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 59

longitudinM f luctuat ion above T¢ as co~ = a(T-To)+g and two t ransverse fluctu- at ion . . . . t requencms as COT 2 = a(T--T~). I f the s t ruc ture factors of the three modes are propor t ional to K i, eqn. (I.2.36), then the scat ter ing intensi ty for large g is given (Cochran 1969) by [K21 sin 2 0 where 0 is the angle between N and k as i l lustrated in

fig. 1.33.

Fig. 1.33

Sketch of constant intensity contour lines as function of wave-vector transfer 14 for 0, cubic ferroelectric or for K within the plane of ferroelectric fluctuations for a [2d] ferroelectric. (After Cochran 1969.)

I n the case of ferroelectrics with two componen t s polarized in say the x and y directions, there are two t ransverse f luctuat ions for k along the z direction, bu t a longitudinal and a t ransverse f luc tua t ion for k in the (x, y) plane. Fo r K in the (x, y) plane the scat ter ing has the same form as shown in fig. 1.33. Al though the te t ragonal phase of BaTi03 is a two-componen t ferroeleetric system, we are unaware of any detailed measurements of the critical scat tering.

1.4.2.6. Acoustic instabilities Near ly all the acoustic instabilities with n ~> 2 have cubic terms in the L a n d a u

expansion of the free energy, and so are predicted to be of first order. The only example (Cowley 1976 a) wi th n = 2 and no cubic terms is an instabi l i ty agains t r/4 and t/5 in the te t ragonM system. The velocities of the acoust ic waves are zero at T¢ for modes with wave-vec to r k perpendicular to the [001] direction and polarized along (001) or with k along [001] and polarized transversely. The former set give rise to a similar scat ter ing pa t t e rn to t h a t of un ia l i a l ferroelectrics, while t h a t of the la t ter set is similar to t h a t of the acoust ic instabilities discussed in § 1.4.1.6.

1.4.3. Phase diagrams

1.4.3.l. Tricritieal points I n §I.4.1.1 the free energy, eqn. (I.4.9), was shown to give rise to a cont inuous

phase t rans i t ion if v ,>0, and a first-order t rans i t ion if u < 0 . The point u = 0 terminates the line ( in the space of u and T) of cont inuous phase t ransi t ions and is

Page 60: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

60 R . A . Cowley

known as a tricritical point. The Landau expansion of the free energy a t this point is

= G O + ~a (T - T~)Q 2 + hQ 6 , (I.4.42) G

so tha t the order paramete r is given by

Q4 = 6~(T c _ T) (I.4.43 Cb)

and the susceptibility below T¢ by

)/(0) = [4a(T¢ - T)] - 1. (I.4.43 b)

The results show tha t within Landau theory the principal tricritical point exponents have values fl = 1/4, ?-- 1 (to be compared with the Landau resfflts for an ordinary critical point, f l = l / 2 , ?=1) . I t also follows from (I.4.43b) tha t the ratio of the susceptibility above T¢ to tha t below T¢ for a constant I T - T e l is 4 within the classical tricritical theory- - twice the Landau result for an ordinary critical point. A fuller description of tricritical points and a comparison with experimental results is given in § ]I.5.3.1.

1.4.3.2. Phase transitions without symmetry brealcing In § 1.3.1 we emphasized the importance of the concept of symmet ry breaking at

structural phase transitions. There are, however, phase transitions where (as a t the l iquid-gas critical point) the symmet ry is unchanged. Consider, to be specific, the Landau free energy for an n = 1 system in an applied field:

G = G O + 2 a ( T - T o)Q2+ uQ4 + hQ6 _ EQ

For u > 0 there is no phase transit ion when E ¢ 0. For u < 0 there is a first-order phase transit ion for small E, and no phase transit ion for large E. There exists a critical value of the field Ec a t which the transit ion is continuous and beyond which no transit ion occurs. The critical field, and the corresponding transit ion tempera ture To are located by the conditions

=o,

aS

and

f l u l \ Si 2

6U 2 Tc = T o + -

5ha

There is no symmet ry breaking a t the critical point, since (by vir tue of the applied field) the coordinate O already has the non-zero value

Page 61: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 61

The order pa ramete r for the phase transition must thus be defined as the difference between the two possible values of Q0 characterizing the two phases tha t coexist on the phase boundary which terminates a t the critical point.

This type of behaviour has been observed in K H 2 P 0 4 by Eberhard and Horn (1975). At atmospheric pressure the phase transit ion is first order and shows a temperature hysteresis of 0"025 K. On increasing the applied field the hysteresis decreases and extrapolat ion of their da ta suggests tha t Eo = 6"5 x 105 V/re.

1.4.3.3. Competing interactions One of the characteristic features of an n > 1 component system is the possible

occurrence of different ordered phases: the n- - 2 system has an ordered phase, I, with ~1=1, ~2=0 and another, I I , with ~1=~2=1/x/2 whose relative stabil i ty is determined by the sign of v, § 1.4.2.1. More than one phase transition m a y occur if there are competing interactions one of which favours one of these phases, while another favours the other phase. This type of competi t ion is responsible for the successive phase transit ions in BaTi03 (Devonshire 1954) and we illustrate the behaviour by extending the free-energy equation, (I.4.33), to include sixth-order terms in the order parameters;

1 G = Go +~a(T- To)(Q~ + Q~ + Q~) + u(Q~ + Q~ + Q~):

+v(Q~ +Q~ +Q~)+hI(QZl +Q~ +Q~) 3

+ h2(Q~ + Q~ + Q~) +h3[Q~(Q~ + Q~) .~ 4 2_}_ 2 4 2 2 02(01 (I.4.44) Q3)+Q3(Q1 + Q:)].

The quadratic te rm and the terms in u and h 1 are isotropic and so do not determine the nature of the ordered phase. We shall determine the structure of the ordered phase by using per turbat ion theory to calculate the energy of the anisotropic terms. Consider three possible structures for which (Q1,Q2,Q3) are given by: I(1,0, 0)Q0, II(1/~/2, 1/~/2, O)Qo and I I I (1/~/3, 1/~/3, 1/x/3)Q o. The contribution of the anisotropic terms to the free energy is

GAs(I) 2 ¢ = (v + haQo)Qo,

I f v < 0 then phase I will tend to be more stable than phase I I or I I I . If, however, h2 and h 3 are negative then they will tend to stabilize phase I I I ra ther than phases I or I I . Since, however, the h-coefficients are multiplied by Q~, there may be a tempera ture at which it is energetically favourable for a transition to occur from the structure favoured by v to tha t favoured by the h-coefficients. I f we neglect h 3 for simplicity then three transitions are predicted by eqns. (I.4.45) if v < 0 and h 2 > 0: d i sordered~phase I when T = T~; phase I ~ p h a s e I I when Qo 2 = -2v/3h2; phase I I --+phase I I I when Q~ = - 6v/5h> Whereas the first transition may be continuous, the second and third transitions are necessarily of first order. I t is interesting to note tha t phase I I can never be stable under the action of the parameters v or h2 alone, but can

Page 62: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

62 R . A . Cowley

be stabilized by a competition between these terms. Devonshire (1954) applied this theo~.~ in detail to BaTi03 and was able to show that it explained the occurrence of the three transitions. However, not surprisingly, a satisfactory quanti tat ive description could be obtained only when the parameters were allowed to be temperature dependent.

I f the parameter v is positive, phase I I I is stable below T c and, if h z is negative there is a first-order transition to phase I, when Q~) = - 3v/4h. In this case phase I I is never stable. In fig. 1.34 the phase diagrams of these n = 3 systems are illustrated.

Fig. 1.34

UNDISTORTED

~ T

V>O

h 2

U UNDISTORTED

T V(O

Phase diagrams for an n = 3 system with Q4 and Q6 anisotropies.

Another type of competition between different phases arises if a stress or strain is applied to the crystal favouring one of the possible lower temperature phases, rather than the other (Bruce and Aharony 1975).

The effect of a [100] stress on SrTiO3 is to split the degeneracy of the three modes 01, Q2 and Q3. The free-energy equation, (I.4.33), may then be rewritten as

1 a = G o + ~ [ a ( T - T¢) (Q~ + Q~ + Q~) + 2plQ~

+p2(Q~ + Q~)] + u(Q~ 4- Q~ 4- Q~):

4 4 4 4-v(Q~ +Q2 4-Q3), (I.4.46)

where the new terms Pl and p2 are proportional to the stress applied along the [100] direction. In the interests of simplicity we shall assume that Pl -= - P 2 - - P 0 , neglect the strain dependence o f u and v and assume v is negative (e.g. SrTi03). in this case both the applied stress and v favour ordering along the cube axis, Qt, if p0 < 0, and Q2

Page 63: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 63

or Q3 if p0 > 0. The phase transitions occur when

2po T = T ¢ - - - tbr po<O

a

T=T¢+ p° for p o > 0 , a

and there is a first-order transition separating the two ordered phases, as shown in

fig. 1.35 (a).

Po

Fig. 1.35

ordering along / [olo] or [ o o /

[loo]

Po ¸ T. ordering / along [ 0 1 1 ] /

Z . ~-T

(a) (b)

Phase diagrams for a [ 100] stressed n = 3 system. (After Bruce ~nd Ah~rony 1975.) (a) The case v<0 (e.g. SrTi03). (b) The case v>0 (e.g. LaAI03).

The behavour is more complicated if v > 0, (as in the case of LaAIO3), because then the applied field favours ordering along the cube axes while v favours the [1, 1,1 ] type of ordering. The essential form of the phase diagram may be determined by assuming that the anisotropic terms are small and using perturbation theory to calculate their effects: a more complete treatment is given by Bruce and Aharony (1975). I f the order parameter of the ordered phase is written (Q1,Q2,Q3)

(cos 0, {2 sin 0, ~3 sin O)Qo where ~2 ± y2 1, then the anisotropic contribution to = S 2 t ~ 3 ~

the free energy is

3po 2 GAN = T Qo cos2 0 + v(cos" 0 + sin 4 0(~ + ~))Qo 4 .

~2 + ~3 = 1 , and hence we can minimize Now since v > 0 and we wish to minimize GAN, 4 -4 GAN with respect to the angle 0 to obtain three solutions:

I. cos 0 =0,

1 Po II . COS 2 0

3 2vQ~

I I I . sin 0 = 0

Phase I is stable when Po > 2vQ2/3 and phase I l l stable when Po < -4vQ2/3 and an intermediate phase in which the ordering vector is more general is stable tbr intermediate values of Po- The phase diagram is sketched in fig. 1.35 (b).

Page 64: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

64 R . A . Cowley

A similar analysis may be applied to the ease of [111]-stressed materials, though the t r ea tment is complicated by the non-orthogonali ty of the different [111] directions. This time, as one might expect, an ' intermediate ' ~phase occurs when v is negative, as shown schematically in Fig. 1.36. A comparison of these results with experimental da ta is given in § II.5.3.3.

Fig. 1.36

phase

~T

(e) (b)

Phase diagrams for a [1111 stressed n= 3 system. (After Bruce and Aharony 1975.) (a) The ease v<0. (b) The ease v>0.

1.4.4. Incommensurate phase transitions

1.4.4.1. The simplest incommensurate phase As ye t we have described only those phase transit ions where the order

parameters , Q~, are real, but in general they may be complex. Consider a te tragonal material in which the order parameter , Qi, has a wave-vector q~ = (0, 0, ~). Then unless the wave-vector is the zone centre or zone bounda~.~, the order pa ramete r Qi will be complex. Init ial ly it seems surprising tha t qs can have a value between 0 and the zone boundary because it is only at these points that symmetry may impose that Vqc0(qj) = 0. There is nothing, however, to prevent Vqc0(qj) = 0 for an arb i t ra ry point in the zone for some part icular combination of interatomie forces. However, in these cases the wave-vector qs of the minimum in the dispersion relation is not fixed by symmet ry and so is usually tempera ture dependent. Both of these features are i l lustrated in figs. 1.37 and 1.38 for K2SeO 4 (Iizumi et al. 1977), where we show the measured dispersion relation and the tempe~bature dependence of q,.

In the theory of incommensurate phase transitions the order pa ramete r Q1 is complex and there is necessarily another coupled order parameter , Q2, with wave- vector -q~. Since, however, the atomic displacements, a(l~) are real, eqn. (I.2.1) shows tha t Qx and 02 mus t be complex conjugates of one another, and it is frequently useful to define new variables to explicitly take aceoung of this condition. The most direct way of doing this is to write

1 1 Q1 = ~ A exp(i~b), Q2 =x/2 A exp(-iq~), (I.4.47)

where the coordinate A is the ampli tude of the distorted wave and qb is its phase and

Page 65: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 65

Fig. 1.37

5.0 Z.B.

4.0 o ' X2

~.o

2.0

, o

I o . . . . t , q ,

0 O.5 1.0 q ( REDUCED UNIT)

The dist)ersion relation of the soft, mode in K2SeO 4 at various temperatures. (After Iizumi et al. 1977.)

Fig. 1.38

I I I I I I

0.12 -- t _ oo3~, -~

• ~ o ~ 0 - o Ti

_ SUPERLATTICE Ii T CRITICAL "5 0.0~ _OO2,~_t REF - o o , ; SCATTE"'.O

0.04 3.Oh~-'

i 0 O2 ~ ~""

L 0 L{I t I I I

so 9 0 Ioo no 120 ,30 ,4-0 TEMPERATURE (K)

1 2~ The temperature dependence of qs=.~(l--cS)~ in K2SeO 4 compared with calculation by

Iizumi et al. (1977).

both are real. An a l te rna t ive approach is to in t roduce new coordinates which are wr i t ten as

1 P1 = / 2 (Ol + O2),

i P2 = ~ 2 (Q2 - Q 1 ) , (I.4.48 a)

where P1 and P2 are again real. The relat ionship between these two coordinate sys tems is

P l = A c o s q ~ while P 2 = - A s i n ~ b . (I.4.48 b)

Page 66: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

66 R . A . Cowley

The free energy of this system is

G = G O + a ( T - T¢)Q1Q2 + 4uQ~Q~, (I.4.49)

where the terms in Q~, Q~, Q~, and Q~ are absent because they fail to satisfy wave- vector conservation. Substituting eqn. (I.4.47) into eqn. (I.4:48) we obtain;

= G o + I ( T - T¢)A 2 + uA4; (1.4.50) G

an expression involving the amplitude A which is identical to the free energy of the singlet system discussed initially, eqn. (I.4.3). We immediately obtain, therefore, that

a A ~ - - ~ u ( T ¢ - T ), T<T¢; (I.4.51)

and analogous expressions for the specific heat .and free energy. The other noteworthy point about eqn. (1.4.50) is tha t it does not explicitly involve the phase angle ¢. The free energy is therefore independent of the choice of ¢.

In the alternative representation with the order parameters, P 1 and-P2, the free- energy equation, (I.4.49), becomes;

G = Go + l a ( T - T¢)(P~ + P~)+,a(P) + p~)2 (1.4.52)

This form of the free energy is identical with that of the n = 2 system, discussed in § 1.4.2.1, except tha t v = 0. The phase transition is therefore to a phase of type 1I, § 1.4.2.1, corresponding to (P 1, P2) = A0 (~ 1, ~2) with the eigenvector ~ arbitrar:7. This continuously v~riable direction in the (Px,P2) space corresponds to the arbi t rary choice of the phase angle q5 in the other representation.

The analysis of § 1.4.2.1 can now be repeated to obtain the susceptibility both above and below the phase transition to give

%L(O)=%T(O)=[a(T- - Tc)] -1 , T>T¢; (I.4.53 a)

%L(0) = [2a(T¢ -- T) ] - ~, T < T¢; (I.4.53 b)

)/T(0) = o0, T < T ¢ , (I.4.53 c)

where %5 is determined by the fluctuations parMlel to the ordering vector, ~, which am frequently called amplitude fluctuations, and XT corresponds to fluctuations perpendicular to tha t direction which have an infinite susceptibility. These fluctuations were called phasons by Overhauser (1971) because they correspond to changes in the phase of the order parameter.

1.4.4.2. More complex incommensurate phases Not all incommensurate phases have only two components of the order

parameter. In order to illustrate the type of behaviour which occurs we examine the properties of two n = 4 systems. The first corresponds to an incommensurate phase in an orthorhombic crystal with wave-vector q~ = (+ ~1, -+ ~2,0) or a tetragonal crystal with q~ = ( + ~, 0, 0) and (0. + ~10). I t is then convenient to introduce two amplitudes

Page 67: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 67

and phases by the requirement that

1 Q1 =/~-A1 exp (i~bl), 02 =Q*,

(I.4.54)

1 Q3 = / ~ A 2 exp (iq52), Q4 =Q*.

The free energy is then of the form

G = Go + a ( T - To) (QIQ2 + Q3Q~) + 4(u + v) (Q~Q~ + Q~Q~)

+ 8uQ1Q2Q3Q4. (1.4.55)

With the aid of the transformation, eqn. (I.4.54), the free energy becomes

1 G=Go+~a(T-Tc)(A~+A~)+u(A~+A~)2+v(A~+A~), (I.4.56)

which is independent of the phase angles q51 and ~b 2. Now, since eqn. (I.4.56) is identical with the n =2 system described in § 1.4.2.1, there are three solutions

I. v < 0 A 1 or A2=Ao, and the other is equal to zero.

II. v = 0 A~+A~=A~.

III . v>O AI=A2=Ao/x /2 .

The susceptibilities can be obtained in the usual way. For structure I, with say A 1 = Ao, there is a longitudinal susceptibility corresponding to fluctuations in A 1 given by

zI(O)=[2a(T~-T)] - t , T<T~, (1.4.57)

while the transverse susceptibility arising from the fluctuations in the phase q51 is infinite. The susceptibilities of the Q3 and Q4 modes are degenerate and given by

u-t-v Z3(0)=Z4(0)- av(Tc_T), T<T°.

If the ordered phase is of type I I I then there is a longitudinal susceptibility given by eqn. (I.4.57), and another amplitude susceptibility corresponding to anti-phase fluctuations of the two amplitudes with a susceptibility given by

2u+v )~n (0) = 2va(T~- T) ' T < To.

and two infinite susceptibilities, corresponding to fluctuations in the phase angles q~ 1 and ~b:. In this system the number of phasons with infinite susceptibility depends on the number of pairs of q~ and - q~ present in the ordered structure. I f there is only one pair, phase I, then there is only one infinite phase susceptibility, but if both modes arc present in the ordered structure, two susceptibilities are infinite.

Finally let us consider another case of an n = 4 system; BaMnF4 which has an 'a' face centred orthorhombic structure and an incommensurate phase described by the four order parameters;

1 1 Ql(q~= (0"39,~,g)), Q2(q~= (-0"39,½,½)),

-g ,g ) ) and Q4(q~= Q3(q~=(0"39, 1 1 (-0'39,½,½)), (Shapiro et al., 1976).

Page 68: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

68 R . A . Cowley

The free energy is now of the same form as eqn. (I.4.55) but with an additional term given by 4w(Q~Q~ + 2 2 Q,,Q2). When this is t ransformed to the form of the free energy given in eqn. (I.4.56), we obtain an additional term wA~A~ cos (2gb I + 2q52). This te rm has the effect of fixing the relative phases between the two pairs of incommensurate wave-vectors. Consequently for the s tructure of type I, discussed above, this term splits the degeneracy between )/3(0) and )~4(0), while for the s tructure of type I I I , there is only one infinite susceptibility which arises from fluctuations which preserve ~b i ÷ (~2~ while the other phase susceptibility is finite and corresponds to fluctuations in 4t + q52. The effect of the Umklapp interaction in BaMnF 4 is therefore to reduce the degeneracy of the susceptibilities, and in part icular to ensure tha t there is only one phase mode with an infinite susceptibility. The number of infinite phase susceptibilities is then the number of pairs of wave-vectors in the ordered structure, n/2, less the number of phase locking terms. There are always phase locking terms which restrict the number of infinite phase susceptibilities to be a max imum of three.

1.4.4.3. Fluctuations of incommensurate phases We now extend the analysis, part icularly of the simplest incommensurate phase,

to include the wave-vector dependence of the fluctuations. The free-energy equation, (I.4.49), is extended to include the wave-vector dependence of the quadratic te rm in Q1Q2 when ibr smM1 k expanding about the tempera ture dependent, q~, gives

G 1 G = o ÷ ~ ~ (a(T - T¢) ÷f l (]d2x ÷ ]~2) ÷f2(k~))Q~(k)Q2( _ k)

4u + ~ ~ Q~(k~)Q2(k3)Q~(k2)Q2(-kl-k2-k3). (I.4.58)

~* k l k2 k3

The fluctuations can be found by transforming to the P 1 (k), P2 (k) coordinates when for T > To the quadrat ic terms in the P(k) become

1 2N ~ ~=1"= Ru(h)Pi(k)Pi( - N)

with

Rii(k ) = a ( T - T¢) + fl (k~ + k~) + f2k~, (1.4.59)

so tha t the Trequencies' of the fluctuations described by P l (k) and P2(k) are equal and given by

~o i (k) 2 = c%(k) 2 = R11 (k). (I.4.60)

The corresponding frequencies below T c can be found by assuming a part icular phase for the ordered structure so t ha t qb = 0 in eqn. (I,4.48 b), when the longitudinal or ampli tude frequency is given by

O)L(k) 2'= 2a(T¢ - T) +fx ( k2 + k~) ~ - f 2 ] c 2 (I.4.61)

and the transverse or phase frequency by

~T(k) 2 =2'1 (k~ + k, ~) +f2 (~) (1.4.62)

In fig. 1.39 we illustrate these frequencies close to T¢. The phase frequencies are lineal' with k below T¢ and their slope is independent of

t empera ture in this model.

Page 69: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions

Fig. 1.39

DISPERSION RELATIONS FOR AN INCOMMENSURATE

PHASE IN THE PLANE WAVE REGIME

69

b \ . . ../ \ t

\ / --,. AMPLITUDE / ,~, / I / ~ \ \ MODE i I /

PHASON

-%s %s WAVEVECTOR o b

The dispersion relations for the excitation frequencies of an incommensurate system with a plane-wave ground state. (After Bruce and Cowley 1978.)

Before discussing experimental observations of phasons, it is worthwhile discussing their nature in more detail. An alternative approach for describing the fluctuations of an incommensurate phase is to consider fluctuations in the amplitude and phase. By analogy with the t rea tment of the fluctuations in the order parameter Q as wave-vector dependent Q(k), we introduce a wave-vector-dependent fluctu- ation in the amplitude, A(k), and phase ~b(k) so that from eqn. (I.4.47) we obtain for small fluctuations

1 Qz(k) = ~ ( A ( k ) exp (i~bo) + iA o exp (i¢o)qS(k)) (I.4.63)

where 4)o is the phase of the ordered structure. I f the ordered structure is chosen as above to have ~b 0 =0 , then to leading order in the fluctuations

Pl(k) =A(k) , P2(k) =AoqS(k). (I.4.64)

P 1 (k) and P2 (k) can now be identified with the amplitude and phase fluctuations, but the identification is accurate only for leading-order terms in the expansion of the ('omplex exponential, exp (i¢(k)); once ~b(k) becomes sufficiently large to invalidate this expansion, then the two developments are no longer identical. In particular, the second-order terms in the expansion of exp (i~b(k)) give a real contribution to the right-hand side ofeqn. (I.4.63) which must be incorporated with the P1 (k). This leads to a large correction to the longitudinal susceptibility: in fact we have shown (Bruce and Cowley 1978) that it varies as 1/te tbr small k below 7'¢, cf. eqn. (I.4.53b). This is an exactly analogous problem to the well-known divergence of the longitudinal susceptibility of a Heisenberg ferromagnet below T¢ on the coexistence curve (Vaks et al. 1967). In our notat ion it arises because fluctuations in the direction of the order parameter in (P1,P2) space are relatively easy and to second order give rise to fluctuations in the length of any component of tha t parameter.

Page 70: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

70 R . A . Cowley

The scattering cross-section for X-ray or neutron scattering is readily obtained using the techniques developed in § 1.2.3 and the definition of the P coordinates in eqn. (I.4.48). The result is

h N S~(K, co) = ~ ( n ( c o ) + l) ~ [A(K + qs + k)

k

x IF(K, q~ + k)l 2 + A ( K - q~ + k) iF(K, - q~ + k)l:] x Im [XL(k, co) -I- XT(k, co)], (i.4.65)

where XL(k, CO) and XT(k,CO) are the frequency dependent susceptibilities of the longitudinal and transverse fluctuations, and the Fs are the mode structure factors defined by eqn. (I.2.36). Above T~, eqn. (I.4.65) reduces to the usual one-phonon cross-section from the soft mode, but below Tz,XL and ZT differ so tha t there are two response functions for each K. The scattering cross-section in principle enables both the phase and amplitude modes to be observed as illustrated in fig. 1.39. Alas we are unaware of any neutron scattering measurements which convincingly vindicate the existence ofphasons in real incommensurate systems. In KeSeO 4 (Iizumi et al. 1977) and BaMnF4 (Shapiro et al. 1976) considerable scattering is observed for the appropriate wave-vector transfers, but it has not, as yet, been possible to distinguish the scattering from the phase modes from that of the normal elastic waves seen emanating from the Bragg reflection of the incommensurate phase. In fig. 1.40 we show experimentM results obtained for the one-dimensionM conductor, KCP, (Carneiro et al. 1976) and this too definitely shows scattering at low energies close to the incommensurate wave-vector, 2kF, but it is not clearly of phason character. In TaSe 2 (Moncton et al. 1975) and NaNO2 (Sakurai et al. 1970) the transitions are more closely of the order~disorder type and so it is not possible with conventional neutron scattering techniques to determine the dispersion relations.

The amplitude and phase modes can also be observed using light scattering techniques. The polarizability tensor is expanded in powers of the coordinates Q1 and Q2 as

Pap(K) = ~ P~(qs + k, - q s - k + K)Q l(k)Q2( - k + K). k

Now since the modes can be determined most directly by one-phonon scattering we are interested in the terms in which either Q1 (k) oz' Q2(k) takes on its average value. I t is useful to expand the polarizability tbr small K so tha t

- =P~a + ~ Pap, ~Kv + •. -

and then since P~(R1, k2) is symmetric in k 1 and k 2

1 P~p(q, + K, - q , ) = P = ~ - ~ P = , , ~ , I ¢ , + . . . .

Wl~en the polarizability is expressed in terms of the P coordinates and the ordered structure is assumed to have qS=0 (eqn. 1.4.48 b) then

0 " 1 Pap(K) = AoP~pP I (K ) + zAo ~ P~ .~K~P E(K ). 7

Page 71: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

S t r u c t u r a l p h a s e t r a n s i t i o n s 71

r g

7

8

5

i ~4 ¢;

3

2

7

6

5

5

2

I

0

F 8

• . , - ~ - - - ' , - - ,

• Y N ¢ ' . - ' - - ' , : . " : - ~ m ] ~ .' . . ; , ~ / . - ' __._; -.: ) / / . . .

• " , , '1 / . ; " . . ' . , ' ~ " Yl'~ ." • 1 " ~ l : :;,,:

• . ~ / ~ . . ~..-... ,' : • / / " / ' - ~ •

• " , , ;V . ; ' ~ /,':'/i/." : : • ~ 7 . ; ' / k / # . " : . - - , '

• "z," ; , , / f , . l lh . . I "\ "

• ; ; / / . , / . _ /;:11i: : , ' "~: . / y / . : . L. . ' iO 9 I.;:/H: ".I ~ . ~"

• ,5~. ' , '> " ; I ' : lhb - ' H ~ ! ,

" , , / . ; ' , , " ,' . " / I - H I: - -

i

OI

i

i

O2

Fig. 1.40

h i . i

. . . . 5 . . . . . "K2 Pl(CNI,IBfo 3 3 2 D2Q .' T : B O K

I .

RESOLUTION(FWHM )

p o 4 o ~

INTENSIT IES- o5!} t 5 2O 30 4 0 OOUNT~ 5o ~

I00 2 0 0 3OO 4 0 0 50O

../;',<.->~.X~ {.

0 3 014 l : [ C ' ]

/i , . . / / . '

o ° / / . "

0.5

• / / . " . , , 5 . . . . . " / / . "

. ~ - Z , ' " i f . . / o • ~ . . . . . . , . . . . . • " > ' ( , ~ . ' : ~ , . ~ , . ' \ - ~ " ! l" 2 ' " ' . % 3 ° ' % ~

. • / / / , ~ . . j / ~ . _ ~ " ~ . " . . . . ~ j - ~ I i : T= 160 K ° L i ~ •

~ ! ! 1 //// 'oo¢o' ~ ' ~iI. Ill-

~ ~ / l i -1 iV" )Ii: 30 .. . . . . . I

!iiE. ~.lt ~oo I , iff, ::11 ooo ....... j

OI 0.2 0 3 014 O ,5 c [c l

Neutron scattering intensity counters for scattering vectors K=~C*+~ in KCP. (From Carneiro et al . 1976.)

This equation shows tha t the amplitude fluctuations are observable with l~aman scattering techniques, while the phase fluctuations can be observed by Brillouin scattering techniques• The frequencies and line widths of the amplitude mode in K2SeO¢, measured by P~aman scattering techniques (Wada et a l . 1977) are shown in

Page 72: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

72 R . A . Cowley

fig. 1.41. Since the phase modes can be observed in Brillouin scattering they mus t have the same symmet ry as acoustic modes, and will therefore interact with them. This interaction can be evaluated (Bruce and Cowley 1978) by expanding the coupling coefficient between the strain ~/~ and the modes (q j) and ( - q j),

in eqn. (I.2.10), as was described above for the polarizability. The result is similar; there is a linear coupling between the phasons and the strain parameters which is determined by

i A o V~p.~ - Ok~ V~p . .

This coupling is proport ional to the order pa ramete r Ao, and will modify the velocities of both the phasons and acoustic modes.

Fig. 1.41

.r (;9

~20 Z bJ

OI bJ n,. h

E3

l--

Z o

W $ • . ° o °

j + o j o ~o o. °'g°'°'° o

I. , -- n 8 0 1130 120

TEMPERATURE (K)

K2Se04

lI i

140

Raman scattering measurements of the longitudinal or amplitude mode in K2Se04. Phase (2) is paraeleetrie, (3) is incommensurate and (4) is a locked-in phase. (After Wada et al. 1977.)

Alas we do not, as yet, know of any direct observat ion of phasons by either Brillouin or ultrasonic techniques. Bechtle and Scott (1977) have suggested tha t some anomalies in the intensity of the Brillouin scattering of the acoustic modes in BaMnF4 arise because of coupling to the phasons but this is not ye t proven.

Finally, we discuss the effect of the phasons on the Debye-Wal le r factor of the incommensurate phase. The Debye-Wal le r factor is determined by the displacement-displacement correlation function for the part icular site in the crystal. When wave-vector is conserved this can be expressed in terms of the normal mode

Page 73: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 73

correlation function ~ <Q(qj)Q(-qj '} as shown by eqn. (I.3.36). In an incom q j j"

mensurate phase, this correlation function is given by

1 <%(l~)ua(l~)}=M~ N ~ e~(~,q~jl)

q l q x J l J 2

x %(K,q~2)<Q(qjjl)Q(q2j2) } expi(ql -q2) . R(K). (I.4.66)

We shall now consider the part of the Dehye-WMler factor dependent upon the correlation function of the amplitude and phase modes described above. When the transformation is made to the P1 and P : coordinates the relevant contribution to (I.4.66) is proportional to

~5/~ [(1 -cos (2q~. (1 (2q~. R(/~)))/~OT(k)2 + R(bc)))/Coz(k) 2] +cos (I.4.67)

where the frequencies are given by eqns. (I.4.60)-(1.4.62). This result is unusuM below T¢ in that it varies with position, R(l~). In an incommensurate phase each atom is displaced a different amount and since the fluctuations correspond to fluctuations in that displacement it is to he expected that the Debye-WMler factor varies with position. The intensity of the Bragg reflections corresponding to the underlying lattice K = z and to the incommensurate phase, 14 = z + qs, have the same Debye-Waller factor determined by

1 1

An alternative approach to the seattel'ing properties and the Debye-WMler factor was developed by Overhauser (1971) using the phase variable ~b(k) directly. The component of the Bragg scattering from the (/~)th atom is determined by (cxp iK. u(/lc)}exp iK. R(/~c). If we consider only the displacements due to the phase fluctuations then

/ 2 V/2 u(/~) = [ ~ 2 Aoe(~,q~))eos(q,.R(l~c)+4)o+4,(l~c)), (I.4.69)

where ~b(l~c) is the fluctuating part of the phase

1 ~b(1K) = ~ q~(k) exp (ik. R(/K)). 0.4.70)

The Bessel function expansion for a complex exponential is

exp(izcosO)= ~ i'J,(Z)exp(in~h), n = - - c o

and gives

<exp (iN. u(/~c))} = ~ i"d,(y) exp (in(q~. R(ll¢) + ~b0))) t l = - c o

x <exp (in4(lrc))>, (I.4.71)

where

y=x/2AoK, e(~, q~j)/M~/2 .

Page 74: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

74 R . A . Cowley

The effective Debye-Wal le r factor for the different terms d, = {exp (inq~(ltc))} is then given, using eqns. (I.4.70) and (I.4.64), as

n 2 2) This result shows tha t the Bragg reflections of the underlying lattice, K ='c, are

given by the terms in eqn. (I.4.71) with n =0 . Since d o = 1, the phase fluctuations do not contr ibute to their Debye Waller factor, while the p r imary Bragg reflections of the incommensurate phase K = • - qs have a Debye Waller factor given by d 1. This is an unusual Debye-Wal le r ti~etor because it is independent of K and deereases to zero rapidly as T~T~ due to the dependence on Ao 2. The higher terms in the expansio~ (I.4.71) correspond to the two-phonon, three-phonon terms, etc., in the expansion of the seattei'ing cross-section as discussed in § 1.3.1.

These results are dear ly very different from those derived using the expansion in terms o f P 1 and P2, eqn. (I.4.68), and so we are left with two questions. First ly why are they so different, and secondly which, if either, is eorreet? The reason for the difference is tha t the expansion in P2 and the phase fluctuations, q~, are only exactly equivalent to first order. Since the Debye-Wal le r factor is a summat ion over infinitely many-phonon proeesses the two expansions give different results because they include terms beyond the first order in a different way. I t is clearly physically appealing to use the phase fluctuation model of Overhauser (1971) because the energy is unchanged by a change in the phase, q~0. Consequently it would seem tha t eqn. (I.4.72) is to be preferred. Close to T~, however, the fluctuating displacement, ~b(l~c), can become large, because the ampli tude A0 becomes very small. Our development, however, has t reated the fluctuations in ~b(l~c) as a harmonie oscillator and this is clearly inappropriate when ~b(l~c)~ re. Consequently we expect the phase mode contr ibution to the Debye-Wal le r factor to be given accurately by eqn. (I.4.72) for T<< T~, but close to To there will be deviations from this behaviour and it will possibly revert more closely to the form obtained in eqn. (I.4.68), which yields the normal Debye-Wal le r factor for T > T~.

1.4.4.4. Secondary order parameters At incommensurate phase transitions there is always a coupling between the

strain and pairs of order parameters . In the simplest model the coupling is given by 2eO, lQetl, which when t lunsformed into the phase and ampli tude coordinates becomes eA211. Consequently the coupling alters the magni tude of the ampli tude and the character of the ampli tude fluctuations but has no direct effect on the phase fluctuations. These effects can be t reated using the methods outlined in § 1.4.1.2, and the results are essentially the same as those obtained there.

In principle it is possible to have an improper ferroelectric in which the pr imary order parameter has an incommensurate wave vector, if the irreducible represen- ta t ion of the order paramete r is doubly degenerate. In this case the ferroeleetric ordering will be accompanied by a part icular choice of the relative phases of the two systems (Cochran 1971 ). Since, however, we are unaware of any real systems of this type we shall not discuss them in more detail.

Of more interest in incommensurate phases are secondary order parameters associated with wave-vectors of the form

qm = mqs + vm, (I.4.73)

Page 75: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 75

where m is any integer. A simple mean field a rgument shows that if there is a te rm in the hamil tonian of the form

then in an incommensurate phase with a distortion of wave-vector qs, there will also be a distortion with wave-vector 2qs. Extending this argument shows tha t distortions will occur ibr all the wave-vectors given by eqn. (I.4.73). We may expect, however, tha t increasing j will involve increasing orders of the anharmonic interactions so tha t the ampli tudes of these distortions decrease rapidly as j increases. Despite this it is necessa W to consider if these secondary distortions can influence the quali tat ive features discussed above.

The free energy of the system within a completely general Landau theory m a y be writ ten (Kwok and Miller 1966)

G = G o + ~ ~ V. <Q(q~ l )> . . . <Q(q~.)}A(ql + .... +q . ) n q . . . . qn \ ~ 1 Jn

J1 • . . j r l

where the qm helong to the set given in eqn. (I.4.73) and the V. coefficients may be tempera ture dependent. Now consider a change in the phase of the distorted w a v e s t o

(Q(qm.)'m)}¢ = exp (im¢)(Q(qmj,,)} , (I.4.74)

then the free energy of the new phase is

(q, q. ox, o , = G o + X F, v. (i(ml+ .... +m.)~) . q. .... s . \ 3 ~ 3 . 1

J1 . • . J n

x {Q(qlJO}. - . {Q(q, j , )}A(ql ÷ . . . +q,) .

But if q~ is t ruly incommensurate , wave-vector conservation implies tha t ml + m2 + . . . + m, = 0. Consequently G 4 = G and the free energy of the incommensurate s tate is unchanged by the t ransformat ion (I.4.74). This then suggests tha t the phason behaviour, namely XT(k)~oo as k+O, will occur even when the effects of these secondary order parameters are included. An explicit example has been detailed by us elsewhere (Bruce and Cowley 1978).

These secondary order parameters break the same symmet ry operations of the high-temperature phase as the pr imary wave-vector, q~. They are therefore secondary order parameters of the second type discussed in § 1.3.2. From a somewhat pedantic point of view, eqn. (I.4.73) defines a set of ~ N 1/3 possible components of the order pa ramete r with different values ofqm, in the same direction, corresponding to different components of the same set of order parameters. A different type of combination of this set of order parameters is employed in the next section to discuss lock-in phase transitions.

1.4.5. Lock-in phase transitions

1.4.5.1. Phase-locked phases In §§I.4.1 to 1.4.3 we discussed the properties of phase transitions for which

symmet ry imposed Vqco(q~j)=0. We shall now discuss the properties of com- mensurate phases tbr which Vqc0(qj) is not necessarily zero. For example, a

Page 76: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

76 R . A . Cowley

tetragonal material for which qs = (0, 0, ~) and ~ = 2n/pc has a commensura te wave- vector but if p > 2 the gradient is in general non-zero. The additional terms in the free-energy equation, (I.4.49), are given by an Umklapp term involving p th order anharmonici ty so tha t

P

G=Go+a(T_T~)QIQ2 2 2 + 4uQ1Q2 + 2 (~-+ 1) V(Q[ + Q~). 0.4.75)

When this expression is t ransformed to the ampli tude and phase coordinates

1 2 G=Go + ~ a ( T - To)A + u A 4 + VAP cospc~. (I.4.76)

This expression depends upon ~b and so the phase of the distorted wave is one of the p different choices of ~b which will minimize eqn. (I.4.76). Strictly speaking within Landau theory, and for T-~ To, only the values of p = 3 and 4 are of interest. When p = 3 there are cubic terms in the free energy and so the transit ion is necessarily of first order, § 1.4.1. l, while when p = 4 the Umklapp term has the form A 4 V cos 4q~. Consequently the ampli tude of the distorted wave is given by

A2o_a(T~ - T) 4( -Ivl)

and the phases of the different possible domains of the distorted wave are q~o = 0, n/2, n, 3n/2 if V is negative and ±n/4 , ±3n /4 if V is positive. The longitudinal susceptibility can be obtained below T~ as

ZL(0) = (2a(T~ - T) ) - 1

while the transverse or phase susceptibility is given by

u-iV] 0.4.77 ) )/T(0) -- 41 Via(To _ T)'

which can be wri t ten in terms of the amplitude, A0, to give the transverse frequency a s

C0T(0)2 = (p2] V[Ag-2). (I.4.78)

These results show tha t the interaction locks the phase of the distortion so tha t the susceptibility of the phase mode is finite and of magni tude inversely proportional to tha t interaction.

Since there is no reason to expect tha t the phonon dispersion relation will have a minimum exactly at ~/p, even if it is close to tha t wave-vector , a continuous phase transit ion will normally give an incommensurate phase with q~ close to ~/p. This will be followed a t a lower t empera ture by a lock-in phase transit ion to the com- mensurate phase with wave-vector, ~/p, as discussed below.

In Nb02, however, the transit ion is apparent ly directly from a disordered phase to a commensurate phase described by a wave-vector It~ = (1, 1 1 " ~, ~) m reciprocal lattice units (Pynn and Axe 1976). Finally, when q~ is commensurate , there are only a finite set of the secondary order parameters with different wave-vectors, and these are then secondary order parameters of the first type because they do not break all the symmet ry elements. In Nb02, for example, the secondary order parameters have qj = (½, 0, 0), or (0,½, 0) or (0, 0, 0), none of which break all the translational symmet ry

[! ± !~ elements which are broken by the distortion with qs-- ~4, 4, 2J-

Page 77: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 77

1.4.5.2. Soft solitons In the preceding section we have shown tha t Umklapp terms in the free energy

lock the phase of the distorted wave for commensurate wave-vectors q~ and hence lower the free energy. We now consider the te t ragonal system discussed in §§ 1.4.4.1 and 1.4.5.1 but with a phase transit ion described by an incommensurate wave-vector qs, which is close to the commensurate wave-vector qo =~/P. The quadratic term in the expansion of the free energy is

where

1 Gz = ~7, ~ [a(T - T~) + fk2]Q1 (k)Q2 ( - k),

/ V (I.4.79)

k-q-q~ and t~o=q0-qs .

The system undergoes a phase transit ion a t T¢ to an incommensurate phase, whose free energy is given (neglecting the higher Fourier components of the distortion) by eqn. (I.4.5 a), as

Ginc = Go - - l ~ u (T¢ - T) z. (1.4.80)

The free energy of the commensura te phase with p = 4 is

a 2 ec°m = G ° - - 16(u- IVI) (Too - T) 2 (I.4.81)

where Too is the phase transit ion tempera ture tha t the commensura te phase would have had if the incommensurate phase did not occur first, T~o = Tc - f ~ / a and V is the lock-in energy, eqn. (I.4.76). For temperatures greater than T r the free energy (I.4.80) is smaller than (1.4.81) and the incommensurate phase is stable, but a t T L these equations predict a first-order transit ion to the commensura te phase where

TL=T¢- I Via

This analysis demonstrates the essential quali tat ive features of a lock-in phase transition. There is a competi t ion between the quadratic t e rm in the free energy (the frequency dispersion relation energy) which layouts ordering a t qs0 and the Umklapp interaction energy which favours a commensura te ordering, qo. At high tempera- tures the second te rm dominates, but a t lower temperatues, A o becomes large, and the interaction term dominates.

Although the above analysis contains the essential physics, i t neglects some interesting and novel features of these phase transitions. These arise because the interaction energy responsible for the lock-in also makes the secondary order parameters very large, and neglecting them as we have done so far in this section is invalid (Moncton et al. 1975).

The effect of the secondary order parameters is most easily calculated by rewriting the free energy in a cont inuum form as an integral over position, r. %n principle we should introduce a position dependent ampli tude A (r) and phase ~b(r); in practice, however, the analysis is more t ractable if we assume A (r) = Ao. This phase- modulation-only approximat ion is justified because the phase or t ransverse suscept- ibility is much larger than the ampl i tude or longitudinal susceptibility, eqn. (I.4.53),

Page 78: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

78 R . A . Cowley

and consequently it is much easier to create local distortions of the phase than of the amplitude. The free energy, eqns. (I.4.76) and (I.4.79), in a real-space representation then becomes (Bruce et at. 1978).

a=Go- l~u(yo- T)2 + G(r)dr, (I.4.82 a)

where v is the volume of the unit cell and the position dependent free energy is

1 G(r) = ~ (VqS(r) -60 ) 2 - V cospq~(¢), (I.4.82 b)

where

V= VA~- 2/f

I f the vector 6 o is along the z axis, eqn. (I.4.82) is minimized when ~b(r) is a function only of z and the phase dependent part of the free energy becomes (for a crystal of length L)

G~ 1 ~L _ 1 2 f A 2 - L ,] o G( z )dz -5ok + ~ 5 o - V, (I.4.83 a)

where now

and

1 //"~d~ 2 G(z)= 2- |~-z ' - \ / V[cosp¢(z ) - 1], (I.4.83 b)

1 k--- ~ (~b(L) -- 4(0)) (I.4.84)

is the effective wave-vector for the incommensurate phase. The problem of finding the nature of the incommensurate phase is then reduced to finding the forms of ~b(z) which minimize eqns. (I.4.83). McMillan (1976) solved this problem by writing ~b(z) as a Fourier series

~b(z) - ~z + ~ H , sin (pn&), n

and then solving numerically for the coefficients ~ and H,. More elegantly, Bak and Emery (1976) observed tha t the functions for which the integral in eqn. (I.4.83 a) is an extremum satisfy the sine-Gordon equation

d2~(z) dz 2 -pPsinp~b(z) ,

one solution of which is the soliton or domain wall (Scott et al. 1973)

~b(z) = O(z)-- 4-tan ~ (exp (pF1/2z)), (I.4.85) P

which separates a region with ~b(z)=0, say, from one in which ~)(z)=2~/p. Substituting this solution back into eqn. (I.4.83) gives the energy of a single domain wall as l/L~G(z) dz = 8 F1/2/p. Consequently the free energy of a regular array of M

Page 79: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 79

solitons with spacing b can be found, taking account of their long distance re ?ulsion, as (Bak and Emery 1976)

G, _ M 8V1/2 [1 + 4 e x p (-pF1/2b)] - k 3 o - F-~ 62 fA , L p 2

Now using the f~et that k= 2M~z/pL = 2~/pb the free energy becomes

G¢ _ M _8__ /2 2 0 + 4ML 8Ft/2 exp P ---ZL -- V+ . 1.4.86) fA~ L p

The first term is the free energy of M non-interacting solitons and i f4V 1/2 >~3o this is positive and the commensurate phase with M = #= 0 is favoured. If, however, 4 V 1/2 < ~50 a state with non-zero kand M exists which is the incommensurate phase. Thus, there will be a continuous transition from the commensurate to incom- mensurate state when 4 F 1/2 = ~6o, which using the relationship between F and the bare interaction V, occurs when

1 Ao = (~260zf/16 V)p- 2

For larger temperatures, smaller Ao, M steadily increases but is limited by the interaction term in eqn. (1.4.86). This interaction has an exponential dependence on 1/M so that the temperature dependence of the order parameter, M, is given by

2xM 2zc F t/2 - - = k = ] ~5o ' (I.4.87) pL log l ~ i / 2

This form is very difficult to distinguish experimentally from the discontinuity of a first-order phase transition as shown in fig. 1.42.

1.0

0.8

0.6

Q4

0.2

Fig. 1.42

0 0 0.2 0.4 0.6 0,8 1,0

Sketch of the order parameter k, units 60, against reduced temperature for a lock-in phase transition. (After McMillan 1976.)

The nature of the structure of the phase function q~(z) in the incommensurate phase is given by

2xm ¢(z) = 40+ + O(z-mb), (I.4.88)

P

Page 80: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

80 R . A . Cowley

where m is the closest integer to z/b. This structure was calculated by McMillan (1976) for the case of p = 3 and is i l lustrated in fig. 1.43.

This type of phase transit ion is quite different from those discussed earlier, because the order paramete r is not the ampli tude of a single Fourier component of the atomic displacements but is the number of domain walls, each of which is well localized in space. This can occur for an incommensm'ate phase because of the large number of possible order parameters for an incommensurate phase, as discussed in § 1.4.4.4. Whenever there is a very large number of order parameters , there is always the possibility of ibrming linear combinations, such as domain walls, with unexpected or non-per turbat ive characteristics.

6~ 5

~ 4.._K~ 5

i 2Tr -3-

/ /

0

Fig. 1.43

i

/

I

X

The structure of the phase function O(x) = O(z) close to a lock-in phase transition with p =3. (After McMillan 1976.)

The scattering cross-section from a distorted wave described by 2A0cos (qoz+dp(z)) can be calculated within the one-phonon approximat ion using the methods of § 1.2.3. The Fourier t ransform of the distortion is given by

F~(K) = F + (K) + F_ (K), (I.4.89)

where K is the z component of K, and

F _ ( K ) - - ~ exp(i(K-qo)z-cb(z))dz, (I.4.90)

which on substi tut ing for q~(z) from eqn. (I.4.88) becomes

M 2 g bAo ~ e x p ( i ( K - q o ) m b - i - - m ~ , F-(K)=2a-a f(K-q°)exp(-idP°)m= 1 \ P /

where f(K) is the form factor of the soliton

f(K) = exp ( i (Kz- 0(z))) dz. (I.4.91) , ) - b / 2

Page 81: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 81

The summation over m yields the total number of domain walls M provided tha t the wave-vector K is given by

2~/1 \ K=T,=qo+~-tp+I), (I.4.92)

where I is any positive or negative integer. The location of these Bragg reflections is illustrated schematically in fig. 1.44.

Fig. 1.44

T b_

C O M M E N S U R A TE

F)=3 p =4

/NCOMMENSURA TE

• , I j , . [ , , , , , , , ,' ," , , ,| ,- ,

- 2 2 - 2 2

Sketch of the Bragg reflections from an incommensurate phase with a soliton ground state.

Now in real systems the distortion of the atoms is the amplitude multiplied by the eigenvector of the soft mode and the crystal is a lattice, not a continuum. If, however, the phase qS(z) is varying only slowly within each unit celt then the intensity of the Bragg reflections can be shown to be given by

NAo2 2 I ( K ) = ~ - F(K,Roj) ~ [ A ( K + T , ) I f ( K + q o - ~ ) r

+aIK-T,)tfIK-qo-~tl2], {I.4.03/

where • is a reeiprocM lattice vector of the underlying lattice and A(K) is given by eqn. (I.2.19).

In principle the form factor of the soliton, eqn. (1.4.91), depends on the detailed structure of the soliton, but if the spacing between them, b, is much larger than the thickness, lip gt/z then they may be approximated as phase step functions, when

[(; )]' f(K-qo) = f ( T I ) = ~ + I sin (re~p)exp (-ire/p). (I.4.94)

The relative intensities of the Bragg reflections eMeulated using this form factor are shown in fig. 1.44.

Page 82: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

82 R . A . Cowley

Using the form factor eqn. (1.4.94), the sum over I in eqn. (1.4.93) can be performed to give the integrated intensity of these Bragg reflections. This is found to be identical with the intensi ty of the Bragg reflection of the commensurate phase. Consequently on approaching the lock-in phase transition, the many Bragg reflections of the incommensurate phase become closer and closer together in K as the number of solitons decreases, but their integrated intensity remains constant.

The phase excitation spectrum of a system of solitons has been calculated by us elsewhere (Bruce and Cowley 1978). In view of the fact tha t there are no direct measurements with which to compare these results we only briefly comment on the properties. There are two types of phase excitations. The first of these corresponds to oscillations of the solitons about their mean positions. Since the solitons are a regular a r ray in one dimension, z, and interact only with nearest neighbours, the dispersion relation for these excitations is given by the usual expression fbr the excitations of a one-dimensional chain. A detailed calculation gives these frequencies as

c o ~ , ( k ) 2 = S f p 2 F e i p ( - p V i / 2 b ) ( 1 - cos (to=b)) +f(#~ + #~). (1.4.95)

This dispersion relation is i l lustrated in fig. 1.45, and ibr small #z, c0s(K) is linear in ]c due to the translational invariance of the soliton array. The velocity of the phasons decreases as their separat ion b increases, and as T - + T L varies as ( T - T L ) t/z. [['he scattering cross-section of these excitations can be evaluated and also decreases to zero as T - + T L, because the number of solitons is decreasing.

The other type of excitat ion consists of fluctuations in the phase of the essentially commensura te regions of the crystal. These excitations have nearly the same dispersion relation and scattering cross-section as the phase excitations of the commensura te phase discussed in § 1.4.5.1, as shown in fig. 1.45.

1

0

1

Fig. 1.45

T=T L

T=TL+A

z~ 1 T=TL + 4A

T=TL +16A

0 0.3 0.6

W~,VEVECTOR k (~-)211

The phase excitation spectrum dose to a lock-in phase transition. (From Bruce and Cowley 197S.)

Page 83: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 83

This type of lock-in phase transit ion occurs in TaSe 2 (Moneton et al. 1975), TaS 2 (Nakanishi et al. 1977) and in T T F TCNQ (Com6s et al. 1976). The theory accounts quali tat ively for the strong satellites observed in TaSe2 and TaS2, but in all eases the transitions are not continuous but of first order, in contradietion with the prediction of the soliton theory.

There are two main defects of the theory (Bruce et al. 1978). Firstly, there is a coupling to the macroscopic strains. Inclusion of this coupling shows tha t it leads to different effects dependent upon whether or not the strain can follow the phase soliton profile. This is dependent upon whether there are acoustic modes of the correct symmet ry to describe the changes in the strains with distance. In our example, Cgtlzz/Oz is given by the longitudinal acoustic modes propagat ing along the z axis and so ~/zz strains can follow the phase solitons and give rise to a change in the effective interaction, a modified F (Bak and Tim onen 1978). In contrast there cannot be tVrlyy/c~z components because there are no acoustic modes propagat ing along z corresponding to t/yy strains. Consequently there can only be a macroscopic r/yy s t ra in , which leads to a te rm in the free energy which is proport ional to M 2 and negative; exactly analogous to the negative contr ibution to u discussed in §I.4.3.1 for conventional phase transitions. In the soliton ease, however, eqn. (I.4.86) shows tha t there is not an intrinsic positive term proportional to M 2, so this coupling to the strain necessarily drives this transit ion to be of first order. We are of the opinion tha t this la t ter type of coupling may always occur in real systems.

Secondly, if the soliton thickness is only a few lattice constants, it will be energetically favom'able for the solitons to be centred on a part icular position within the unit cell. The spacing between the solitons will then be an integral number of lattice spacings so tha t b = da, and the wave-vectors of the Bragg reflections will be given by

Under these conditions the sys tem will undergo a sequence of first-order transitions between these different commensura te states until it eventually reaches the pr imary commensurate phase with T I=2~r/ap. This sequence of first-order transitions is sometimes known as a Devil 's staircase (Aubry 1977).

The lattice may also produce a very large lock-in energy even though it does not favour any part icular d. This may then have a large effect on tile phase excitations of the system, as the phasons or soliton modes, fig. 1.45, will then have a gap which can be very large compared with the dispersion in the energy.

Throughout this section we have assumed tha t there are only two corn ponents of the order pa ramete r for the commensura te phase _+ qo- In practice the lock-in phase transit ion in TaSe 2 has three pairs of wave-vectors and so there is the possibility of three different types of solitons with interactions between them. As yet comparat - ively little work has been done on these more complicated systems. Nakanishi and Shiba (1977) suggest tha t the phase transit ions are of first order. Bak (1978) argues tha t the interaction energy between the different sets of solitons gives rise to a te rm in the free energy which is proportional to M2.. I f this interaction energy is negative, a t t ract ive, the transit ion is always of first order. I f it is repulsive then there will probably be a transit ion to a phase with only one set of these solitons frozen into the crystal followed by another first-order transit ion to a phase with all three sets of solitons.

Page 84: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

84 g . A . Cowley

Long period structures also occur in alloy systems; for example, Oriani and Murphy (1958) showed tha t in CuAu there was a s tructure with long periodicity between 650 and 690 K. At 650 K this undergoes a lock-in phase transit ion to the well-known ordered CuAu structure. The structure with the long periodicity consists of ordered regions of ordered CuAu, of about 5 unit cells in length, separated by domain walls from the next phase. This transit ion was discussed in terms of an instabil i ty of the crystal against domain walls by Inglesfeld (i972). A similar s tructure has been postulated by Pynn (1978) for the s tructure of the co-phase in Zr and Ti alloys. I n his model the co-phase in Zr0.sNb0.2 consists of the bcc structure distorted by a commensura te wave-vector ,

2 ~ ( 2 2 2~

The three different domains corresponding to this deformation are then separated by domain walls which are pinned by the local fluctuations in the concentrations of Nb and Zr a toms and so are varying distances apart .

1.4.5.3. Lock-in without soft solitons There are lock-in phase transitions to those wave-vectors for which the

dispersion relations co(q) are necessarily symmetr ic abou t q0. In these cases the linear term in Vq%(t) in eqns. (I.4.82 a, b) is absent and so the theory developed in § 1.4.5.2 must be reworked. In the neighbourhood of one of these more symmetr ic wave- vectors the dispersion relation can be wri t ten in the form

2 F 2 2 l 4-] 2 co(q)2 = co(qo) +.fl L -- 5okz + ~-kz J +f2 (kx + ky2) • (I.4.96)

which exhibits a minimum for k along the z axis a t the point kz = _ 50. Bruce et al. (1978) have discussed this type of phase transit ion using the type of approach developed in the previous section. The equation, analogous to the s ine-Gordon equation is a fourth-order differential equation, which does not, to our knowledge, have a simple analytic solution. The solutions were, however, studied numerically by looking, in particular, ibr a solution of a soliton-like character. Such solutions were found but the energy of these solutions is such tha t the commensura te phase is unstable against the format ion of an incommensurate phase with a nearly plane- wave structure before it is unstable against the format ion of these soliton-like structures. Accordingly a first order phase transition is predicted.

There are several examples of this type of lock-in phase transition. NaNO: (Yamada et al. 1968) has an incommensurate phase with a wave-vector qs ~ 2~/8 and a lock-in transit ion to a ferroeleetric phase. The transit ion is of first order and no harmonics of the incommensurate wave-vector have been observed. In thiourea (McKenzie 1975) a similar phase transit ion is again first order and occurs a t 180 K, while weak first harmonics (about 2% of the pr imary distortion) were found.

1.4.5.4. Improper lock-in transitions Some lock-in phase transitions are associated with the onset of ferroeleetricity, as

found in K2SeO4 (Iizumi et al. 1977) and (ND4)2BeF4 (Iizumi and Gesi 1977). In

Page 85: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 85

these cases the lock-in terms in the free energy are of the form

P 2(7 + ~)W(Q~ + Q~)Q, (I.4.97)

where Q is the ferroelectric secondary order parameter . In an incommensurate phase with wave-vector qs terms of the form of eqn.

(I.4.97) will necessarily give rise to distortions of the secondary order paramete r with a wave-vector 4- q = _ (pqs + ~). This gives rise to a lock-in energy only if 4- q are the same wave-vector when there will be a constructive interference between these two distortions resulting in a lower energy for the resulting commensura te phase. Of course, there will also be the possibility of lock-in due to terms of the form

2(p+ 1)Vt(Q21 p .~_ Q~ p) ,

for these wave-vectors. This t e rm is usually non-zero bu t has frequently been neglected in the theory of these materials.

When a plane-wave ground state for the incommensurate phase is assumed the theory can be evaluated as given in detail by Iizumi et al. (1977) for K2Se04, and the analysis is very similar to the t r ea tmen t of secondary order parameters in § 1.4.3 and of lock-in phase transitions a t the beginning of §I.4.5.2. Since, however, the dispersion relation is not symmetr ic about qs, this analysis requires fur ther development to examine if soliton-like structures can modify the behaviour. This approach has been developed by Bruce et al. (1978) and soliton-like structures may occur in the phase of the distorted wave. Since, however, the coupling is through a secondary order pa ramete r the free energy includes a negative te rm in M 2, where M is the number of solitons. Consequently, as discussed in § 1.4.5.2, the transit ion is always of first order because there is no positive te rm in M 2 in the soliton interactions.

S hiba and Ishibashi (1978) have also analysed this problem using a Fourier series expansion of the distortions Q~, Q2 and Q. They find a continuous transit ion with properties very similar to tha t of soliton theory given in § 1.4.5.2. We believe this is because their numerical work is not sufficiently accurate.

The lock-in transitions in both K2Se04 (Iizumi et al. 1977) and (ND4)2BeF 4 (Iizumi and Gesi 1977) are of first order. In the former case, p = 3, relatively strong first harmonics of the pr imary distortion were observed, while in the la t ter case, p = 2, no harmonics have as ye t been reported.

§ 1.5. M~cRoscoe~c THEORIES In this section we review the different microscopic models which may be used to

calculate the phenomenological parameters of Landau theory. The statistical mechanics of the models is discussed in only the most e lementary way, but we focus on the extent to which the models provide a more microscopic picture than Landau theory. Alas the very success of Landau theory in correlating experimental data means tha t any microscopic model with a few parameters may well be able to fit a large amount of data, but the model m a y or may not be a correct microscopic description. We therefore examine the different models critically and find tha t very few can be said to be a justifiable microscopic and quant i ta t ive description.

Page 86: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

86 R . A . Cowley

1.5.1. The weakly anharmonic crystal models

1.5.1.1. The soft mode In crystals with large and temperature independent band gaps, such as SrTiOa,

structural phase transitions of the displacive type are almost certainly correctly described by the anharmonic crystal model. The theo .ry of the weakly anharmonic crystal was developed in the early 1960s by Maradudin and Fein (1962) and by Mar~dudin and Flynn (1963) as reviewed by Cowley (1963). I t was applied to ferroelectrics by Silverman and Joseph (1963) and by Cowley (1964).

The basic result is tha t the one-phonon susceptibility or scattering cross-section can be obtained from the expressions for a harmonic crystal if the harmonic frequencies ~o(qj) are replaced by

(~(qj, ~o)2"= ~(q j)2 + ~ (q j , ~o), (I.5. l)

where E(q~,(9) is the anharmonic part of the self-energy and depends on the frequency, o), used to s tudy the mode in an experiment.

There are three contributions to the self-energy to lowest order in the anharmonici ty as shown in fig. 1.46. The first contribution arises from the thermal strain, T t/~p, and is obtained from eqn. (I.2.10) as

E v ( q j ) = 2 ~ V ~ ( q ; . q ) ~ . (I.5.2)

This contribution can be obtained from measurements of the strain dependence of the frequencies of the mode and a knowledge of the thermal strain, l~ecently Samara et al. (1975) have shown tha t ET(clj) decreases with increasing temperature for all the ferroelectrie materials they studied, while it increases for all other types of transition. We are u~aware of any reason to expect this to be a universally correct observation.

Fig. 1.46

(b)

(g)

Diagrammatic representation of the anharmonic contributions to (a)-(d) the self-energy of a normal mode and (e)-(g) the quartic term in the Landau expansion of the free energy.

Page 87: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 87

The second contribution to the self-energy in fig. 1.46 arises from the quartic terms in the anharmonic expansion and contributes, using eqns. (I.2.5) and (I.2.7),

E4(q j )=6h ~ V ( ~ - q q ~ - - q l ~ ( 2 n ( q l j i ) + l ) (1.5.3) qlJl J Ji Jl ] c°(qlJl) '

which at high temperatures hco(qijl)<</cBT is proportional to T. This term usually gives a contribution to the self-energy which increases with increasing temperature.

The third contribution shown in fig. 1.46 arises from the cubic anharmonici ty and yields

• " J l J 2 / I

x (co(q ljl)co(q 2J2))- l[R(co) +/uS(co)l, (I.5.4)

where

R(co) = (n I -k n a + 1) (((o 1 + co2 + co)p i _ (co _ col --co2)p i)

+ (n l - n 2 ) ((co - c o l + co2)p 1 _ (co + col - co2) ; 2)

and

S(co) = (n i -~7b 2 -~ 1) (5(00 --COl --(02) --'6(CO-[- (2) 1 -[- (/)2))

-~ (nl --%2) ((~(CO -]- col --0)2) --5(CO --COl -[- CO2))~

where col =co(qlJi), etc. This term has both real and imaginary parts which depend on the frequency co. The static real par t E3(qj, 0) is negative and increases in magnitude as the temperature increases. The imaginary par t is odd in co and so at low frequencies can be writ ten as proportional to co.

The change in the free energy of the crystal/unit cell due to a static distortion, Q(qj), is given by (Cowley 1964)

1 z G = G O + ~ ( q j , 0) Q(qj)Q( - q j). (1.5.5)

This result shows tha t the free energy of the distorted crystal is determined by the same frequencies as the static susceptibility. They may therefore be equated with the frequencies of eqn. (I.4.16) as

co(k) = &(k + q~j, O)

The calculation of the static frequencies, 5)(q j ,0) , is then a calculation of the parameters of the quadratic term in the Landau expansion, eqn. (I.4.1). The Landau theory is recovered if c5(q~, 0) z = 0 at T¢, and is positive for higher temperatures; the change in c5(q~j, 0) with temperature gives the parameter a.

These results imply firstly tha t the harmonic frequencies are imaginary, so that the crystal is only stabilized by the anharmonic terms. Secondly the Landau theory is obtained only if the anharmonic self-energy is analytically well behaved at T¢. The former is the case if the terms in the summations over (qiJi) in eqns. (I.5.3) and (I.5.4) for which (qlJl) are the soft modes make a negligible contribution to these summations. Strictly, this approximation is invalid close to T~ as discussed in detail in paper II , but Landau theory is valid if these terms are relatively small which is probably the case for tempera-tures not too close to T~.

Page 88: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

88 R . A . Cowley

At low frequencies we can make use of the low frequency expansion of E(q~, co) to give

cb (qsJ, co)2 = co(qsj)2 + E(qsj, 0) - 2ico70,

where 70 is a constant. This gives the one-phonon response function as a classical oscillator, eqn. (I.3.22), but is only valid for the range of co for which it is possible to replace E(q~j, co) by its low-frequency expansion. We shall discuss this fur ther in paper I I I , but briefly eqn. (I.5.4) suggests tha t it is valid for frequencies co < a typical aeoustie mode zone boundary frequency provided tha t in eqn. (I.5.4) there are not many terms for which co(qljl)=co(q2J2)-

The higher order terms in the Landau expansion can be obtained in a similar way as shown schematically in fig. 1.45 for the quartic term. Since the leading terms in these expressions are independent of temperature , the Landau theory approxim- ation, tha t u is independent of temperature , is clearly a reasonable approximat ion close to T¢. Weakly anharmonie per turbat ion theory provides explicit expressions for the parameters of Landau theory in terms of the anharmonic interactions and normal modes. There is, however, still a quant i ta t ive problem because in most simple materials, the anharmonic corrections to the frequencies are a few per cent for temperatures up to the Debye temperature . We need therefore to unders tand why the corrections are larger in materials undergoing structural phase transitions and to do so with reasonable interatomic interactions. In the next section we describe a t t empts to do this for SrTi03.

1.5.1.2. Models of SrTi03 There have been m a n y a t t emp t s to develop microscopic models to describe the

behaviour of SrTi03. Since, however, we are concerned in this section to examine if the behaviour of SrTiO3 can be obtained using reasonable values of the interatomic forces we shall discuss only the most detailed model used, which happens (!) to have been developed largely by Bruce and Cowley (1973). Stirling (1972) measured the phonon dispersion relations for SrTiO3 and fitted these to obtain the parameters of a shell model for the interatomic forces in SrTiO3. This model which had short- range Ti-O and Sr-O interatomic forces and much weaker forces between neighbouring oxygen atoms, provided an accurate description of the frequencies and eigenvectors of the normal modes in SrTi03. The model was then extended to include the third and fourth derivatives of the interatomic tbrces between the T i -O and Sr-O atoms. The third and fourth derivatives of the S ~ O potential and the third derivat ive of the Ti-O potential were fitted to the anharmonic terms in the Landau theory of the l l 0 K transit ion in SrTiO3 as developed in §§I.4.2 and 1.4.3 and initially by Slonezewski and Thomas (1970). With these three parameters it was possible to obtain reasonable values of the Landau parameters and to calculate the different contributions to the self-energy of the soft mode. In fig. 1.47 we show the frequency dependence of E 3 (qsJ, co) a t 300 K. I t shows considerable s tructure but for co < 3 THz, the real pa r t is well approx imated by a constant and the imaginary par t is a lmost linear in frequency as required for the classically damped oscillator to be an accurate description of the soft mode. The different contributions to the self-energy of the soft-mode are shown as a function of tempera ture in fig. 1.48. As expected the thermal strain contr ibution is small, and E 3 and E 4 are both large and to a large extent cancel. I t is clearly inappropr ia te to calculate one of these contributions without also including the others. The agreement between experiment and

Page 89: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 89

~L m

Fig. 1.47

I • / r /

i l l I I

! % j

t !

t I

' ' 2'O 2'~ ~'O 3'~ 5 I0 15

FREQUENCY (THz)

E 3 (q)', ~o) of'the R 25 mode as a function of co for SrTi03 at 300 K (after Bruce and Cowley 1975). The real part is a dotted line and the imaginary part a solid line.

calculation shown by fig. 1.48 is very gratifying, as is the agreement of several other properties shown in table 1.2. The theory also gives a good description of the widths of the infra-red active q = 0 modes as shown by the calculated refleetivity in fig. 1.49. Since the numerical values of the anharmonie force constants were not larger than would be expected from the usual 1/r 12 behaviour of short-range forces, the results of these calculations were very satisfactory and showed that at least for SrTiO3, the weakly anharmonic model provided a satisfactory description of the zone boundary soft mode.

The temperature dependence of the q = 0 modes also involves the tburth derivative of the Ti-O force. When this was obtained from the observed temperature dependence its value turned out to be large and furthermore the other properties which depend upon it, such as the non-linear dielectric constant, were not given accurately by the model, l%cently Migoni et al. (1976) have suggested that the strong two-phonon Raman scattering arises in a shell model from a very anharmonie shell- core coupling and further tha t this coupling contributes to the temperature dependence of the ferroelectric mode in perovskites. This may explain why our calculations of those effects which depend upon the fourth derivative of the Ti-O forces were unsatisfactory, but as yet Migoni et al. have not included the other anharmonic effects in their theory.

Although further calculations will no doubt be performed for SrTi03, we believe that these results have shown that anharmonicity is the origin of the temperature dependence of the soft modes in these materials. Furthermore the anharmonic parameters for the interatomic tbrces are very much the size to be expected from work on alkali halides. We are left then with the basic question as to why the effect of the anharmonicity is so much larger in SrTi03 than in, say, KBr. In SrTiO3 the mean

Page 90: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

90 R . A . C o w l e y

Fig. 1.48

m z i.u

12

l o

J J f I f f I i f I

f J

J J

2-

o

- i . . . . . . . . . . . . . ' ' ' ' ' " " ' " " ' "" "" " " " • " "

I i J | i i I 0 5 0 1 0 0 150 2 0 0 2 5 0 3 0 0

IEMPERATURE (OK)

/ /

I /

J J

J J

f J

J

f

J J J

• . . . . .

• . . . . . • . . . . . .

I I I

3 5 0 4 0 0

The different contributions to the seKenergy of the R25 soft mode in SrTi03. The dashed line gives Z 4, dotted line E 3, fine dashed line Y r, and the solid line c~ 2. The circles are experimental results. (After Bruce and Cowley 1975.)

Table 1.2. Properties of SrTiO3. (After Bruce and Cowley 1973.)

Quantity Units Experimental Calculated

Order parameter, Q~/T¢ - T 10- s A2 K - t 0"165 0-167 coL/o9 v 9"4 9"25 g a m a n frequency ratio, 2 2

Ratio of g a m a n frequency to frequency above To, co2/co(qj) 2 2-4 2"33

dT¢ 1 0 - 2 K b a r -1 0'17 0"178 d,p

dTo 10 z K bar- 1 0"33 0"294 d ] ) ( 1 1 D

Thermal expansion at 300 K 10 -6 K - 1 10'0 8'76

Elastic distortion, ~ - 10 5 K - ' 0"83 0'77 To - 7'

Gruneisen 7 103 5-67 5-17 Change in soft-mode self' energy

between 110 and 4 5 0 K Tttz 2 4"3 4-86 Soft-mode damping c o n s t a n t ~'0 at 110 K T H z 0"1 0-06

Line width of q=O optic modes at 85 K and 47 K T H z 0"09 0"08

<0-06 0.01 0-15 0.17

Page 91: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 91

Fig. 1.49

7O

&

50

D

1o

i i i i i i i |

6 10 14 18 22 26 30 30 50 70 90 110 130

W A V E L E . G T . CM,CRONS

' ' 2 ' 0 ' ' ' , 3O 15 1 5 3

FREQUENCY (THZ)

The reflectivity of SrTiO3 compared with measurement of Spitzer et al. (1962). (After Bruce and Cowley 1975.)

frequency of' the normal modes is about 8 THz, so t ha t the self-energy effect for the soft mode a t 450 K, fig. 1.47, is only about 8~o of the mean frequency squared. This is comparable to the anharmonic self-energy in alkali halides at these temperatures. Consequently in SrTiO 3 the anharmonic effects are not large, but the harmonic frequencies of the self mode are very small. Possibly this arises because perovskites have three types of ion and only one free parameter , the lattice constant. I t is then very likely tha t one of the a toms is only a loose fit in its site and so can easily be displaced from its h igh-symmetry position. This type of a rgument would seem to be nearly as far as we can go in explaining why many perovskites exhibit structural phase transitions.

1.5.2. Order~lisorder systems

1.5.2.1. The time-dependent Ising model A spin representat ion for describing order-disorder systems was introduced in

§ 1.2. We shall now discuss the calculation of the parameters of the Landau theory for this model and the t ime dependent properties using the theory of Glauber (1963) and as developed by Suzuki and Kubo (1968). In their model each spin interacts with a heat bath, which induces spontaneous spin flips. I t is then assumed tha t each spin relaxes back towards its local equilibrium configuration with a t ime constant, %, which in the simplest form of the theory is independent of the configuration of the neighbouring spins.

In the interest of simplicity we consider a system with one spin in each unit cell, S(1)= _½, when the hamiltonlan of the system in an external field E(I) can be

Page 92: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

92 R . A . Cowley

obtained from eqn. (1.2.11) as

H = 2 ~ c~2 (ll')S(1)S(l') - ~, E(l)S(l), l l ' l

where we have assumed tha t ~bl(l ) is included in the external field. The local molecular field on the site l is

Em(l ) = E(l) - 4 ~ +2(ll')<S(l')> I '

and hence within the simplest t ime-dependent model the equation of motion is

d 1 ~OdT<S(/)} = - <S(/)} + 2 tanh (flE,,(l)/2). (I.5.6)

The solution of this equation, is well known. The susceptibility above T¢ is obtained by expanding the tanh function to leading order when, af ter a Fourier t ransformation,

)/(q,~o)=)/(q)(l+ico~(q)) ~, T>T:, (I.5.7) where the static susceptibility is

1 )~(q)-4(kBT-J(q))'

and the interaction J(q) is given by

J(q) = - ~ qb:(//') exp (iq. (R(/) - R(/'))). l '

The relaxation t ime is given by

z(q) = ~(1 - f i J (q ) ) i (I.5.8)

The static susceptibility, eqn. (I.5.7), shows tha t the phase transit ion will occur for the wave-vector q: for which J(qs) is a max imum and at the tempera ture T c such tha t kBT c = J(q:). The Debye relaxation form has a relaxation time: z(q), which for q = q s diverges as ( T - T ¢ ) -1

We can obtain the parameters of the Landau theory for these systems by comparing these results for the susceptibility with those obtained in § 1.4, and also by comparing the expressions for the order parameter , Qo =N-1/2(S(qs) ) for T < To. The la t ter is obtained from eqn. (1.5.6) by expanding the tanh to third order terms to give

Q2= 3 ( k B T - J ( q s ) ) k2T2

4J(qs)3

when we can identify the parameters of the Landau theory as

and close to T c

R(k) = 4 ( k , T - J(q~ + k)), (1.5.9)

4 u=~kBT¢ . (I.5.10)

Page 93: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 93

Consequently the parameters of Landau theory can be obtained with this model once J(q) is known, and the relaxation t ime ~0. A feature of this model which is simpler than the weakly anharmonic crystal model is the way in which the same interaction J(q~) determines both T¢ and u, whereas in the other model there is no direct relationship between them.

1.5.2.2. The tunnelli~2 model In the introduction we described an order-disorder system by using a spin

variable S(l) which, for example, described the orientation of the NO2 molecular groups. In some systems quantum-mechanical tunnelling between the two wells lifts the degeneracy of the two orientations and the wavefunctions are either the symmetric or an t i symmetr ic linear combinations. This type of two level system can always be described by an S =~ spin system but the spin is now a vector, in this case the S ~ component describes the location of the ion and the S x component the tunnelling between the sites. A careful discussion of the derivat ion of the effectiye spin hamil tonian for the KH2PO 4 system has been given by Tokunaga and Matsubara (1966) and is reviewed by Blinc and Zeks (1974). In the simplest case the hamiltonian of the tunnelling model with one spin/unit cell is

H = 2 Y. ~2 (ll')S~(0S~(1 ') + ~ (2nSX(l) - E(1)S-~(1)) (I.5.11) l l ' 1

where ~ is the tunnelling integral, and we have assumed tha t the external field couples only to the z component of the spin.

This hamil tonian can be solved within the molecular' field approximat ion as described in the last section and by Blinc and Zeks (1974). The molecular field is a v e c t o r

Em(l ) = ( - 2 ~ , 0, - 4 ~ ~b2(ll' ) <S=(I')} + E(1)) I '

and the static susceptibility becomes

1 z(q) 4(fl eoth f l ~ - J ( q ) ) (I.5.12)

This shows tha t a phase transit ion occurs when coth tic ~ =J (qs ) /~ , and tha t the ordered phase is described by the wave-vector qs for which J(q) is a maximum. This result reduces to tha t of § 1.5.2.1, namely kBT c = J(qs) where g2 is small, and as the tunnelling integral becomes larger there is a decrease in To. This explanation for the transition tempera ture being much lower in K t l2PO 4 than in KD2PO 4 was the original mot ivat ion for the development of the tunnelling model (Blinc 1960).

The parameters of the Landau theorv are given by

~o(k) 2 = R(k) = 4(~ coth fl~ - J(q~ + k)), (I.5.13)

which close to T¢ becomes

R(k) =4[( L\ sinh fl~F~ (fic~)/~2ks(T-Tc)-J(q'+k)+J(qs)] , (I..5.14)

showing tha t the parametei ' a depends on the tunnelling integral. The quartic parameter , u, in eqn. (I.4.1) can be found by calculating the

tempera ture dependence of the order pa ramete r using the molecular field theory.

Page 94: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

94 R . A . Cowley

The result, after some algebra, is tha t

2J3(q~)[ flcJ(q~) ] u = ~ 1 cosh2 (fl¢~)

(1.5.15)

which reduces to the result (I.5.10) in the Ft-~0 limit. These results show tha t expressions for the Landau parameters may be obtained

from the tunnelling model. The effect of the tunnelling is to decrease T c and a, and to i n c r e a s e u .

In addition to these static results, the tunnelling model has also been used to calculate dynamical properties. Within the random phase approximation the Heisenberg equations of motion for the spins lead to a tunnelling frequency for T > T c (Brout et al. 1966, Blinc and Zeks 1974) given by

h2(DT(k) 2 = 4 ~ ( ~ - J (q~ + k) tanh fl~). (I.5.16)

For k = 0 this frequency decreases to zero as T - T o , but is given by a different expression to the 'frequencies' of the Landau theory, eqn. (I.5.13). These tunnelling frequencies are quantum mechanical in origin and not .the same as the Landau frequencies. This is because in the tunnelling model the spin is a vector and the tunnelling corresponds to the application of a field in the x direction. This causes a par t of the spin to align along x, with a consequential reduction in the fraction which aligns along z, and hence in the susceptibility which depends directly only on the z component.

Before proceeding it is interesting to compare the tunnelling model with the Glauber model discussed in the previous section. In both models a mechanism is put forward by which the particle at site l changes its spin. In the Glauber model, this is by contact with a heat bath when a thermal fluctuation permits the particle to surmount the potential barrier between the two spin values, while in the tunnelling model the particle tunnels through the barrier quantum mechanically. Clearly the former process will tend to be the dominant process at high temperatures and with heavy molecular groups~ while the lat ter will be the more likely at low temperatures and if only light atoms are moving at the transition. This is the case at a number of orientational transitions.

1.5.2.3. Spin~phonoT~ interaction In order-disorder materials there is often a coupling between the spin and the

displacements of the atoms. For example, in NAN02, the force on a displaced Na ion is clearly dependent upon the orientation of the neighbouring NO2 groups (fig. 1.3). The interaction can be expanded as a double power series in the spins and the displacements, but we assume (although it is certainly not always valid) tha t only the leading term is important,

HI= ~ W~(ll')u~(1)S(l'), (I.5.17) ctll '

where for simplicity we are not including tunnelling effects in this section~ and W~(ll') gives the change in the force on atom 1 when the l' orientation is altered.

I f the motion of the displacements is very much more rapid than that of the spins the displacements will take up average positions which are dependent on the

Page 95: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 95

particular spin coordinates. These average positions are most readily found by rewriting the interaction in terms of Fourier coordinates

H,= ~ tl"(qj).Q(qj)S(-q), (I.5.18) qJ

when the instantaneous equilibrium positions of the displacements are given by

W(qj) Q(q j ) - °~(qJ) 2 S(q), (I.5.19)

and the normal modes of vibration are now oscillations about these displaced positions, but in this approximation their frequencies are unchanged.

Using eqn. (I.5.19) to eliminate the Q(qj) in the hamiltonian for H t and tile phonon hamiltonian, leads to an effective interaction between the spins which in Fourier space is given by

j (q)= J(q)_1_ y lW(qj)l 2 (I.5.20)

8 T • co(q j) 2 "

U~ffortunately, this result is not completely correct because J(q) is derived from the interactions between different sites l and l'. An interaction between a spin and itself cannot contribute to the ordering of the spins and consequently tha t par t of the phonon modulated interaction must be subtracted from the effective interaction. The correct result (Kanamori 1960) is

1 I w(q./)12 1 I w(q'J)t: g ( q ) = J ( q ) - 7 ~ - -t- ~ q~j -. (I.5.21) o j c0(qj) 2 c0(q~j) 2

The effect of the spin-phonon coupling is then to lead to a modification of the interaction between the spins. This is of particular importance in tha t the acoustic mode part of eqn. (I.5.2t) may lead to a long-range interaction between the spins which falls off only as 1/R 3 (Sugihara 1959).

Another consequence of the coupling is tha t the susceptibility or scattering is now a coupled phonon-spin problem. For example, the scattering cross-section from the spin variables, eqn. (I.2.45), is now modified to give an effective structure factor

Fs(K,qj)=Fs(K, qj) ~ F ( K qj), (q3)

showing that there will be an interference between the scattering from the spin and the phonon modes.

Much of the theory discussed in this section was developed initially for coupled spin-phonon problems in magnetism. I t was applied to order~disorder systems by Sakurai el al. (1970) and by Yamada et al. (1972) and to the tunnelling model by Kobayashi (1968).

FinMly, in tiffs section we have assumed that the normal modes of vibration could follow the spin coordinates. This is not possible for the long wavelength acoustic modes and in the tunnelling model may not be possible for other modes. The dynamics can then only be discussed once the dynamical response of both spin and normal mode systems is known and their coupling will give rise to effects similar to those described for coupled acoustic and optic modes in § 1.4.1.5.

Page 96: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

96 1~. A. Cowley

1.5.2.4. Applications to real systems Throughout this article NAN02 has been used as an example of an order-disorder

system. The first application of the order-disorder theory to NAN02 was by Yamada and Yamada (1966) to explain their diffuse X-ray scattering measurements (fig. 1.19). The NO2 groups have a dipole moment which may be aligned along the positive or negative b axis (fig. 1.3). They calculated the dipolar interaction between the rigid dipoles and found that this par t of J(q) = - D ( q ) gave a contribution with a maximum very close to the wave-vector of the incommensurate phase, and a transition temperature only 50% higher than the observed one (fig. I. 19). When they introduced short-range forces between neighbouring dipoles they were able to produce a vel:y good agreement with their experimental results with only three parameters specifying the three nearest neighbour forces.

Unfortunately further developments have questioned the validity of this agreement. In the ferroeleetric phase of NaNOz the Na are displaced from the mid- points between the NO2 ions (Kay 1972). This means that in addition to the direct dipole dipole interaction there is also a phonon-modulated interaction as described in the previous section. When Sakurai et al. (1970) used models describing the normal modes of NaN02 to include this extra interaction between the NO2 groups, they were no longer able to obtain a stable incommensurate phase. The situation is therefo~,e unsatisfactory.

The dynamics of the ordering in NAN02 was studied by H a t t a (1968) using dielectric measurements (fig. 1.13). These results are well described by the model described in §I.5.2.1 with a relaxation time, ~o, given by Eyring's reaction rate theory as

2=h ~0 = ~ exp (UB/kBT),

where UB is the potential barrier between the two positions of the N O 2 groups. Assuming that UB is weakly dependent upon temperature but approximately 7"4ksTN and taking J(q) from the work of Yamada and Yamada (1966) described above, Yamada el al. (1968) obtained a good description of the experimental results as shown in fig. 1.13.

The ammonium halides crystallize in the CsC1 structure, and the tetrahedral NH4 molecule has two possible orientations depending upon which four of the neighbour- ing halide ions are closest to the H atoms. These two orientations can be described by a spin variable as introduced in § 1.2. The distorted phase of NH4Br and NH4I is the y-phase which has an antiferrodistortive arrangement of the spins described by a wave-vector q~=(½,½,0)2~/a. On the other hand NH4C1 and NH4Br at low temperatures form the a-phase which is a ferrodistortive phase, q, = 0.

The direct interaction between the NH 4 ions is an oetupole~octupole force and the corresponding J(q) was calculated by Nagamiya (1943). The result is shown in fig. 1.50 and favours the ferrodistortive phase. The X-ray structure determination of NH4Br by Terauchi et al. (1972) showed tha t the Br ions become displaced at the transition. Consequently Yamada et al. (1972) developed the theory using the spin- phonon approach described in § 1.5.2.3. Although the resulting theory inevitably involved a large number of parameters, it did show tha t the spin-phonon coupling favoured an antiferrodistortive ordering as illustrated in fig. 1.50. The total effective exchange interaction J(q) might then favour either ferrodistortive or antifer- rodistortive distortions dependent upon the size of these two terms.

Page 97: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transi t ions 97

Fig. 1.50

4

2

~2 ~o

'0 x

Z bJ

-E

/ 7 " - . ! "">/"

,.T.. (~)

J(~)

\

\ \'__....~\

F (C ,~ o) M (, /2 J/a C } R

The exchange interaction between spins in the ~mmonium halide NH4Br after Yamada et al. (1972). The direct interaction is given by . . . . , the spin-phonon by . . . . . . . , and the total effective coupling by - -

Another mechanism favouring the ant idistort ive structure was proposed by Hiiller (1972), who suggested tha t the octupole on the NH4 ion produced an electronic dipole on the neighbouring halide which then produced an electric field on the other NH4 ions. This is clearly similar to the phonon mechanism except tha t the phonon mechanism produces an ionic dipole instead of an electronic dipole. At present the relative magni tude of these different effects is still uncertain, as reviewed by Vaks and Schneider (1976).

Possibly even more suecesstul has been the theory of the phase transitions in methane. There are several phase transit ions a t which the different methane molecules within the unit cell order in position. J ames and Keenan (1959) suggested tha t the force was an oe tupo l~oe tupo le force and this gives the correct structure (Press 1972), and the anisotropy of the critical scattering (Hiiller and Press 1972). Although the agreement with exper iment is not perfect, it is sufficiently sat isfactory to suggest tha t the nature of the force between the methane molecules is largely understood.

In other molecular materials such as N2 the tbrm of the intermoleeular interaction and the effect of the electrostatic forces in polarizing the ions is less well understood. For some recent work on the problem of obtaining a potential for Nz, we refer to the work of Kjems and Dolling (1975) and of Raieh and Gillis (1977).

Finally we comment on the validi ty of the tunnelling model for describing the KH2PO 4 (KDP) class of materials. The original mot ivat ion behind the development of the tunnelling model was the ease with which it explains the large isotope effect (T c for KH2PO 4 is 123 K, and is 220 K for KD2P04) since the change in mass will clearly make a large difference to the tunnelling integral. This was supported also by the results of dynamical Igaman scattering experiments which showed for KH2PO 4 an overdamped mode but with a considerable spectral width (fig. 1.22), while for KD2PO 4 the spectral width was a t least an order of magni tude smaller (White et al. 1970). This is in agreement with the predictions of the tunnelling model if f~ is much larger for K H z P O 4 than for KDzPO 4, a l though the simple model fails to explain why

Page 98: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

98 R . A . Cowley

the mode is overdamped. These successes were also supported by the result of the measurements of the decrease in the transition temperature with increasing pressure by Samara (1971) and of an underdamped soft mode at high pressure by Peercy (1973). These results were assumed to arise because the H - I t distance decreased with increasing pressure and hence the overlap integral rapidly increases.

The difficulties for' the theory are tha t all the atoms move in the ferroelectrie motion (Skalyo et al. 1970), consequently there is a strong coupling between the hydrogen tunnelling and the other normal modes. Although this can in principle be included using the theory of Kobayashi (1968), this of necessity requires the introduction of many parameters, which makes agreement almost inevitable. Paul et al. (1970), in fact, conclude tha t it is not possible to obtain a good agreement between experiment and theory by assuming tha t the only effect of deuterat ion is to change the tunnelling integral. Fur thermore Nelmes et al. (1978) have performed a structure determination under pressure which suggests tha t the difference between the minima in the H potential is not pressure dependent. I f this result is confirmed it invalidates the simple tunnelling-model explanation of the pressure dependent results. In conclusion Nelmes et al. (1972) show that the structures of KHzPO 4 and KD2PO ~ differ so it is hardly to be expected that the model will be appropriate if only the tunnelling integral is assumed to depend on deuteration.

This leads us to the question of whether the tunnelling model is a good description of the K D P materials. The motion of the H atoms is certainly correlated with the motion of all the other atoms in the structure. This coupling alters the effective exchange coupling, § 1.5.2.3 (Kobayaski 1968), but its effect on the tunnelling integral gt is never discussed. Since the neighbours of an H atom move, the potential is locally not symmetric about the centre making the tunnelling to the other side more difficult. If, however, a cooperative tunnelling model is adopted the effective mass of all the atoms is clearly sufficiently large that tunnelling will be relatively unimportant compared with the hopping described by the Glauber model, § 1.5.2.1. Certainly the results on K D P could be described by a soft mode with a large and vet:y temperature dependent damping, and a frequency which was altered greatly on deuteration or by the application of pressure. Clearly further work is needed before we have a microscopic theory of these systems which.lie between the displacive and order-disorder extremes.

1.5.3. Electron-phonon models

1.5.3.1. The formalism I t is frequently convenient to divide the interatomic force between two ions into

two parts; a direct ion-ion interaction and an interaction mediated by the electrons, although the distinction between the ions and the other electrons is dependent upon the particular problem and assumptions. The calculation of the frequencies of the normal modes, or of the energies in the spin representation, is then a very complex subject as shown by the reviews in Hor ton and Maradudin (1974). In this section we a t tempt only to outline the essential points of the development wi.th particular reference to the effects which may lead to structural phase transitions.

The potential seen by an electron in the crystal at the point, r, may be expanded as a power series in the normal-mode coordinates

W(r)= W°(r) + ~ W(qj, r)Q(qj)+ . . . . (I.522) qJ

Page 99: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 99

The change in the potential at r due to a normal mode (q j), then causes a change in the electron distribution at the point r' which results in a change in the energy of the normal mode. I f we assume that the symmetry is such that mode q j couples only to itself, then the electron-phonon contribution to its self-energy is

E E (q j ) = - ~ W(qj, r ) W ( - q j , r')z~(r, r') drdr', ([ .5.23)

where )f(r , r') is the electron density response function. The frequency of the normal mode is given by

co(q j) 2 = coB(q j) 2 -t- EE(qj), (I.5.24)

where coB(q j) is the contribution to the frequency arising from the direct ion ion interaction, and also from the second-order terms in the expansion of W(r) in powers of Q(qj).

I t is often convenient to rewrite eqn. (I.5.23) in terms of Fourier components by making use of translational invariance when

ZE(qj) = - ~ W(q÷v,j)W(-q-'c',j))f(-q-~,q+z'), (I.5.25)

where

W(q ÷ ~,j) = ~ W(qj, r) exp ( - i(q .1. ~). r) dr. (I.5.26)

I t is of interest to compare this result with the developments of §§ 1.4.1.6 and 1.5.2.3. In the latter case o)(qj) z is the inverse susceptibility of the phonon normal mode with which the spins are inteiucting; eqn. (I.5.20) for J is then of the same structure as eqn. (I.5.25). The extra complication of the sum over • and v' arises because the electron distribution is defined for all r, whereas the phonon displace- ments or spin variables are defined only at the lattice points.

Equat ion (I.5.24) shows that a structural phase transition will result if E ~ becomes sufficiently negative, while eqn. (I.5.25) shows that this may occur if either the e l ec t ro~phonon interaction, W, is sufficiently large or the electronic suscept ibility is sufficiently large. In particular eqn. (I.5.25) shows tha t whenever there is an instability of the electron gas, ) f ~ , then unless the electron-phonon coupling happens to be zero, there will always be a structural phase transition which occurs at a temperature somewhat in excess of the temperature at which the electronic system alone is unstable (cf. the difference between the clamped and free transition temperatures in § 1.4.1.6). We are therefore concerned with the possible instabilities of the electronic system, which may take on a variety of different forms; for example, Jahn-Tel ler instabilities, charge density wave instabilities, spin density waves, Wigner crystallization and the Mort metal- insulator transition.

Before discussing some of these transitions in more detail, it is, however, useful to collect together some expressions (Sham 1974) for the electron-phonon interaction and the electronic susceptibility for systems with electron energy bands. I f the electron-ion pseudo-potential is v~(r - r(/~)) then the electron-phonon interaction is given by

( ] ~:/2 W(qj, r ) = ~ \ N ~ / e(K, q j ) . V, (,~)v~(r - R(l~)) exp ( iq . R(lK)). (I.5.27)

Page 100: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

100 1~. A. Cowley

I f the electrons do not interact with one another and are described by Bloeh waves with a wave-vector k and band index 2 and have energy v(k2), then the electronic susceptibility becomes

~ n(k2) - n(k + q2') z ° (q+z , -q--z)= - - ~ + ~ ~ ~ k ~ q ~ ~

(2kl exp (i(q + z ) . r)lk+ qX} (k +q,~'l exp ( - i ( q + z ) . r)lk;~ } , (I.5.28)

where n(k~) is the Fermi occupation number. In practice electrons interact with one another through a potential whose Fourier

t ransform we write as vE(q). Within the random-phase approximat ion the electronic susceptibility is then given by a mat r ix equat ion in z and z' which we can write schematically as

Z E = Z ° - X°v~x ~. (1.5.29)

The form of vE(q) is in general unknown: within the Har t ree approximat ion vE(q) is given by the Fourier t ransform of the Coulomb interaction, but usually modifications are made to this form of vE(q) to incorporate the effects of exchange and correlation a t least approximately .

1.5.3.2. Jah~Teller systems The simplest and most elegant examples of the cooperative J a h n Teller effect

occur in the rare earth zircons as reviewed in detail by Gehring and Gehring (1975). In these systems, DyVO~, TmV04 and TbVO~ for example, the 4f electrons on the rare earth ions are well localized and have a strong spin-orbi t splitting. The lowest J mult iplet is then split by the crystal field which, in the h igh-symmetry phase of these materials, is a te t rahedral field. In the simplest case, TmV04, the ground state of the T m ion is a degenerate state whose degeneracy can be lifted by a distortion of the te t rahedral crystal field.

Since the 4f electrons are well localized this problem is part icularly simple because we can calculate the electronic response, Z E, for each rare ear th ion independently when, using eqn. (I.5.23), we obtain

ZE(qj) : _ [ W(qj)I2ZE.

In § 1.5.2.1 we calculated the susceptibility of a two-levels spin S = ½ system. We can use the results ibr TmV04 if we neglect the interactions between the different spins, when Z ~ =¼fi, where we have labelled the two electronic levels by the spin variable. The phase transit ion then occurs a t the tempera ture T c given by the m a x i m u m value of I W(qj)[2/(4l~B~o'(qj)2), which also determines the wave-vector and branch index of the soft mode.

In the rare ear th zircons the degeneracy of the electronic ground state is split by a shear of the surrounding tetrahedron, and the structure is such tha t the most energetically favourable mode is the acoustic shear mode determined by the elastic constant C11 -C12 . In other instances the soft mode is more complicated as for the Jahn-Te l le r distortion of the Cu ions in K2PbCu(N02)6 for which the unstable mode has an incommensurate wave-vector q~= (0"425,0"425,0) 2u/a (Yamada 1977).

Jahn-Te l le r phase transitions may also occur even if the electronic ground state of the high- temperature phase is not degenerate. Consider a two-level system with a

Page 101: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 101

splitting fl, while the distortion of crystal field causes transitions between these states. This is then an isomorphous problem with the tunnelling model discussed in §I.5.2.2, and the electronic susceptibility can be obtained from eqn. (I.5.12) with J = 0. The high- temperature phase is stable only while

tl coth fltl < W(qj)2 4~oB(Rj) 2"

which shows tha t there is a critical value of the splitting beyond which there is no ordering. This type of behaviour occurs in DyV04 and TbVO4 a l though the crystal field structure is more complex than the doublet discussed above (Gehring and Gehring 1975). In magnet i sm this behaviour is known as an induced momen t system (Grover 1965). This approach has been remarkably successful in explaining the properties of the rare ear th zircons. Although it is not possible to calculate the electron-phonon interactions W(qj) from first principles, a very successful theory has been developed with relatively few adjustable parameters and this theory gives a good account of a very large quant i ty of experimental data, as reviewed by Gehring and Gehring (1975).

More complications arise in the theory of the rare ear th pnictides because they are metals or semi-metals: the conduction electrons give rise to an interaction between the different rare ear th ions. Quali tat ively the theory can then be developed as above but using the full interacting susceptibilities of §§ 1.5.2.1 and 1.5.2.2 for the electronic susceptibility instead of the non-interacting susceptibilities used above, but it is then almost impossible to make calculations.

Finally this theory may be invalid for the 3d transit ion metal ions because in these systems there is a very strong coupling between the electronic levels and the surrounding ionic enviromnent so tha t keeping only the leading terms in the expansion (I.5.22) is invalid. I t is then possible to proceed by treat ing the interaction between an ion and its local environment using cluster methods, and coupling up the clusters using a mean field theory, as described by Thomas (1977). Of course difficulties then occur when the cluster around one 3d ion contains ions which are also par t of the cluster surrounding another rare earth ion.

1.5.3.3. Charqe density waves In this section we discuss the structural phase transitions associated with charge

density waves in metMlic systems, as elegantly reviewed in more detail by Friedel (1977). The electronic susceptibility is dependent upon (eqn. (I.5.28))

n(k2) - n(k + q2') (I.5.30) I ( q ) = - ~' v ( k 2 ) - v ( k + q ~ ' ) "

kZ.£'

This expression and hence X ° will become large if the electron energy denominators are small. In metMlie systems this occurs part icularly when the states (k + q2') and (k2) are both a t the Fermi level. Now, for any direction of q, there will be a limiting value of q = K F such tha t for q > K F both (k+q,U) and (k2) may be at the Fermi surface but not for q < K F or vice versa. We therefore expect a special contribution to l(q) when q is close to these wave-vectors. There are three cases of importance.

(a) I f the two Fermi surfaces have a tangent plane but different curvatures, as found in a [3d] free-electron system, the I(q) becomes

I(q) = K 1 + K:p log IPl, (I.5.31)

Page 102: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

102 R . A . Cowley

where K 1 and K2 are constants while q [I KF and p = [q - KF[. This shows tha t I (q) has an infinite derivat ive (Kohn 1959) when q = KF, and hence I(q) cannot be a m a x i m u m for these wave-vectors, in this instance a charge density wave will not occur with qs = KF.

(b) I f the two Fermi surfaces have a common tangent and one common curvature then

I(q) = K t + K2/(K 3 +/)1/2). (I.5.32)

This is the form of the expression obtained ibr a 12d[ free electron system and shows tha t there may be a m a x i m u m in I(q) when q = Kv. This suggests tha t there may be a charge density wave instabil i ty with qs = K r.

(c) I f the Fermi surfaces have a common tangent and two common curvatures, then

+ log IPl,

which is the divergent expression characteristic of a. [ld] electron system (Peierls 1955, Froehlich 1954). This suggests, neglecting electron-electron interactions, tha t a structural phase transit ion will always occur in this case, and the behaviour is known as an example of perfectly nesting Fermi surfaces.

These concepts originated with the work of Jones (1934) on the crystal s t ructure of alloys. In our cases (b) and (c), the lattice distortion, if it occurs, opens up a gap in the electron energy spectra a t the Fermi level. Consequently there are new Brillouin zone boundaries and a lowering of the total electron energy in the distorted phase.

Before discussing the application of these concepts in practice, it is necessary to mention the role of the electron-electron interactions. For simplicity we consider the case of nearly fl'ee electrons, when z~(q + , , - q - ~ ) becomes diagonal in the wave- vector, and fur thermore we neglect the effect of Umklapp processes. We can then write eqn. (I.5.25) as

EE(qj) = _ [ W ( q j ) ] 2 ](q) 1 + v E ( q ) l ( q ) '

which then leads to a condition ibr the stabil i ty e t a normal mode from eqn. (I.5.24) as (Chan and Heine 1973)

1 ]W(qj)[2 v ~ ( q ) < - - . (I.5.34) ~oB(qj) 2 I(q)

where vE(q) is the Fourier t ransform of the electrm~electron interaction and in the Har t ree approximat ion is positive and given by eZ/%q z. Equat ion (I.5.34) has the unexpected proper ty tha t in the absence of e lectron-phonon interaction, there would be no charge density instabilities of an electronic system because I(q) is necessarily positive (Friedel 1977). I t shows fur thermore tha t charge density wave instabilities occur even in [ld] systems only if[W(qj)[z/cgB(qj) 2 is sufficiently large. A completely convincing theory of a s t ructural phase transit ion can be carried out only if there is a careful calculation of the electron phonon interaction W(qj), the susceptibility I(q) and the electron-electron interactions vE(q).

In practice this p rogramme has not been carried out to completion for any system. Many of the calculations have concentrated on calculating the core of the non-interacting electronic susceptibility, I(q), by using simple models and more

Page 103: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 103

recently by using the results of detailed band structure calculations. In the case of the [ld] conductors such as KCP(K2Pt(CN)413r0.33H20) (Comgs et al. 1973 a) and TTF-TCNQ (Comgs et al. 1973 b) the distorted structure has a wave-vector K F along the chain direction which is exactly consistent with the expected value of K F from the known electron density. Consequently there is no doubt tha t the distorted structure results from the peak in the electronic susceptibility for q = N F predicted by eqn. (I.5.33). Electron-electron interactions play a role however not only for the interactions along the chains, bu t also in giving rise to forces between the chains (Barisie 1971) without which there would not be a transition to a [3d] ordered structure. Clearly until the theory of the interacting [ld] electron gas is more perfectly understood we cannot hope to have a completely satisfactory theory of these materiMs.

The layered transition metal diehaleogenides have distorted phases described by incommensurate wave-vectol~. These wetv initially assumed to arise from cylin- drical parts of the Fermi surface associated with the layered structure, giving rise to a corresponding peak in I(q). This model, however, fails to explain why different examples of the series have very similar periodicity for the charge density waves; NbSe2 and TaSe2 both have incommensurate phases with q s = ( 1 - 5 ) a * / 3 with 5~0-02. Fur thermore eMeulations by Myron et al. (1977) of I(q) using a self- consistent muffin tin potential for TaSe2 show only a vel3r weak peak for q ~qs, and similar calculations for TaS2 give a peak at larger wave-vectors than q~ (fig. 1.51). Clearly the nature of the eleetron-phonon interaction, and of the electron-electron interactions in these materials is of ei'ueial importance.

In the A15 compounds, NbaSn and VaSi, the situation is also confused. The first model proposed to explain the transition was put forward by Labbe and Friedel (1966). Essentially their model describes the shear as arising from an inequivalenee of the electron energy bands in the different directions of the distorted etwstal. There is then a redistribution of the electrons and a consequent lowering of the energy. This effect is included in the calculation of I(q) in eqn. (I.5.33), but is only large enough to give an instability if the density of states at the Fermi surface is large. Labbe and Friedel obtained a large enough density of states by postulating [ld]:like bands arising from the chains of metal atoms in the A15 structure. The symmetry of the A 15 structure also causes a degeneracy of the electron energy bands at the [100] zone boundary. If the crystal distorts, this degeneracy is lifted and Gorkov (1973) suggested that if the Fermi surface was nearly [ld] and very close to the [100] zone boundary this splitting would account for the transition. Unfortunately Mattheiss's calculations (1965, 1975) of the band structure failed to substantiate these simplified models of the band structure and in more recent theories, such as those of Bha t t (1977), it is difficult to obtain a simple physical picture.

Calculations on other systems have been less ambitious.

1.5.3.4. Other charge density waves Besides the charge density waves associated with the Fermi surface effects

described above, there are a number of structural phase transitions associated with other types of electronic instability. In this section, we only briefly indicate some of these transitions.

The first example is the charge ordering which takes place in F%O4. In the high- temperature phase the iron ions have two sites; a tetrahedral site which is occupied

Page 104: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

104 g . A . Cowley

Fig. 1.51

7 . 0 1 ' I q I ' I ~ ] ' ] ' ] ' I ' ] ~ I ' / ' [ ~ I

~ 6 . 5 z

5.5

0

5 . C -

4.~ 0

0

I 5 5 P

IT TaS 2 O O 0 00000

0 0000

0

0 000 -

0

0

( a )

, I ~ I ~ ] t ] i ] i I I I ~ I I 1 - 9 II 13 15 17 19 21 25 25

q

9 . 0

8.5

:E 8 O o

7 5

Q:

u~ 7 0

~ 6 5 G

6 O

j , i E I ,

_ IT To Se 2

o o '%, - ocd~Ct~oO °o -

- ooooo °~°% o -

- - 0

o o O - -

- c o 0

(b) _

F I 3 5 7 9 I] 13 15 17 19 21 25 25 M q

/(q) =z(q) for TaSe 2 and TaS2 for q along the TM direction. Whereas the curve for TaSe~_ shows a weak peak at the incommensurate wave-vector, that of T~S2 does not. (After Myron et al. 1977.)

by an Fe 3 + ion and two octahedral sites occupied by Fe 2 + and Fe 3 + ions. In the high- temperature phase the distribution of Fe 2 + and Fe 3 + ions is random but below the Verwey transition at 123 K the Fe 2 + and Fe 3 + ions take up an ordered arrangement (Verwey and H a a y m a n 1941, Hamilton 1958). In both phases the electrons are localized but the hopping conductivity is drastically reduced at the phase transition. Since this phase transition is clearly associated with the charge distribution, there is a corresponding structural distortion. Both of these effects have been detei~fined using neutron diffraction techniques as reviewed by Shirane (1977).

This transition is qualitatively not very different from a metal-insulator transition (Mort and Zinamon 1970). Basically in a system with one electron/atom, the correlation-energy favours a situation in which each electron is situated on a different a tom in an antiferromagnetic configuration. I f the overlap wigh the neighbouring ions is sufficiently small there will be a gap between this insulating state and one in which the electrons can propagate through the crystal. This possible change in the electron wavefunctions causes a change in the electron density and hence may give rise to a structural phase transition.

Another example of where a change in the electronic properties gives rise to a structural phase transition is in the fluctuating valence systems, such as SINS,

Page 105: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Struettural phase transitions 105

Sm/YS or Ce/Th alloys. In these systems (Varma 1976) the atomic 4flevels are close in energy to the 5d-6s bands and as composition, t empera ture or pressure are varied, the electrons may take up either the localized 4f states or the band states. In SmS alloyed with YS, for example, a t 19(~o of YS there is a first-order transition from a semiconducting phase with localized 4f electrons to a metallic phase of mixed 4 f and 5d-6s states (Tao and Hol tberg 1975). This transition is associated with a 12(yo decrease in volume. As discussed in §I.4.1.5~ a bulk modulus instabili ty has no critical fluctuations and is necessarily of first order due to cubic terms in the free energy. These conclusions are in excellent accord with the available experimental results for the acoustic modes in the Sm/YS system by Penney et al. (1976).

Finally we discuss magnet ic transitions. I f e l e e t ro~phonon interactions are neglected then eqn. (I.5.34) suggests t ha t electronic systems will only show phase transitions if rE(q)<0. This is indeed the case for magnetic transit ions because exchange interactions are a t t ract ive, a n d eqn. (I.5.34) with W(qj )= 0 is the well- known Stoner (1946) criterion for i t inerant electron magnetism. The coupling to the phonons requires, however, more discussion. In the absence of spin-orbi t interaction W(qj) = 0 and there is no linear coupling between the spin density and the phonon modes. The phonon modes play the role of secondary order parameters. Inclusion of the spin-orbi t coupling allows a non-zero interaction, but this is frequently very much weaker than the corresponding coupling to charge density waves.

1.5.3.5. Dynamic effects and temperature-dependent effects In § 1.5.3.1 the expression for the electronic response was evaluated assuming

tha t a static field was set up by the normal modes. In principle, it is comparat ively easy to generalize this by expressing the results in terms of the dynamic electronic susceptibility, which is given by the same expression as eqn. (I.5.28) but with the denominator replaced by v ( k 2 ) - v ( k + q 2 ' ) + o ) . This suggests tha t the dynamic response will only be significantly different from the static response if the frequencies co of the normal modes are comparable with the difference in energy of the two electron states. In most instances nearly all the electronic energy denominators are comparable with the Fermi energy and consequently much larger than he) except for a very few electronic transitions between electron states close to the Fermi surface. In these cases there will be veiT7 little damping of the normal modes by the electronic excitations, and very little difference between the static and dynamic sus- ceptibilities. There are two instances where this may not be the ease. First ly in [ld] conductors, KCP and TTF-TCNQ, the density of states a t the Fermi surface is large and phonon damping may result as possibly illustrated in fig. 1.40. Secondly in the insulating Jahn-Te l le r systems discussed in § 1.5.3.2, the energies of the electronic excitations are relatively small and well defined (Gehring and Gehring 1975) so tha t coupled mode spectra are observed similar to the behaviour discussed in § 1.4.1.6.

Finally we discuss the tempera ture dependence of the phonon frequencies due to the electronic susceptibility. In many theories such as those of the A15 materials and of the transi t ion-metal dichaleogenides, a t t empts have been made to explain the tempera ture dependence of the normal modes as arising from the change in the Fermi factors in )/~(q) (eqn. (I.5.28)). This would appear to be unreasonable except in [ld] conductors and insulating J a h n Teller systems because the Fermi tempera ture is much higher than the transit ion temperature . I t is surely more reasonable to explain the tempera ture dependence as arising from the anharmonici ty described in § 1.5.1 and to make use of the electronic terms only to explain the unstable zero tempera ture 'harmonic ' frequencies of the normal modes.

Page 106: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

106 R . A . Cowley

This combined approach has now been shown to be the appropriate one in describing the temperature and dopant concentration dependence of the soft mode and phase transition in doped SnTe. Pure SnTe has the NaC1 crystal structure at high temperature but below about 120K has a phase transition to a ferroeleetric rhombohedral phase (Kobayashi et al. 1976, Pawley et al. 1966). This phase transition temperature is decreased as the carrier concentration is increased and reaches 0 K for a concentration of 8 x 1026 m 3 The soft mode has been studied below Tc for several samples by Sugai et al. (1977 b), fig. 1.52. The first theory of the concentration dependence by Kristoffel and Konsin (1968) a t tempted to explain the results entirely in terms of the temperature dependence of the electronic response of the free carriers, but more recently Sugai et al. (1977 a) found it necessary to include anharmonicity as well. This is basically because the Raman spectra, fig. 1.52, show that the frequency of the TO mode has much the same temperature dependence, irrespective of the concentration, and arises from the anharmonieity, while the frequency at zero temperature is carrier dependent. Finally we comment tha t phonon interaetions were also found to be of crucial importance in explaining the observed temperature dependence in KCP (Sham and Pat ten 1976).

Fig. 1.52

L 500 "~ I P cm-3 ; ~1 2 2x1020

, , , m

o3oo1- \ \ o 2o ~m-'

z 100

,~, 0 50 100 K

TEMPERATURE

The temperature dependence of the TO frequency squared in doped 8nTe. (After 8ugai et al. 1977 m)

1.5.4. Conclus ions

The basic conclusions of this section have to be that the microscopic theory of structurM phase transitions is in an unsatisfactory state. In part this reflects the very success of the Landau/soft mode theory which in terms of a few arbi t rary parameters is able to correlate a very large amount of experimentM data. I t is then far too easy to invent a ~microscopic model' with a similar number of parameters which not surprisingly also correlates a large amount of experimental data. A microscopic model is only justified if either it predicts the parameters of a Landau theory To, a, (qsJ), etc., without introducing new parameters, or else also predicts the results of a large quant i ty of other microscopic data such as phonon-dispersion relations or optical reflectivity of a metal.

Page 107: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural pha~e transitions 107

On these strict tests we have made very little progress. Except for the [ld] metals and, less c(mvbwingly, some order~tisorder systems we can barely predict the wave- vector q~ with any cotdidence let alone the transition temperature, T¢. We hope that this review, by cataloguing our inadequacies, will stimulate further work to develop more sat isfacto~ microscopic theories, in the hope not just of improving our understanding, but also that we may make better use of the important technological properties of materials associated with structural transitions.

REFERENCES ALEFELD, B., BIRI% M., and HEIDEMANN, A., 1969, Naturwissenschaften, 8, 410. AL,~-NIELSEN, J., and DIETRICH, 0. W., 1967, Phys. Rev., 153, 706, 711,717. ANI)I~RSON, P. W., 1960, Fizika Dielectri~ov, edited by G. I. Skanavi (Moscow: Acad. Nauk.

ss~t) AUBRY, S., 1977, Proceedings of the Conference on Stochastic Behaviour in Classical and

Quantum Systems, Como, June 20-24. BAK, P., 1978, Proceedings of the Symposium on Soliton Strvcture and Dynamics in Condensed

Matter, Oxford, 22-29 June. BAK. P., and E.~IERV. V. J., 1976, Phys. Rev. Lett.. 36, 978. BAN, P.. and TI.~tONE~. ,J.. 1978. J. Phys. ('. 11, 4901. ]3ALLENTYNE, J. M., 1964, Phys. Rev. A, 429, 136. BABN(', S., 1971, Solid St. Commun., 9, 1507. BARNOSKI, M. S., ~nd BALLENT~/'NE, J. M., 1968, Phys. Rev., 174, 946. BnCHTLE, D. W., ~nd SCOTT, J. F., 1977, J. Phys. C, 10, L209. BHATT, R. N.~ 1977, Phys. Rev. B, 16, 1915. BUIDn, V. G., and HEGDE, M. S., 1972, Phys. Rev. B, 5, 3488. BLI~C, I~., 1960, J. Phys. Chem. Solids, 13, 204; 1971, Structural Phase Transitions and Soft

Modes, edited by E. J. Samuelson, E. Anderson ~nd J. Feder, (Oslo: Universitetsforl~gel), pp. 97-132.

BLanc, R., and MALL M., 1969, Solid St. Commun., 7, 1413. BLINC, R., and ZEKS, B., 1974, Soft Modes in Ferroelectrics and Antifer~velectrics (New York:

Elsevier). BOR~, M., ~nd HUANG, K., 1954, Dynamical Theory of Crystal Lattices (Ox~ford University

Press). BOKSA, F., 1971, Structural Phase Transitions and Soft Modes, edited by E. J. Samuelson, E.

Anderson ~nd J. Feder, (Oslo: Universitetsforl~get), p. 371. BORSA, F., and ]:~IGAMONTI, A., 1972, Physics Lett. A, 40, 399. BRODY, E. M., and CUMMINS, H. Z., 1968, Phys. Rev. Lett., 21, 1263. BROUT, R., MULLER, K. A., und T~OMAS, H., 1966, Solid St. Commun., 4, 507. BRUCE, A. D., and AHARO~¥, A., 1975, Phys. Rev. B, 11,478. BRUCE, A. D., and COWLEY, R. A., 1973, J. Phys. C, 6, 2422; 1978, Ibid., 11, 3609. BRUCE, A. D., MURRAY, A. F., and COWLEY, R. A., 1978, J. Phys. C, 11, 3591. CARNE~RO, K., SHIRANE, G., WERNER, S. A., and KAISER, S., 1976, Phys. Rev., 13, 4258. Cnh~, S. K., and HnT~E, V., 1973, J. Phys. F, 3, 795. C()Cu~¢A~, W., 1960, Adv. Phys., 9,387; 1969, Ibid., 18,157; 1971, Structural Phase Transitions

and Soft Modes, edited by E. J. Samuelson, E. Anderson and J. Feder, (Oslo: Universitetforl~get), p. l ; 1973: The Dynamics of Atoms in Crystals (London: Edward Arnold).

COLLI]:~S, M. F., and Tnn, H. C., 1973, Phys. Rev. Lett., 30, 781. CoM~s, P~., LAMBERT~ M., ~nd GuildEr, A., 1970, Acta crystallogr. A, 26, 244. COMICS, R., LAMBEI~T, M., LAU~OIS, H., und ZELLER, H. R., 1973a, Phys. Rev. B, 8, 571. CoM~s, R., LAMBERT, M., and ZELLER, H. l~., 1973 b, Phys. Stat. Sol. (b), 58, 587. COMICS, R., SHAPIRO, S. M., SJ~IRANE, G., GARITO, A. F., and I-IEEGER, A. J., 1976, Phys. Rev.

B, 14, 2376. COWLEY, R. A., 1963, Adv. Phys., 12,421; 1964, Phil. Mag., 11,673; 1968, Rep. Prog. Phys.,

31, 123; 1976a, Phys. Rev. B, 13, 4877; 1976b, Phys. Rev. Lett., 36, 744. Co\vLi~:v, R. A., Axe, J. D., and hzu~m, M., 1976, Phys. Rev. Lett., 36, 806.

Page 108: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

10S R . A . Cowley

COWLEY, I~. A., and :BRUCE, A. D., 1978, J. Phys. C, 11, 3577. COWLEY, I~. A., BUYEP~S, W. J, L., and DOLLING, G., 1969, Solid St. Comm,wt~.. 7, 181. Ct~M~INS, S. E., 1970, Ferroelectrics, l, ll. DARL~SGTON, C. N. W., and O'CONNOR, D. A., 1976, d. Phys., C, 9, 3561. DEVONSglRE, A. F., 1949, Phil. Maq., 40, 1040; 1954, Adv. Phys., 3, 85. DORNER, B., AXE, J. D., and StoRAgE, G., 1972, Phys. Rev. B, 6, 1950. DORNER, B., BAUER, K. H. W., and GRIMM, H., 1974, FerroeIectries, 7, 291. DVORAK, V., 1974, Ferroelectrics, 7, 1. EBERHA~D, J. W., and HORN, P. M., 1975, Solid St. Commun., 16, 1343. FLEURY, P. A., SCOTT, J. F., and WO~LOOK, J. M., 1968, Phys. Rev. Lett., 21, 16. FLEURY, P. A., and WORLOC~, J. M., 1968, Phys. Rev., 174, 613. FRANKE, V., and HEGENBARTH, E., 1974, Phys. Stat. Sol., 25, K17. FgIEDEL, J., 1977, Elect~vvn o/nd Phonon Interactions andPhase Transitions, edited by T. giste

(New York: Plenum). FROEHLICH, H., 1954, Pr.oc. R. Soc. A, 233, 296. GEHRING, G. A., and GEHRING, K. A., 1975, Rep. Prog. Phys., 38, 1. GLAUBER, P~. J., 1963, J. Math. Phys., 4, 294. GORKOV, L. P., 1973, Soviet Phys. JETP Lett., 17, 379. GaOVER, B., 1965, Phys. Rev. A, 140, 1944. HAMILTON, W. C., 1958, Phys. Rev., 110, 1050. HARADA, J., PEDERSEN, T., and :BARNEA, Z., 1970, Acta crystallogr. A, 26, 336. HARLEY, R. T., and MACFARLANE, P~. ~i,, 1975, J. Phys. C, 8, [,451. HATTA, I., 1968, J. phys. Soc. Japan, 24, 1043. HORNREICH, I~. M., LUEAN, M., and SHTRIKMAN, S., 1975, Phys. Rev. Lett., 35, 1678. HORTON, G. K., and MARADUDIN, A. A., 1974, Dynamical P'~vperties of Solids (New York:

Elsevier). HosmNo, S., and MOTE¢I, H., 1967, dap. J. appl. Phys., 6, 708. HiILLER, A., 1972, Z. Phys., 254, 456. HtTLLER, A., and PRESS, W., 1972, Phys. Rev. Lett., 29, 266. HUTCmNGS, M. T., 1964, Solid State Physics 16 (New York: Academic Press), p. 227. hZUMI, M., AXE, J. D., SHIRANE, G., and SmMAO~:A, K., 1977, Phys. Rev. B, 15, 4392. hzvMI, M., and GESI, K., 1977, Solid St. Commun., 22, 37. INGLESFELD, J. E., 1972, J. Phys. F, 2, 68. JAMES, H. M., and KEEI~AN, T. A., 1959, J. chem. Phys., 31, 12. JOHNSON, C. J., 1965, Appl. Phys. Lett., 7, 221. JONA, F., and S m ~ E , G., 1962, Ferroelectric Crystals (New York: Pergnmon). JONES, H., 1934, Proc. R. Soc., 144, 225. KAMINOW, I. P., and DAMEN, T. C., 1968, Phys. Rev. Lett., 20, 1105. KANAMORI, J., 1960, J. appl. Phys., 31, 145. KA~ZlG, W., 1951, Helv. phys. Acta, 24, 175. KAY, M. I., 1972, Ferroelectrics, 4, 235. KJEMS, J. K., and DOLHNG, G., 1975, Phys. Rev. B, 11, 1639. KOBAYASHI, K. K., 1968, J. phys. Soc. Japan, 24, 497. KOBAYASHI, K. L. [., KATO, Y., KATAYAMA, Y., and KOMATSUBARA, K. F., 1976, Phys. Rev.

Lett., 37, 772. KOHN, W., 1959, Phys. Rev. Lett., 2, 393. KRISTOWEL, N. and Ko~slS, P., 1968. Phys. Stal. Sol., 28,731. KWOK, P. C.. and MH~LER, P. B., 1966, Phys. Rev., 151,387. LAEB~, J., and FRIEDEL, J., 1966, J. Phys., Paris, 27, 153,303, 708. LANDAU, L. D., 1937, Phys. Z. SowjUn, 11, 26. LANDAU, L. D., and LIESHITZ, E. M., 1959, Statistical Physics (London: Pergamon). LI~ES, M. E., and GLASS, A. M., 1977, Principles and Applications" of Ferroelectrics and Related

Materials (Oxford University Press). Li)THI, B., and MORAN, T. J., 1970, Phys. Rev., 2, 1211. McKENZlE, D. R., 1975, J. Phys. C: 8, 1607. McMILLA~, W. L., 1976, Phys. Rev. B, 14, 1496. MARADUDIN, A. A., and FEIN, A. E., 1962, Phys. Rev., 128, 2589. MARADVmN, A. D., and FLYN~, P. A., 1963 a, Phys. Rev., 129, 2529; 1963 b, Ann. Phys., 22,

223.

Page 109: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

Structural phase transitions 109

~/[ATTHEISS, L. F., 1965, Phys. Rev. A, 138, 112: 1975, Phys. Rev. B, 12, 2161. MEaAW, H. D., 1947, Proc. R. Soc., 189, 261. MERZ, W. J., 1953, Phys. Rev., 91,513. MEYER, G. M.~ and COWLEY, R. A.~ 1976, Ferroelectrics, 11,479. MEZEt, F., 1972, Z. Phys., 255, 146. M1GONI, R. BILZ, H., and BAi~ERLE. D., 1976, Phys. Rev. Lett., 37, 1155. MONCTON, D. E., AXE, J. D., and Dt SALVO, F. J., 1975, Phys. Rev. Lett., 34, 734. MOTT, N. F.~ and ZINAMON, Z., 1970, Rep. Pwg. Phys., 33, 88l. M[~[,LER, K. A., 197 l, Structural Phase Transitions and Soft Modes, edited by E. J. Samuelson,

E. Anderson and J. Forbes, (Oslo: Universitetsforlaget), pp. 61-96. M/~LLER, K. A., and BERLINGER, W., 1971, Phys. Rev. Lett., 26, 13. Mi)LLER, K. A., BnnI,INGER, W., and SLONCZEWSKL J. C., 1970, Phys. Rev. Lett., 25, 734. Mi)LLER, K. A., BERLINGER, W., and WALDNER, F., 1968, Phys. Rev. Lett., 21,814. MYRON, H. W., RATH, J., and FREEMAN, A. J., 1977, Phys. Rev., B, 15, 885. ~AGAMIYA, T., 1943, P'~vc. phys. math. Soc. Japan, 24, 137; 1944, Ibid., 25, 540. NAKANISHL S., and SHIBA, H., 1977, J. phys. Soc. Japan, 43, 1839. NAKANISHI, K., TAKATERA, H., YAMADA, Y., and SHIEA, H., 1977, J. phys. Soc. Japan, 43,

1509. NELMES, R. J., E[RIKSSON, V. R., and ROUSE, R. D., 1972, Solid St. Commun., 11,261. NELMES, R. J., KENNEDY, N. S. J., BAHAR[E, E., and HEWAT, A. W., 1978, Lattice Dynamics,

edited by M. Balkanski (Paris: Flammarion), p. 705. 0RIANI, R. A., and Mum'~¥, W. K., 1958, J. Phys. Chem. Solids, 6, 277. OVERHAUSER, A. W., 1971, Phys. Rev. B, 10, 3173. PAUL, G. L., COCHRAN, W., BUYERS, W. J. L., and COWLEY, R. A., 1970, Phys. Rev. B, 2, 4603. PAWLEY, G. S., COCHRAN, W., COWLEY, R. A., and DOLLING, G., 1966, Phys. Rev. Lett., 17,

753. PEERCY, P. S., 1973, Phys. Rev. Lett., 31,379. PEIERLS, 1% E., 1955, Qnantnm Theory of Solids (Oxford University Press). PENNEY. T., MELCHER, R. L., HOLTZBERG, F. and GUNTHERODT, G., 1976, A.I.P. Conference

Proceedings, 29, 392. PRESS, W., 1972, J. chem. Phys.. 56, 2597. PY.~N, t~., 1978, J. Phys. F. 8, 1. PYN~-, R., and AXE, J. D., 1976, J. Phys. C, 9, L199. RAICH, J. C., and GILLIS, N. S., 1977, J. chem. Phys., 66, 846. RAMAN, C. V., and NEDUNGADI, T. M. K., 1940, Nature, Lond., 145, 147. REHWALD, W., 1973, Adv. Phys., 22, 721. RISTE, T., SAMUELSON, E. J., arid OTNES, K., 1971, Structural Phase Transitions and Soft

Modes, edited by E. J. Samuelson, E. Ander'son and J. Feder, (Oslo: Universitetsforlaget), pp. 395-408.

SAKSENA, B. D., 1940. P~w'. ~atn. Acad. ~ci. India. A. 12, 93. SAK('I~.~I. J.. ('owIA,:v, R. A., sod DOHAN(;, G., 1970, J. phys. So~'. Japan, 28, 1426. SAMARA, a. A., 1971, Phys. Rev. Lett., 27, 103. SAMARA, G. A., SAKUDO, T., and YOSHIMITSU, K., 1975, Phys. Rev. Lett., 35, 16. SAWADA, A., UDAGAWA, M., and NAKAMURA, T., 1977, Phys. Rev. Lett., 39, 829. SCOTT, J. F., 1974, Rev. rood. Phys., 46, 83. SCOTT, A. C., Cttu, F. Y. U., and MoLAUOHLIN, D. W., 1973, Proc. Inst. electron. Engrs, 61,

1443. SHAM, L. J., 1974, Dynamical Properties of Solids, Vol. 1, edited by G. K. Horton and A. A.

Maradudin (New York: Elsevier), p. 301. SHAM, L. J., and PATTON, B. R., 1976, Phys. Rev. Lett., 36., 733; 197(i, Phys. Rev. B, 13, 315. SHAPIRO, S. M., COWLEY, R. A., Cox, D. E., and EIBSCHULTZ, M., 1976, Oak Ridge Laboratory

Report Conf~760601~P1, p. 399. SHENOY, G. K., 1973, P~vspec.tives in MSssbauer Spectroscopy (New York: Plenum Press), p.

141. SnIBA, H., and ISHIBASHI, Y., 1978, J. phys. Soc. Japan, 44, 1592. SmRANE, G., 1974, Rev. Mod. Phys., 46, 437; 1977, Electron Phonon Interaction and Phase

Transitions, edited by T. Riste (:New York: Plenum). SHIRANE, G., JONA, F., and PEPINSKY, R., 1955, Proc. Inst. Radio Engrs, 43, 1738.

Page 110: Structural phase transitions I. Landau theory IN PHYSICS, 1980, VOL. 29, NO. 1, 1-110 Structural phase transitions I. Landau theory By l~. A. COWLEY Department of Physics, University

110 Structural phase transitions

SHIUA~E, G., and YAMADA, Y., 1969, Phys. Rev., 177, 858. •ILVERMAN, B. D., and JOSEPH, R. I., 1963, Phys. Rev., 129, 2062; 1964, Ibid., 133, A207. SKALYO, J., FRAZER, B. C., and SI4IRA~E, G., 1970, Phys. Rev. B, l, 278. SLONCZEWSKI, J. C., and THOMAS, H., 1970, Phys. Rev. B, 1, 3599. SPITZER, W. G., MILLER, l~. C., KLEINMAN, D. A., and HOWARTH, L. E., 1962, Phys. Rev., 120,

17t0. STANLEY, H. E., 1971, lnt,~vduetion to Phase Transitions and Critical Phenomena (Oxford

University Press). STIRLI~O, W. G., 1972, J. Phys. C, 5, 27tl. STONER, E. C., 1946, Rep. Prog. Phys., l l , 43. SUGAI, S., MURASE, K., KATAYAMA, S., TAKAOKA, S.~ NISHI, S. and KAWAMURA, H., 1977 a,

Solid St. Commun., 24, 407. SUGAI, S., MURASE, K., and KAWAMURA, H., 1977 b, Solid St. Commun., 23, 127. SUGIIIARA, K., 1959, J. phys. Soe. Japan, 14, 1231. SuzuKI, M., and KuBo, R., 1968, J. phys. Soc. Japan, 24, 51. TAO, L. J., and HOLTZBERG, F.~ 1975, Phys. Rev. B, 11, 3842. TERAUCm. H., NODA, Y., and YAMAI)A~ Y., 1972, J. phys. Soe. Japan, 32, 1560. THOMAS, H., 1977, EleetronPhonon interactions and Phase Transitions, edited by T. Riste

(New York: Plenum). TOKUNAGA~ M., and MATSUBARA, T., 1966, P~vg. theor. Phys., Kyoto, 35, 581. UNOKI, H., and SAKUDO, T., 1967, J. phys. Soc. Japan, 23, 546. VAKS, V. G., LARKI~, A. I., and PIKIN, S. A., 1967, Zh. ~ksp. teor. Fiz., 53, 1089 (English

translation: 1968 Soviet Phys. JETP, 26, 647). VAKS, U. G., and SCttNEIDER, J. E., 1976, Phys. Stat. Sot. (a), 35, 61. VAN HOVE, L., 1954, Phys. Rev., 95,249. VAR~IA, C. M., 1976, Rev. rood. Phys., 48, 219. VERWEY, E. J. W., and HAAYMAN, P. W., 1941, Physics, 8, 977. WADA, M., SAWADA~ A., iSHIBASHI, Y., and TAKAGI, Y., 1977, J. phys. Soc. Japan, 42, 1229. WAGNER, H., 166, Z. Phys., 195, 273. WHITE, K. I., TAYLOR, W., KATIYAR, R. S., and KAY, S. M., 1970, Physics Lett. A, 33, 175. WORLOCK, J. M., and OLSON, D. H., 1971, Proceedings of the International Conference on Light

Scattering (Paris: Flammarion). p. 410. YA~IADA, Y., 1977, Electron-Phonon Interactions and Phase Transitions, edited by T. Riste

(New York: Plenum). YAMADA, Y.: FUJII, Y., and HATTA, I., 1968, J. phys. Soc. Japan, 24, 1053. YAMADA,, Y., MORI. M., and NODA, Y., 1972, J. phys. Soe. Japan, 32, 1565. YAMADA, Y., and YA~IADA, T., 1966, J. phys. So('. Japan, 21, 2167.