Spin Glasses - An Experimental Introduction
Spin Glasses - An Experimental Introduction
an experimental
introduction
Schrtig in dem eisernen Gitter
der drehende Spin
am Drehsinn
errtitst du die Seele
J. A. Mydosh
1993
UK Taylor & Francis Ltd, 4 John St., London WClN 2ET
USA Taylor & Francis Inc., 1900 Frost Road, Suite 101, Bristol PA
19007
0 J. A. Mydosh 1993
All rights reserved. No part of this publication may be reproduced,
stored in a retrieval system, or transmitted, in any form or by any
means, electronic, electrostatic, magnetic tape, mechanical,
photocopying, recording or otherwise, without the prior permission
of the copyright owner and publishers.
. ..
PREFACE Vlll
INDEX 251
Preface
Phenomena that exist in nature should not only be observed and classified
but describable by models or theories which then lead to a better
understanding and predictability of the phenomenon. Often, as is the case
with the spin glasses, experiment has serendipitously uncovered some
unusual or unexpected effect which at first glance baffles the anticipations
of our physical intuition. As a result of the novel experimental properties,
the phenomenon requires further study and a preliminary picture or concept
to represent it. Here a name should be given. With the accumulation of
data mathematical notions are needed to quantify the results and an incipient
theory is born. The phenomenon, its various experimental manifestations,
the many (hopefully) specific examples which can be found, and the
embracing theory, all developing over the course of many years, create a
problem area, in our case the ‘spin-glass problem’. Fresh input comes from
the drawing of various analogies with other areas, new anomalies are
discovered by more sophisticated experiments, and theoretical progress
leads to a deeper understanding and predictive power. Hence the problem
is nicely on its way to being solved.
Frequently there is a long pause after an important discovery and the
recognition of its significance. Time and ideas are necessary to set the stage
for further advancement and the delay can be great. With some luck the
given phenomenon will become a major effect, i.e., an important topic of
contemporary research. Hectic efforts, measured by the large number of
workers presenting their results at countless conferences and workshops,
are demanded. And all this could continue for many years before a (near)
solution is found or the ideas and curiosity run out. Thus, for a particular
period the area becomes highly fashionable or ‘in-physics’, and then after
its fame and topicality diminish, the problem, if still unsolved, may remain
as an interesting, yet pmsC research topic. Nevertheless during these intense
years, a great deal was learned and this knowledge will remain as an
integral part of physics and even appear in tomorrow’s text books.
So it was and is with the spin glasses. Around the late 1960s there were
some unusual effects observed on a series of magnetic alloys (rather simple
and traditional ones such as iron in gold and manganese in copper).
X Preface
However, then the time was not right but at least a name was coined. Only
in the mid 1970s did theory using some radically new ideas come to grips
with the experiments and the possibility of a sharp phase transition. Around
1975 ‘all hell broke loose’ and this began the fame period of the spin glasses
with great activity and much progress. Here analogies and spin-offs increased
the pace. After 1985 the tempo slowed down and a third phase of tranquility
appeared with more sophisticated theories, subtle new experiments and
‘ideal’ spin-glass materials. The phenomenon had run its course to
establishment so that now is perhaps a good time to try and tell its story,
naturally in a phenomenological way.
The purpose of this book is indicated in the title. It is meant to introduce
the phenomenon and physics of spin glass (Chapters 1 and 2) via an
experimental approach (Chapter 3). The method is to use simple physical
pictures and concepts based on the empirical observations, rather than
employing theoretical models with the associated mathematics. Hence,
figures and tables will be more common than equations and derivations.
Materials, i.e., the real spin glasses, will also play a major role (Chapter
4). In spite of this particular emphasis the theory of spin glasses is an
essential part of the story and its ideas, techniques and advances will be
portrayed in an elementary fashion (Chapter 5). How close does experiment
come to theory? Chapter 6 considers the recent advances in finding and
measuring spin glasses which closely mimic the theoretical models.
Why write such a book on spin glasses? At least four good reasons can
be given.
1 The spin glass has become a fundamental and general form of magnetism;
its general occurrence as ‘magnetic ordering’ occupies the third place
after ferromagnetism and antiferromagnetism.
2 Randomness, frustration, glassiness and amorphousness represent very
important phenomena in contemporary physics.
3 Radically new concepts and ideas are needed to describe the basic
experimental behaviours.
4 There is a richness of analogies with many other areas extending from
astrophysics via molecular evolution to zoology. A few of these analogies
are examined in Chapter 7.
The spin glasses were surely one of the first precursors to the physics of
complexity - a new area which has recently been established.
For the above reasons a simple, less theoretical treatment of the problem
should be made available for the non-theorist or even non-physicists, and
also the younger (graduate) students in order to serve as an introduction
to the subject and its basic physics. Thus, the level of the book is low; only
a general background in solid-state physics and elementary magnetism is
required beyond the undergraduate courses, for the remaining necessary
concepts are explained, many in Chapter 2 and others as needed. Emphasis
Preface Xl
is placed on experimental facts and physical notions with always a real spin-
glass system in mind. The powerful developments in the theory - the new
statistical mechanics - will be sketched and compared to experiment but
not discussed in mathematic details or via long derivations. Concerning the
theory of spin glasses there are a number of texts and review articles already
available here and it would be unwise and unnecessary to compete with
these excellent theoretical exposes. What is missing is the elementary level,
rudimentary explanation of the spin-glass problem for just about everyone.
More so the novice or newcomer to the field and the expert from another
area are the welcome readers. Perhaps the real spin-glass connoisseurs
would value the book as a reference to the salient experiment features and
an indication of the theoretical lines and directions of progress.
The author has been too long occupied with the phenomenon of spin
glasses. First there were the quiet years of experimentation, then the hectic
decade when it seemed like everyone was working on the problem, and
finally, now that the dust has settled a bit, a chance to look back at the
past 25 or so years and see what knowledge has been gained and what we
really understand. Although my research emphasis has changed to other
topics and other problems, nevertheless, my ‘first love’ remains the spin
glasses. Before forgetting everything it’s time to pause for a moment and
in retrospection write this book as quickly as possible. The season is
certainly ripe, the need is great and a new generation of students is moving
into this research area. But such an undertaking is not meant as the final
word; the field is not finished. The spin-glass problem has not been fully
solved, there is still movement, albeit slow, towards a more complete
theoretical description with yet newer ideas. Also experimentation is
continuing with novel forms of spin glasses and even more subtle effects.
Hopefully then a more finalized version of this spin-glass text will appear
in the next ten years. Thus my apologies now for presenting the yet
incomplete and modifiable ‘truth’.
Although the main ideas and threads of thought for this book were
conceived at the Kamerlingh Onnes Laboratory in Leiden, most of writing
took place at the University of California in Santa Barbara during a
sabbatical visit of six months. I must thank V. Jaccarino for his kind
hospitality at UCSB. Also I would like to acknowledge the support and
encouragement of M. H. Vente during the rather long period of gestation,
writing and correcting. M. Nieuwenhuijzen-Verschoor did a superb job in
typing the manuscript. There were simply too many discussions and
interactions some of them long-forgotten during the past 25 years which
moulded my views on the spin glasses. I should mention the experimental
years at Fordham University in New York and Institut fur Festkorperfor-
schung in Jtilich, and the changing times at Leiden. All these places and
their staffs, my colleagues, contributed to a rather particular and sometimes
personal interpretation of the spin glasses which is related in this book. I
hope it reads a bit like an adventure novel as so many parts have seemed
to me, none the least of which was the actual writing. Read well.
Introduction to randomness
and freezing
r( 0 0 0 0 + Origin
Fig. 1.1 Frozen spin glasswithJ, > 0 (ferromagnetic) andJz < 0 (antiferromagnetic).
The zig-zag links indicate a broken bond; the spin in the heavy circle is frustrated.
1.1 What is a spin glass? 3
one with all antiferromagnetic interactions as shown in Fig. 1.2. Here there
is total confusion among the surrounding spins once the direction of the
centre spin is chosen. In terms of energy consideration, it will not make
any difference whether the frustrated spin points up or down. You can
imagine the complicated ‘frozen-frustrated’ situation in a real 3D lattice.
With such simple pictures we can visualize the ground (T = 0 K) or
frozen state of a spin glass. Its basic ingredients are randomness (here in
site occupancy), mixed interactions (we shall return to this point later), and
frustration: many possible ground state configurations. In order to gain a
deeper insight into this new state of magnetism, we should consider how it
is formed out of the randomly churning spins at high temperature. This
‘freezing process’ and its associated dynamics will be introduced in a later
section of this chapter.
Let us now give a working definition of a spin glass based upon the
above schematic picture. We may define a spin glass as a random, mixed-
interacting, magnetic system characterized by a random, yet co-operative,
freezing of spins at a well-defined temperature Tf below which a highly
irreversible, metastable frozen state occurs without the usual long-range
spatial magnetic order. A key word in our definition is ‘co-operative’ which
usually implies ‘ordering’ or even a phase transition. The question that the
spin glasses pose is a paradoxical one: order out of randomness? How? As
we shall see later there are a number of caveats to our definition, but in
any event we now have a foundation. From now on, the term spin glum is
used to generically represent the class of materials exhibiting the frozen-
state transition. We shall become more specific as we progress.
The two most important prerequisites of our spin glass are firstly the
randomness in either position of the spins, or the signs of the neighbouring
couplings: ferro ( t t ) or antiferromagnetic ( t J ). There must be disorder,
site or bond in order to create a spin glass, otherwise the magnetic transition
will be of the standard ferromagnetic or antiferromagnetic type of long-
e l 0 l
Fig. 1.2 Completely frustrated triangular (hexagonal) lattice with all antiferromag-
netic interactions (Jr < 0).
4 Introduction to randomness and freezing
As so often happens and long before the name spin glass was coined,
nature already provided us with the magnetic alloys. They are composed
of magnetic impurities bearing a moment or localized spin and occupying
random sites in a non-magnetic host metal. The spin glasses were accidentally
discovered with such binary alloys. Here we want to control the concentration
x of these impurities so that they can interact with each other. Now if we
take Fig. 1.1, make it a 3D metal, with a real lattice structure, e.g. face-
centred cubic, and then randomly distribute the little arrows on a proper
fraction of sites, we will most probably find a spin glass at low temperatures.
The host can be just about any non-magnetic metal that dissolves the ‘good
moment’ elements such as Mn, Fe, Gd, Eu, etc. The archetypal specimens
of the metallic (site-random) spin glasses are Cur-,Mn, and Aur-,Fe,.
These noble-metal alloys are also called canonical spin glasses. Site
substitution of magnetic solute for non-magnetic solvent should occur
completely randomly, i.e., without even short-range order of a chemical or
atomic nature. For then we can treat the system as statistical and use
gaussian probabilities. With a proper choice of elements this can be
accomplished to a reasonably fair approximation.
Chemical compounds both insulating and conducting can also be made
random. This is attained by diluting into one of the sub-lattices a magnetic
species in place of a non-magnetic one. Some prime examples here are
Eu,SrI-,S (a semiconductor) and LaI-,Gd,A12 (a metal). Many, many
hundreds of additional combinations can be formed in the laboratory. And
this plethora of systems has made the spin glasses a general and common
type of magnetic ordering.
Moreover, the above random-site occupancy of the alloys, mixed insulating
compounds and ‘pseudo-binary’ intermetallic compounds can be mimicked
in a more modern and violent way. One takes a standard intermetallic
compound, e.g. GdA12 and YFe2, and destroys the crystalline lattice by
making it amorphous. Techniques such as splat-cooling, quench-condensation
and sputtering are employed to de- or un-crystallize the compound. Figure
1.3 illustrates the loss of a periodic lattice with an integer ratio (or fraction)
magnetic to non-magnetic sites. A l-to-l ratio was chosen for simplicity in
1.2 How to create randomness? 5
Fig. 1.3 Amorphous spin glass for which the disordered lattice sites are 50%
occupied with magnetic moments.
Fig. 1.3. Really what matters now is the random apportionment of distances
between the magnetic impurities. The same gaussian distribution function
P(r) occurs here as in our original random-site alloys and the mixed
compounds. The variance or distribution width may be narrower for the
amorphous materials sketched in Fig. 1.3, nevertheless, it is the positional
randomness that creates an impurity-spin distance distribution which is the
first and basic ingredient of our would-be spin glass.
So much said for site disorder. Let us now consider another possibility
to produce a spin glass. Suppose we view the perfect lattice of Fig. 1.4,
every site dot is occupied with a spin. Assume that the magnetic interaction
is only between nearest neighbours (short-range) and imagine that this
single coupling J alternates randomly in sign with +J values as we move
through the lattice. Thus, there are only parallel or antiparallel bonds as
illustrated by the ‘dash’ or ‘zig-zag’ in Fig. 1.4. This is the random-bond
type of system and it was only recently that their existence in real materials,
e.g. the compounds Rb,Cu,-,Co,F, and FeI-,MnXTi03, were unearthed
to give reasonable approximations of 9 couplings. But more about these
random-bond spin glasses when we arrive at the theoretical models for a
spin glass. Suffice it to say, there must be disorder in the constitution of a
spin glass: either site randomness with a distribution of distances between
the magnetic spins, or bond randomness where the nearest-neighbour
interaction varies between parallel coupling +.I and antiparallel coupling
-J. If the system is truly random, the average value of all the exchange
bonds is zero. So we now know how to create the all-important randomness.
Introduction to randomness and freezing
1.3.2 RKKY
For the magnetic alloys the conduction electrons lead to stronger and
longer-range indirect-exchange interaction. This is the now-famous Ruder-
man, Kittel, Kasuya, Yosida (RKKY) interaction whose hamiltonian is
X = J(r)Si * Sj. Embedding a magnetic impurity, i.e., a local moment (spin
Si), in a sea of conduction electrons with itinerant spin s(r) causes a damped
oscillation in the susceptibility of the conduction electrons, and thereby a
coupling between spins (Si and Sj) according to
J(r) = 673J2N&)
cos (2k,r)
(w9’ 1
where Z is the number of conduction electrons per atom, J the s-d exchange
(1.1)
constant, N(E,) the density of states at the Fermi level, kF the Fermi
momentum and r the distance between two impurities. This reduces to
J(r) = Jo COS(%r+d’)
C&r)”
(1.2)
at large distance. A phase factor 4, has been included to account for the
charge difference between impurity and host and the former’s angular
momentum. Such oscillatory behaviour of J(r) or really the Pauli suscepti-
bility, which in the free-electron model has spherical symmetry, is illustrated
by the two coupling schemes in Fig. 1.5. Notice that the (l/r)” fall-off is
sufficiently long-ranged so that it can effectively reach a number of nearest-
neighbour sites. Now suppose we put a second magnetic impurity with spin
Sj at one of the neighbouring sites. This spin will produce its own RKKY
polarization and the two conduction-electron-mediated polarizations will
overlap in such a way as to establish a parallel or an antiparallel alignment
of the two spins. These situations are sketched in Fig. 1.5 by plotting the
Pauli susceptibility of the conduction electrons. Note that the sign (+ = t t
and - = t J ) of the impurity coupling varies with distance. If we combine
this property with the site disorder (various separations between the spins),
discussed in the previous section, we have generated a random distribution
of coupling strengths and directions. A computer simulation for some half-
a-million RKKY bonds, Jij, coupling a random-site magnetic alloy generates
the probability function P(Jij) shown in Fig. 1.6. Important is the near
symmetry in the number of + and - bonds. Here the required factor of
‘competition’ among ferro and antiferromagnetic exchanges is obtained in
8 Introduction to randomness and freezing
2000-
1000-
500.
xx)-
-FT
.z 100 -
3
=.
5 50-
&
2
7 1
a mI
10 1
I
5 t
a natural and common way, namely via the oscillating RKKY. This is why
the whole spin-glass problem started with the magnetic alloys. We must
once again emphasize the combination of site disorder with the + and -
RKKY interaction causing a mixture of competitive bonds that will eventually
lead to frustration in some of these bonds.
1.3.3 Superexchange
1.3.4 Dipolar
(1.3)
Thus far we have listed the mixed or competing interactions which are to
be combined with the random site occupancy to form a spin glass. But
what about the -+.I random-bond mechanism mentioned at the end of
section 1.2 and illustrated in Fig. 1.4? Now we must rely on using two
different magnetic species, in our previous examples Cu (which in an
insulator, but never in a metal, can be magnetic S = i) and Co, and Fe
and Mn. Let us examine the insulating, superexchange Rb&u,-,Co,F,
system. Here the tetragonal crystal electric field experienced by the Co
(S = s) splits the various spin states such that only the lowest S = $ doublet
is populated at low temperatures. Hence both the Cu and Co ions have
effectively S = i. The crystal-field splitting also causes a single-ion anisotropy
(see next section) which forces the spins to point only up or down along
the z (c-axis) direction. This simple, up or down, collinear orientation of
the moments we call an Ising character. At this point we bring the nearest
1.4 Magnetic anisotropy (plays a role) 11
We have already been forced into using the notion of magnetic anisotropy
which offers a preferred direction for aligning the spins. The simplest form
of this anisotropy previously used in Fig. 1.2 is the Ising model in which
all the spins are constrained to point either up or down. This is caused by
a very strong uniaxial anisotropy in one direction, usually. taken as the z-
axis. Alternatively, the spins can also be forced to lie in a 2D plane, then
we have x-y anisotropy. Remember that the Heisenberg model is, of course,
fully isotropic, and the spins can align in any direction, even non-collinear,
of the 3D space. These various types of magnetic anisotropies will become
0.6 r-
Fig. 1.8 Bond probability function P(J) for RbzCuI...,CoxF4 with x = 0.29. From
Dekker (1988).
12 Introduction to randomness and freezing
xg = -D~((s:)’ (1.4)
SITS,
= -CDi(Sf)’ (l-6)
i
The word frustration and its illustration have already appeared at the very
beginning of this chapter. Figure 1.2 sketched the basic triangle and
antiferromagnetically bonded Ising spins. The result was six degenerate
ground states (all with the same energy) for which there is always one
unsatisfied bond. Try this exercise yourself. The concept of frustration,
implicit in many physical and biological phenomena, was first clearly used
and quantified in the description of the spin glasses. It is the essential
ingredient to establish the special ground state. So let us look a bit closer
at frustration.
Suppose we have a square or plaquette of four Ising spins with four &J
bonds as shown in Fig. 1.9. For the left-hand configuration (Fig. 1.9(a)) of
+ +
++ +
-I u
(a) lb)
Fig. 1.9 Square lattice with mixed interaction +:F, -:AF. (a) unfrustrated plaquette
@ = 1 and (b) frustrated plaquette Q, = -1.
14 Introduction to randomness and freezing
two + bonds and two - bonds, all the bond energies are satisfied and
there will only be a two-fold-degenerate ordered state. This arbitrariness is
caused by the initial choice of the first spin which could be set either up
or down. Here we say the plaquette is unfrustrated. Moving over to the
right-hand configuration in Fig. 1.9(b) (three + and one - bonds), all the
bond energies cannot simultaneously be satisifed. One spin remains frustrated
or one bond is broken no matter what we do. Notice how the degeneracy
increases to &fold (count them for yourself). For this trivial example eight
energetically equivalent ground states are possible. In a mathematical way
we may write the frustration function as @ = II sign (Jii) = ? 1, where the
minus sign denotes a frustrated system. For the case shown in Fig. 1.9,
Q = J1J2J3J4 = +l for the unfrustrated left-hand side and -1 for the
frustrated right-hand side.
If there is a great deal of frustration in our system, many of the plaquettes
will be frustrated. Connecting such plaquettes with a set of minimal distance
lines gives a measure of the excess energy caused by the frustration. The
task, and it is not an easy one, is to find the minimum length between the
various frustrated plaquettes and thereby determine the lowest ground-state
energy. There is great similarity between this ‘optimization’ problem and
others such as the intriguing ‘travelling salesman’ problem (see Chapter 7).
Note that the frustration creates a multidegenerate, metastable, frozen
ground state for the spin glass. And as we shall soon see such a ground
state leads to many interesting and unique experimental properties. But
now we should ask the question - ‘is frustration alone enough to generate
a spin glass according to our working definition?’ The answer is ‘No, it
must be coupled with disorder or randomness, site or bond, our first basic
ingredient’. An antiferromagnetically coupled, regular triangular lattice
(which is full frustrated) will not exhibit the co-operative freezing transition
of a spin glass. More likely, a slow ‘blocking’ will occur to a ground state
which is dominated by large magnetic fluctuations. Furthermore, there
seems also to be the need for mixed (+ and -) interactions supplementary
to the randomness in order to attain our spin glass. Thus, and this aspect
is still under contention, a system with only antiferromagnetic interactions
plus disorder will not yet be a good, co-operative spin glass. The ‘extra
necessary’ ingredient would be to include some ferromagnetic bonds.
Support for our standpoint comes from very recent experiments on KagomC
(triangular-like) lattices with only antiferromagnetically coupled ions which
do not show generic spin-glass behaviour.
Let us recapitulate the essence of the previous sections to secure our
point of view. Randomness or disorder is indispensable and can be created
by varying the distances between the magnetic species or by randomizing
the bonds among a periodic set of magnetic ions. The mixed
(ferro-antiferromagnetic) interactions are essential to install the competition
and ensure co-operativeness of the freezing process. Anisotropy seems to
be required to establish preferred directions along which the local spins can
1.6 The freezing process (a first look) 15
At this point it would be instructive to try to present our first naive picture
of the evolution of a spin glass as we reduce the temperature from far
above the freezing temperature Tf to far below. Our description is based
on experimental observation which will be explicated and discussed in
Chapter 3. We now shall hope to gain some rudimentary insights into the
underlying physics of the freezing process.
Of course, as T + CQ,there will simply be a collection of paramagnetic
spins, i.e., independent and rapidly rotating arrows in the ‘chaos’ caused
by the high temperature. When the temperature is lower from T % Tf many
of these randomly positioned and freely rotating spins build themselves into
locally correlated units or clusters, even domains, which can then rotate as
a whole. These ‘clusters’ may simply be ferromagnetically coupled resulting
in ‘giant’ or superparamagnetic moments. Or they may form from a strongly
localized overlapping of RKKY interactions, in which case, because of the
(+) parallel, (-) antiparallel oscillations, the net moment will be proportional
to the square root of the number of spins taking part. The remaining spins,
those not belonging to any cluster, are independent of each other, but help
to transmit interactions between the clusters allowing for changes in the
cluster sizes and their relaxation or response times. This formation of
clusters can be tracked with various experimental techniques and is a direct
consequence of the randomness and mixed interactions.
Now as T + Tf the various spin components begin to interact with each
other over a longer range, because the temperature disorder is being
removed. The system seeks its ground-state (T = 0) configuration for the
particular distribution of spins and exchange interactions. This means a
favourable set of random alignment axes generated by the local anisotropy
into which the spins or clusters can lock. However, as we learned in the
previous section frustration plays its role and a multi-degenerate array of
ground states presents itself for the system to choose from. Since there is
a spectrum of energy differences between the frozen states, the system may
become trapped in a metastable configuration of higher energy.
What type of transition or transformation occurs at the freezing of a spin
glass? Is it a phase transition with Tf a critical temperature (T,), treatable
within the theories of critical phenomena ? Or is there a gradual blocking
of the cluster rotations so that the system slowly settles into the frozen
state? A third possibility exists somewhere in between these two extremes,
where a new type of ‘random’ transition occurs requiring a very different
theoretical description. How are the spin glasses and their freezing related
16 Introduction to randomness and freezing
General magnetism
Historical background
References
In order to better present the spin glasses and establish their place in
magnetic phenomenon, we should first consider how they evolved and
where they belong with respect to other allied magnetic effects. Some
perspective at this point is important and a number of cognate terms are
in need of introduction. This chapter will briefly survey the related or
surrounding phenomena from which the spin glasses developed. The main
portion of Chapter 2 is devoted to the dilute magnetic alloys. Moment
formation via the virtual-bound-state model and the Kondo effect along
with the giant moments will be introduced. Afterwards we will discuss the
various types of behaviour which appear in the different concentration
regimes extending from very dilute to fairly concentrated. This is to show
exactly where the spin glasses lie with respect to their disorder. Here a
valuable new concept for us is percolation which distinguishes the spin
glasses from the more-concentrated, long-range-ordered, random systems.
Then we shall view the magnetic glasses, since amorphous magnetism is
certainly a part of the spin-glass problem and many of these materials
exhibit the novel ‘re-entry’ phenomenon. Finally, the notion of superparamag-
netism is explored and its applicability to the spin glasses perused, and to
conclude, the insulating and semiconducting systems will be mentioned and
identified.
ivJ(J + l)g$$
X= 3kB T (2.1)
Fermi energy. This means they will be occupied with different amounts of
spin-up d-electrons and spin-down d-electrons. A reduced Coulomb repulsion
results in an upwards shift of the up-vbs relative to the down-vbs. This
further increases the overlap and is shown in the left-hand side of Fig. 2.2.
The net vbs occupancy of up-spin minus down-spin will determine the ‘local’
d-moment which could easily be non-half integral. Thus, any value of spin
or moment could appear in a dilute magnetic alloy. Or, if we even push
our parameters further IV,,] + UN,, = 0, we arrive at the symmetric case
of up-spin occupancy equal to down-spin occupancy, and no net spin, i.e.,
the impurity loses the magnetism; it is effectively dissolved in the sea of
conduction electrons. Figure 2.2 (right-hand side) depicts this non-magnetic
possibility.
The main consequence of the virtual-bound-state model is that an atomic
magnetic impurity may not even form a moment when introduced into a
metal. Everything depends on the given set of parameters. If it does, then
any (not just half integer) value of spin or moment may be generated. The
model had its heyday in the early 1960s when the problem of moments
formation was a paramount issue. What the model does not answer is how
the moment forms as a function of temperature and what effect this
temperature dependence will have on the experimental properties.
We shall look further into these questions in the next section. But before
concluding this brief sketch of the vbs model, let us mention the
Schrieffer-Wolff (S-W) canonical transformation which allows us to combine
the various model parameters into an effective exchange interaction J. In
1966 S-W showed that in the strongly magnetic limit [the case displayed in
Fig. 2.1 (right-hand side)]
(2.5)
Fig. 2.2 Schematic diagram (left-hand side) of partially magnetic vbs. Note how
the magnetic moment can form with any spin value. Right-hand side represents the
symmetric case of equal spin-up and spin-down occupancies (U = 0) and thus no
net magnetic moment.
24 Description of related concepts
The subject of dilute magnetic alloys may be divided into two classes of
interest. Firstly, non-interacting or very-dilute magnetic impurities dissolved
into a non-magnetic host can be classified under the general heading ‘Kondo
effect’ - a localized antiferromagnetic interaction of an isolated or single-
Fig. 2.3 The effect of covalent mixing on the vbs which siphons off the moment
and causes antiferromagnetic coupling (J < 0) between the local moment and
conduction electrons.
2.2 Kondo effect and weak moments 25
Table 2.1 Isolated impurity Kondo (spin fluctuation) temperatures for important
dilute alloy systems
Ag
Au 300
- K 10
1 mK <10e6
<10d6 K 0.25 K 500
- K >lOOO
- K
MO
Rh 50 K 50 K 1000 K simple
- 10 K 1K 25 K I
Pd 100 K 10mK 20 mK 0.1 K I exchange
Pt 200 K 0.1 K 0.3 K 1K enhanced
26 Description of related concepts
of states at the Fermi surface. For T < TK there is a gradual loss of local
moment as the conduction electrons begin to form a surrounding cloud of
oppositely (antiferromagnetically) polarized spin. This situation, i.e., that
of a quasi-bound state, is sketched in Fig. 2.4. Its formation is not a phase
transition, but a slow transformation on a logarithmic temperature scale.
In the calculations, this broad-temperature transition to the quasi-bound
state is heralded by the appearance of troublesome logarithmic divergences
which reduce the response of the moment to external fields and enhance
the electron scattering cross-section. Note that the entity comprising the
quasi-bound state does not remain fixed in its orientation. Instead, it churns
about as a whole, a rotating, yet compensated, moment.
Experimental manifestations of the Kondo effect T < TK are:
(i) a loss of magnetic moment, the magnetization falls below its free-
moment value and the susceptibility is less than its Curie value
NP2 x (2.7)
XK = 3kB(T + TK) < xc = 3k,T
where p = gp,n[S(S + l)];.
(ii) The enhanced scattering rate creates a logarithmic upturn in the
resistivity at low temperature.
Fig. 2.4 Static picture of the Kondo bound state with local moment surrounded
by oppositely polarized conduction electrons.
2.2 Kondo effect and weak moments 27
‘\--/ T min
T
Fig. 2.5 The resistivity minimum in the temperature dependence of a dilute alloy.
Notice that the Kondo contribution becomes saturated (‘unitary scattering limit’) as
T+= 0.
at k,T, above the Fermi energy with a height a l/T,. It is this impurity
(Kondo) resonance which creates the many interesting and anomalous
properties.
What we neglected to mention before is that the specific heat of a Kondo
system is large and has a broad maximum at approximately TK. This along
with the magnetization and susceptibility has now been calculated precisely
from the above exact theory. The trouble was then to find a suitable, real,
ideal material whose thermodynamic properties could be directly compared
to theory. A similar difficulty exists, as we shall later see, with the spin
glasses. Finally, (La,-,Ce,)A12, where TK = 0.2 K and S = $ for the crystal-
field split ground state, was deemed appropriate. And a very nice,
quantitative fit between experiment and theory resulted, at least for the
measurable thermodynamic quantities, magnetization, susceptibility and
specific heat. Yet what remains to be calculated are the transport properties:
the resistivity increase and the giant thermoelectric power associated with
the Kondo effect. Up to now there have been no exact entire-temperature-
range solutions to compare with the large amounts of experimental data
available.
The occurrence of the Kondo effect is a hindrance to spin-spin interactions
and magnetic ordering. In Table 2.1 a collection of Kondo temperatures is
given for some transition-metal alloy systems which do indeed exhibit
interactions and ordering effects at sufficiently high concentrations. At very
low concentrations and below TK the moments are ‘weak’ and cannot simply
interact with each other. However, the trick here is to bring the impurities
sufficiently close to each other and turn on the RKKY interaction. For, we
can greatly eliminate the role of the Kondo effect and weak moments by
increasing the concentration and creating good moments via a strongly
magnetic local environment, since when interactions are taken into account
T,(single impurity) % T,(pair of impurities) % T,(triple), etc.
3pB(bare moment)
O.l- ’
F PdZnn
E _ Pd 3nn
7J
z PdSnn
2
.c
o-
Fe I I I 4 I I I I I I
0 L
2 !I, for f.cc. Pol 6 8 A
Since we had our first glimpse at what the spin glasses were in Chapter 1,
let us now consider what they are not. The moments must be good, not
weak - no significant Kondo effect. The concentration must be large enough
so that there is enough competition from the mixed interactions - not the
ferromagnetic predominance of the giant moments and not long-range,
spatial ordering of the usual ferro- or antiferromagnetic transitions.
This last condition is quite interesting. Mathematically, it may be expressed
as ZiSi = 0 (no ferromagnetism), but (S,), f 0 where ( ), represents a time
average over a long observational time t, and in addition
&~(SJ,eXp(ik’Ri)=O asiV+m
I
This equation states that there is 120 long-range antiferromagnetic ordering,
i.e., the directions of the spins are not related to a particular wave vector
k, as is true for even the most complicated of antiferromagnets with lattice
vectors Ri.
Another way of viewing the spin ordering is via their Ising +l spin-spin
correlation function
where cSG is the spin-glass correlation length. The crucial questions here
are: does the spatial correlation length & + ~0 as T + T,, and does the
decay of the ‘order parameter’ governed by a characteristic time become
infinitely long signifying a phase transition? The first would imply the
second as in a conventional second-order phase transition. However, the
spin glasses are different and we do not really know the final answers to
the above questions. Long is not infinite. So let us now hypothesize a phase
transition and study its consequences. It follows that we can replace the
time averages by the ensemble averages of statistical mechanics and define
an order parameter for the frozen spin-glass phase:
4 = ((SJ$)c # 0 for T I Tf (2.11)
where the inner brackets represent ensemble (or thermal) averages while
the outer brackets are for a configurational average over the set of random
interactions. 4 is called the Edwards-Anderson spin-glass order parameter
and will appear often as we tread further into the models and mysteries of
spin-glass theory.
2.4 Spin glasses (a second visit) 31
Since the magnetic impurities have random positions in the lattice, the
effective molecular field, H, may be calculated as a superposition of
contributions from all the impurities in the alloy, due to the infinite range
of the RKKY interaction. Furthermore, H has a distribution of magnitudes,
P(H), and a random distribution of directions because of the oscillatory
nature of the RKKY interaction. A computer simulation has been made
for P(H) using an fee lattice with about 100000 sites, 324 being occupied
with spins (x = 0.3%), and an RKKY interaction performs the coupling
among these vector spins Si. The results for P(H) are shown in Fig. 2.7.
The reader should compare this figure to Fig. 1.6 which gives the probability
of RKKY interaction strengths. Notice that P(H) + 0 as H + 0 meaning
that at each lattice site occupied by a spin there will be a finite internal
field having both a magnitude and direction. Although P(H) generates the
same function for a large enough spin system, it is not required that all the
spins point in exactly the same direction for each simulation or ‘freezing’.
The freezing temperature Tf of a spin glass may be viewed from a very
simple model. By considering only the envelope of the RKKY oscillations,
J’(r) = J,lr3, we have the strength of the interaction as a function of
distance. For a low-concentration alloy the average distance of impurity
separation is r E (l/x)“. Now we use the thermal disorder energy, k,T, to
destroy the correlations between the spins, on average, P apart. In other
words, the exchange energy caused by the RKKY interaction from a spin
Sr, acting at a distance P on another spin S2 must be greater than thermal
disordering for freezing to occur. At the freezing temperature,
This means that no matter what the concentration is, there will always exist
a temperature for which the RKKY interaction will prevail and produce
this random (due to the oscillations) freezing. Here only good moments are
assumed, i.e., no Kondo effect or TK = 0.
A further remark about these dilute alloys and the RKKY interaction
has to do with the relation (xp3) = constant. If a measure of length is used
which is related to the system of impurities, then we can ‘scale’ the
properties for one concentration x1 to those of another x2. Our length scale
is now the average distance between impurities J, so the volume f3 contains
a constant number of magnetic impurities equal to xf3. This has particular
significance for the thermodynamic properties. Magnetization, susceptibility
and specific heat can be expressed in the form
ii4
-= HT c
and p- - H T withT,ax.
x f(- x ‘x-) x fi- x ‘x-)
(2.13)
Experiments in the dilute limit [just above the concentration, x,, where the
Kondo effect plays a major role, viz. TK = T,(x,)] verify these ‘scaling
laws’ and thereby prove the efficiency of the RKKY interaction in creating
the spin-glass state. We shall return to the roots of this scaling at the
beginning of Chapter 5.
................................... ....
..\......~~..~.......r ...\
.. ..w.ir(........~~) ...... ..-CO..
b....-w.+.......~/#f.......~~~ ..
.. ..wb~......I...i#/p..f....~\.~* .
... ..h.~../....irf*b....~~.\.~ ..
..~..i..i...................~.\ ...
......... ................... ..*hv
../...~.~.q....f...........-..b ....
....... ..*~.j........~ ..........
... ..Y...../*: ..................
........ ..~.~.~~.............\ ...
.. ../.....C1’b......j ...........
\.......b.......f..b..~ ........
.... ../ ........................
..b.*
........ ................
..................... ..~~.....#..~
..~...............~~-~.....-......
. . . . . . . ..m.r....i~~.........
.......
.........
. ..#.b........b.-.
Fig. 2.8 Spin glass with mictomagnetic cluster for a 2D square lattice.
For the sake of completeness, let us proceed and further increase the
number of magnetic atoms. At some given concentration for the particular
crystal structure, each magnetic site will have at least one magnetic nearest
neighbour. When such a macroscopic connection or uninterrupted chain
extends from one end of the crystal to the other, the percolation limit (that
concentration for which there is a non-zero probability that a given occupied
site belongs to an unbounded cluster) has been reached. The schematic 2D
lattice shown in Fig. 2.8 is far below the site percolation concentration
which, for a square lattice with only nearest neighbour coupling, requires
58% of the total sites to be occupied. Similar critical concentrations for
percolation may be obtained for other 2D and 3D crystal structures with
various near-neighbour interactions.
Another type of percolation is bond percolation where an interaction or
bond which couples two particular sites is present or absent. Then the
closed bonds must continue in a macroscopic path through the crystal. A
simple physical picture for bond percolation is conduction through a network
of on/off switches. How many switches between the sites of our square
lattices (Fig. 2.8) need be closed before a current flowing in one end of the
network will flow out the other? 50% is the answer.
Percolation applied to magnetic systems is more complicated. Let us
begin with a 3D fee lattice and simple nearest-neighbour interactions, and
randomly fill the lattice sites with magnetic atoms; if they are nearest
neighbours, they will align parallel; if not, they are uncoupled. When 17%
of the sites are occupied an infinite ferromagnetic cluster will form. This
represents long-range ferromagnetic ordering and a proper phase transition
at a Curie temperature TC. Certainly there will be modifications in the
critical phenomenon due to the percolation behaviour and its fractal nature,
and also because of the many ‘free’ spins which are not (or only weakly)
coupled to the infinite cluster. With nearest-neighbour antiferromagnetic
bonding in the same 3D fee crystal, a much larger concentration is needed.
Here the element of frustration or unsatisfied orientation enters, because
the fee triangular symmetry is unfavourable for long-range antiferromagnetic
order and many more spins are needed to build this coherent two sub-
lattice structure in fee space. Now the percolation limit is = 45% site
occupancy above which we have a long-range antiferromagnet with a distinct
T N*
We can take the following step in our percolation approach and allow
mixed interactions, remembering this is the essential ingredient for our spin
glasses. For example, return to Fig. 1.1 where the first neighbour was
ferromagnetic and the second, antiferromagnetic. There does exist an
‘infinite’ coupled cluster, but now without long-range ‘periodic’ order. The
spins are frozen randomly depending on the initial condition. Here we
might say that the spin glass correlation length &o + 03, but recall the
2.7 Concentration regimes 35
After all this discussion we should try to summarize the various concentration
regimes of different behaviour for a metallic magnetic alloy. It must now
be clear that the concentration is most important in determining the
magnetic state of the alloy. Figure 2.9 offers a schematic representation for
such a concentration regime division. At the very dilute magnetic concen-
tration (ppm) there are the isolated impurity-conduction electron couplings
which result in the Kondo effect. This localized interaction (if J < 0) causes
a weakening or fluctuation of the magnetic moment, and below the Kondo
temperature, the moment disappears and the impurity appears non-magnetic.
Thus, the Kondo effect prevents strong impurity-impurity interactions which
are a basic necessity of the spin glasses. Nevertheless, the local-environment
model tells us that there will be pairs (or triplets, etc.) which possess a
much lower TK than the singlets. These ‘good’ moment pairs may then
interact with other pairs and give rise to spin-glass behaviour. Therefore,
in principle, there is no lower concentration limit to spin-glass consequences,
until we run out of measurement response or temperature. Usually, however,
we take TK = T&J as the concentration, x0, below which the Kondo
effect plays a large role in modifying the spin-glass behaviour.
Up to a concentration of few thousand ppm (- 0.5 at. %), the scaling
approach (see section 2.4), based on the (1/r)3 decrease of the RKKY
exchange is particularly appropriate for describing the problem of interacting
impurities. This means that the measurable properties (magnetization,
susceptibility, specific heat, and remanences) are universal functions of the
concentration scaled parameters T/x and H,,,lx, and in addition Tf c( x.
While an effective way to scale the various data, such a treatment tells us
nothing about the freezing process and the possibility of a phase transition.
At low temperature T -@ Tf and with the incorporation of anisotropic dipolar
interactions, also m(llr)3 and thus scalable with x, the magnetic behaviour
was initially explained within the existing model of NCel superparamagnetism.
Superparamagnetism (see below section 2.10) is the temperature and field
cluster formation
Exp. properties-
..,..,
::;,:.:: variables 8
= 50 ppm = i at. % = 10 at. % Magnetic
percolation ‘p9,
T, = To(c) limit %
3
RKKY RKKY + short range correlation direct
Dipolar
interactions
Fig. 2.9 Various concentration regimes for a canonical spin glass illustrating the different types of magnetic behaviour which occur.
2.8 Metallic glasses: amorphous magnets 37
This name has been given to a new class of non-crystalline solids which are
conducting and contain magnetic elements. Most of the metallic glasses are
naturally based on rather large concentrations of transition elements (TM),
e.g. FesoP2, and COAST, so long-range ferromagnetism is a common
occurrence. It is only when the TM-concentration is reduced by non-
38 Description of related concepts
Fig. 2.10 General (schematic) T-x phase diagram for a dilute magnetic alloy.
At the percolation boundary of phase diagram, recalling Fig. 2.10 and its
dashed lines, a strange behaviour takes place, if the long-range order is
ferromagnetic. The spin-glass phase does not simply stop but persists into
the ‘ferromagnetic’ phase. This effect is shown for a real material Au,-,Fe,
in Fig. 2.11. The strangeness is related to the observation that the more
disordered state appears from the more ordered one as the temperature is
reduced. According to experiment it would seem that the T = 0 phase is a
cluster glass created out of the ‘percolated’ infinite ferromagnetic cluster
which has broken up into large but randomly frozen clusters. Hence, the
apparent long-range order is destroyed although a gamut of large ferromag-
netic clusters is still present - conforming to our previous definition of a
cluster glass or mictomagnet. The word re-entrant, something of a misnomer,
means here that the cluster-glass phase develops from a ferromagnetic state,
and thus it re-enters the frozen (disordered) phase out of another, not
paramagnetic, state.
There are many different systems especially among the amorphous metallic
magnets which show this re-entrant behaviour. So the phenomenon is
general, yet a good theoretical model is hard to find. Various suggestions
have been proposed, one of which is the appearance of a temperature-
dependent random anisotropy (dipolar?). At low temperatures this anisotropy
grows sufficiently large that it severs many of the weak links holding
together the rather tenuous x z xP infinite cluster. Then the anisotropy
points these divided units along the randomly preferred directions. Another
model involves a freezing of transverse spin components. At the ferro-
magnetic transitions T, the longitudinal components order, but because of
the local mean field being inhomogeneous, a random freezing perpendicular
to the applied field occurs at a lower temperature Tf.
Recent experimental work indicates that the ‘infinite-cluster’ ferromagnetic
state is not truly long-ranged. Instead, it consists of many extremely large
40 Description of related concepts
l Mossbauer
200
l Susceptibility
0 Neutron scattering
0 Specific heat
f
A. Magnetic resonance
160
/
para
I
120 ferro
0 I
e-
l- I
4
I
80 -
b
: / 1
O 4 8 12 16 20 24
x(at%FeI
Fig. 2.11 Magnetic phase diagram of AuFe constructed from the anomalies observed
in various experiments according to symbol; from Coles et al. (1978).
clusters which are hard to tell from infinite. These large, but disordered
from each other, clusters then dissociate into the collection of smaller
clusters at the re-entry transition. Such processes occur without the usual
type of percolative phase transition.
2.10 SUPERPARAMAGNETISM
M~+]=M,(n)[(S+t)coth((S+Q~)-Icoth(!!)]
(2.15)
Since each particle contains many spins, we require the limit of the Brillouin
function as S + 03. Can you remember how this function depends upon S?
For S becoming large the initial slope decreases and there is substantial
rounding as the argument increases making saturation more difficult.
Confirmingly, experimental data nicely fit the Brillouin function with S = 03,
at least if T is not reduced too far. Since the measured magnetization for
our assemblage of particles is zero when no external field is applied, and
follows the Brillouin function for infinite spin when a field is swept, the
name superparamagnetism aptly fits.
Now we lower the temperature to look for additional effects. Here we
must include the consequences of anisotropy in an energy-barrier model as
depicted in Fig. 2.12. There are many possible sources of the anisotropy:
magnetocrystalline, shape, dipolar, surface, etc. The anisotropy constant
times the particle volume forms the energy barrier height, E, = KV which
can be overcome with either field M . H or temperature kBT energy. Hence,
two minima appear with oppositely directed spin, i.e., two easy orientations
of magnetization. The relaxation time of the magnetization between
these two states is given by thermal activation (Arrehenius law)
0 IT
Fig. 2.12 Anisotropy barrier for an up/down superparamagnet. Note the thermal
activation energy (kBT) is too small to cause rapid barrier hopping.
42 Description of related concepts
Related concepts
Kondo, J. (1969) in Solid state physics, Vol. 23 (eds F. Seitz and D. Turnbull)
Academic Press, New York, 183.
Heeger, A. J. (1969) Ibid, 293.
Rado, G. T. and Suhl, H. (eds) (1973) Magnetism V, Academic Press, New York.
Nieuwenhuys, G. J. (1975) Adv. Phys., 24, 515.
Mydosh, J. A. and Nieuwenhuys, G. J. (1980) in Ferromagnetic materials, Vol. 1
(ed. E. P. Wohlfarth) North Holland, Amsterdam, 71.
Moorjani, K. and Coey, J. M. D. (1984) Magnetic glasses, Elsevier, Amsterdam.
Stauffer, D. and Aharony, A. (1992) Introduction to percolation theory, Taylor and
Francis, London.
Spin glasses
Mezard, M., Parisi, G. and Virasoro, M. A. (1987) Spin glass theory and beyond,
World Scientific, Singapore.
Fischer, K. H. and Hertz, J. A. (1991) Spin glasses, Cambridge.
Chowdhury, D. (1986) Spin glasses and other frustrated systems, World Scientific,
Singapore.
Binder, K. and Young, A. P. (1986) Rev. Mod. Phys., 58, 801.
Maletta, H. and Zinn, W. (1986) in Handbook on the physics and chemistry of rare
earths (eds K. A. Gscheidner, Jr and L. Eyring) North Holland, Amsterdam.
Rammal, R. and Souletie, J. (1982) in Magnetism of metals and alloys (ed. M.
Cyrot) North Holland, Amsterdam.
Mydosh, J. A. (1981) in Magnetism in solids: some current topics (eds A. P.
Cracknell and R. A. Vaughan) SUSSP, Edinburgh.
van Hemmen, J. L. and Morgenstern, I. (eds) (1983) Heidelberg colloquium on
spin glasses, Vol. 192 Lecture Notes in Physics, Springer, Berlin.
van Hemmen, J.L. and Morgenstern, I. (eds) (1987) Glassy Dynamics Vol. 275,
Lecture Notes in Physics, Springer, Berlin.
References
Coles, B. R., Sarkissian, B. V. and Taylor, R. H. (1978) Phil. Mag. B, 37, 489.
Walker, L. R. and Walstedt, R. E. (1977) Phys. Rev. Lett., 38, 514.
Walker, L. R. and Walstedt, R. E. (1980) Phys. Rev. B., 22, 3816.
3
Basic spin-glass
phenomenon
We have been considering for the past two chapters the spin glasses and
their entourage of cognate phenomena. After having spent so much time
and energy introducing the encompassing physics and basic principles of
the spin glasses in a rudimentary and qualitative way, we now should turn
to experiment in order to find out exactly how it supports and confirms
these notions, i.e., our physical picture, of a spin glass. Accordingly,
measurements on real systems and their proper interpretation are the
windows through which we gain our knowledge. Let experiment be our
guide - after all it did come first.
We shall attempt in this chapter to survey salient experimental features
of a spin glass beginning with the high-temperature measurements and then
reducing the temperature to T = Tf. Here we examine the novel critical
behaviour and the various manifestations of a phase transition. Finally, the
important experiments characterizing the low-temperature frozen state
(T < Tf) are viewed with respect to its unique properties. Our three
temperature regime approach is arbitrary, yet effective, since three very
different types of behaviour are encountered. Our choice of a few from the
many thousands of experiments published on spin glasses is geared to bring
our physical picture into focus and keep it generic - not a collection of
exceptions. Thus, we want to embed our notion of a spin glass firmly in
reality before proceeding with models and theories which, of course, should
give the ultimate explanation of the phenomenon and lead to a final
understanding.
Perhaps the dullest part of the spin glasses is their behaviour far above the
freezing temperature. For, we would expect simple paramagnetic effects in
this limit, or a collection of freely rotating, independent magnetic moments.
Using a paramagnetic model we can compare this basis to the generic spin-
46 Basic spin-glass phenomenon
3.1.1 Susceptibility
temperature (K)
Fig. 3.1 The reciprocal susceptibility as a function of temperature: the dashed lines
are a linear extrapolation to determine 8 and the solid line represents the fit to
the model calculations; from Morgownik and Mydosh (1981).
3.1 High-temperature (T % TJ experiments 47
(3.2)
O(x) = x
table of
flrst 11
:Onf IQUP
QPO”P5
OP D t.c c
lattlco
.
M
.-*
v
v
r
-
Yi
” ‘*
.-?4
Fig. 3.2 P as defined in the text versus temperature. The deviation from C-W
behaviours at Td is illustrated in the inset; also shown are the first 11 configuration
groups for an fee lattice. From Morgownik and Mydosh (1981).
Fig. 3.2. The hope is to obtain a better than Curie-Weiss fit, at least a bit
further down in temperature. But such a model cannot be accurate as Tf is
approached due to the co-operative interaction effects - pairs and triplets
will no longer suffice. A few years ago such a model was turned loose on
a variety of dilute alloys (i < x < 5 at. %) and even included the deviations
from randomness due to chemical clustering.
The model was a computer calculation of all possible energy levels for
the various cluster configurations according to the hamiltonian
(3.5)
i<j i
r I I I I I I
d , I I I
0 JO
0 2 4 6
concentration (at % Mn)
Fig. 3.3 Effective Bohr-magneton number p. and the paramagnetic Curie tempera-
ture 0 as a function of concentration in CuMn: p. and 0 are determined from
least-squares fitting of the susceptibility for temperatures above Td; from Morgownik
and Mydosh (1981).
(3.6)
(3.7)
The fitting parameters for this configurational cluster model were the
various exchange interactions as a function of neighbour shell (or distance).
The solid lines in Figs. 3.1 and 3.2 are the model calculations. They do
indeed fit better than the Curie-Weiss law. However, they cannot be used
too close to Tf. The net outcome was an estimate of the exchange parameters
J,,, extended, with certain modifications, to fifth nearest neighbour. Such is
illustrated in Fig. 3.4 along with the expected RKKY conduction-electron
polarization for a Mn impurity in Cu. There is more ferromagnetism present
in the model calculations than in the RKKY interaction. It would seem
that for the close by impurities (first or second neighbours), d-like conduction
electrons are mediating the indirect exchange interaction, thus causing the
50 Basic spin-glass phenomenon
o Au
- Fe
V Cu
- Mn
1
l -Au Mn
o -Pt Mn
L s?L---\-.3 45671
-1 “”
0 0.5 1.0 1.5 2.0
interatomic distance (units of lattice constant)
Fig. 3.4 Exchange parameters J,, as function of neighbour distance: the dashed
line represents the RKKY-conduction electron polarization; from Morgownik and
Mydosh (1983).
to subtract the unwanted components and obtain the magnetic specific heat.
Both the measurement technique itself and the various subtraction procedures
are laborious and imprecise, thereby making the high-temperature spin-
glass specific heat a tough quantity to determine accurately. So we must
work with low T,s or dilute systems, and see how far we can increase the
temperature before the error bars become too large.
Very careful measurements and extraction of the magnetic component
do exist for a number of canonical spin glasses. Figure 3.5 shows a now
classic set of data for CuMn 0.3 at. % where Tf = 3.0 K. We shall return
to this figure a number of times in the coming sections. Suffice it for now
to note the broad maximum in C,,, above Tf (T,,, = 1.4 Tf) and the gradual
fall off at increasing temperatures T + 4 or 4.5 Tf. Too bad we cannot go
yet higher in temperature. The entire decrease with increasing T becomes
less pronounced, quite smeared, and the maximum disappears as an external
magnetic field H is applied. Qualitatively the data exhibited in Fig. 3.5 are
generic - always a broad maximum above Tf with a slow decrease at the
highest temperatures and field smearing.
In order to compare these long-range RKKY spin-glass results with those
of a short-range insulating spin glass, we display the measurements on
20.0 I I I
2790 ppm CuMn 1
0 0
0 75
I I I
0 5.0 10.0 15.0 20.0
T (K)
Fig. 3.5 Magnetic contribution of the specific heat of CuMn spin glasses with 0.3
at. % Mn plotted versus temperature in various magnetic fields (Tf = 3.0 K); from
Brodale et al. (1983).
52 Basic spin-glass phenomenon
EuO.$$,$ (Tr = 1.7 K) in Fig. 3.6. Be careful now for the data are plotted
&IT versus T in the main part and log C, versus log T in the inset. Once
again the same general type of behaviour is found (see inset): a rounded
peak at = 1.8 Tf followed by a drop at higher T. Now this decrease seems
stronger than the RKKY spin glass. The main portion of Fig. 3.6 exhibits
the field smearing as before and also indicates the magnetic entropy S,
from the areas under the various field-dependent curves. Remember that
TC
s, = “dT= cRln(2S + 1) (3.8)
I0 T
Even a rough glance at the C/T versus T plots would tell that a great deal
of the magnetic entropy is lost or frozen-out far above Tf. And the applied-
field shifts even more entropy from the low-temperature portion to the high
temperatures. We recall that there is a constant amount of entropy available
in the sample via equation (3.8), due to the fixed number of magnetic
degrees of freedom. The larger the field the more degrees of freedom lost
at a given tied T(> Tf). Crude estimates suggest that even in zero field
7O-gO% of the degrees of freedom are already forfeited above Tf. This
doesn’t leave too much entropy, only 20-30%, for a big effect at Tf.
-10
I;
t ke
05 t \
1' I '!.
L T(K) 4
Fig. 3.6 Specific heat per Eu atom divided by k,T versus temperature Tin various
applied fields (units in tesla) for Eu,Sri-,S with x = 0.40. Inset shows the zero-
field behaviour in a log-log plot. The freezing temperature Tf is indicated by an
arrow. The upturn of the curves for C/T, seen at very low temperatures, is
interpreted as being caused by a Schottky term due to magnetically decoupled small
clusters of Eu spins. From Meschede et al. (1980).
3.1 High-temperature (T * Td experiments 53
3.1.3 Resistivity
0.161 I I I I I I
T [Kl
Fig. 3.7 Open circles: measured magnetic specific heat for CuMn (2.4 at. %).
Experimental data from Wenger and Keesom (1976). Solid line: fit to the
configuration cluster model with the J,s determined from the susceptibility fit: (a)
low temperature regime; and (b) model prediction up to room temperature.
100 200 30 0
T(K)
Fig. 3.8 Temperature dependence of the magnetic resistivity for spin glass CuMn:
the arrows represent Tf determined from the low-field a.c. susceptibility; from Ford
and Mydosh (1976).
to couple the spins, and thereby reduce the spin-disorder scattering. All
this occurs far above Tf and as a precursor to it.
In order to gain some insight into the physics of the resistivity maximum
we must develop a theory of this competition, i.e., the Kondo effect in an
interacting spin system. Recall our formula (2.6), for the single impurity
Kondo temperature or energy
(3.9)
xPS(S + 1)
A RKKY = (3.10)
EF
T, = $ARKKy In
(3.12)
S,(q) =
exp [iq .
Tz(Si)(Sj) rij]
131
This static part gives rise to an elastic scattering (w = 0) which may be
further divided into coherent (Bragg) at particular q-values or diffuse
(smeared over q-space) scattering contributions. If there is long-range
magnetic order, we find sharp Bragg peaks at the wave vector for the
ordering q. For a ferromagnetic q = 0 so we simply obtain Bragg peaks
commensurate with the given crystal structure. This is how the percolation
concentration for AuFe was established by detecting the appearance of
these Bragg peaks. Note they are never seen in a spin glass. If there is
short-range magnetic or chemical ordering, diffuse scattering will emerge
58 Basic spin-glass phenomenon
(3.18)
Cu-7.3 al.% Mn
1
Cu-52 a1.U Mn
r=lbrOZmeV
-4
@-3.Oat.XMn
r=l.9to.2 mcV
-2
r = 1.910.2 mev
Fig. 3.9 S(q, o) versus w for q = 0.08 b;-l at 300 K for several CuMn alloys. The
central region of the elastic peak is omitted in the diagram. The continuous curves
represent best fits to a lorentzian spectral function in the energy range -2.5
meV < o < 2.5 meV. From Murani (1978a).
process longer than -lop6 s will be seen as a static event. The situation is
even worse with energy analysis, for, now the specific instrumental energy-
resolution, AE, determines the static onset. Using the uncertainty relation,
any relaxation process with characteristic time T 2 hlAE will be indistinguish-
able from a completely static or elastic one. For today’s highest resolution
spectrometers (smallest AE), T of 10-7-10-8 s is achievable. Consequently,
the dynamic or inelastic neutron-scattering technique is a ‘fast observer’ of
the spin glasses. In order to take into account the above distinction, we
separate our dynamic structure factor into a truly dynamic part and an
‘instrument-caused’ static or elastic part. Accordingly,
Sb (q, w) = 2F20 ho
g2p$ 1 - exp (-holksT)
(3.21)
and
(3.22)
Consider first of all the static structure factor S: (q, w) of equation (3.22).
For the I (= h/T) values which most contribute to the integral, ho 4 k,T.
Then
(3.23)
(3.24)
This part of the total structure factor we call the ‘elastic’ scattering, since
due to the instrumental resolution AE, it appears as a static o = 0 process.
Note that all relaxations, which are slower or longer than T = h/AE = lop8 s,
are lumped into this term. If there are no slow relaxations, then 9s (q) + 0
except if we are near a Bragg peak or a q-region of diffuse (short-range)
scattering.
Now consider the dynamic structure factor & (q, o) in equation (3.21).
Here we use the quasi-static approximation such that ho G kBT and
integrate over all o. Thus
(3.25)
3.1 High-temperature (T * TJ experiments 61
(3.26)
x(4,7) = (3.27)
The $6 (q) we obtain after integrating over all energies is called quasi-
elastic scattering. It represents all the fast fluctuations and rapid relaxation
processes in the paramagnetic state. If this state is destroyed by spin-spin
interactions, then the quasi-elastic contribution will decrease.
The key concept compelled by experiment in all this formalism is the
distribution of relaxation times N(T) being very fast and b-function like at
high temperature, and then evolving as T + Tf to longer time scales via a
broad distribution. Well, what can we measure to get a better handle on
N(T)? One usually picks a series of momentum transfers or q-values at
certain arbitrary (see below) positions in the Brillouin zone. Very low q-
values mean small-angle scattering and here one could also obtain information
on the local spatial correlations and their distributions or ranges. With q
constant we then determine the scattering cross-section by sweeping the
energy transfer w. Now the big problem is separating the elastic (o = 0 _’ AE)
central peak scattering from the surrounding quasi-elastic cross-section. By
fitting the shoulders (outside the elastic peak) to a lorentzian function, we
arrive at a distinction: quasi-elastic cross-section, namely, what is contained
in the lorentzian; and elastic cross-section, namely, the remaining sharp
central peak.
Figure 3.10 shows such a separation for the scattering cross-sections of a
AuFe (10 at. %) sample. At high temperature (150 K), the elastic part is
practically zero, yet there is a lot of quasi-elastic (or fast) scattering present.
This increases with decreasing q as greater regions of real space, and
thereby more spins become involved. Then as the temperature is reduced
the elastic scattering grows at the expense of the quasi-elastic scattering
because more slow relaxation times are fed into the time window of the
instrumental resolution. The smooth elastic increase continues down across
T,, while the fast quasi-elastic scattering slowly decreases as the number of
fast relaxation processes is eliminated.
We have been emphasizing in the present discussion spin relaxation times
and their distribution which shifts to longer times and broadens. This is a
natural consequence of our neutron scattering analysis in energy (0) or
time (t) space. However, we can also relate the physical process to cluster
formation or the growth of correlated volumes of spins. The bigger the
cluster size, the slower its relaxation time. Hence, as the fast single spins
become bound into clusters, they lose their ‘fastness’. In any case, scattering
intensity in a spin glass is always transferred from quasi-elastic to elastic as
62 Basic spin-glass phenomenon
l
ELASTIC SCATTERING Se(4
#
50 100 150 20
T (K)
Fig. 3.10 The elastic scattering and the integrated quasi-elastic scattering S,(q) as
a function of temperature for three q-values for the AuFe (10 at. %) alloy. The
vertical solid line marks the temperature of the maximum in ac-susceptibility x(O)
and the dashed line the temperature of the maximum in x(q) from the neutron
results. From Murani (19786).
these clusters are formed and slowed down. The neutron scattering sees
them directly. And as we traverse Tf in the next section we can even arrive
at some more quantitative results,
Similar neutron measurements have been performed for CuMn
(5-25 at. %), but now the total (elastic plus quasi-elastic) scattering cross-
section was determined at different temperatures as a function of momentum
transfer q. Here the whole of q-space, i.e., the Brillouin zone, was scanned
for both magnetic and nuclear (non-magnetic) scattering. This refinement
was made possible by neutron-polarization analysis for which the magnetic
scattering occurred with a spin-flip of the incoming polarized neutron and
vice versa for the non-magnetic scattering. By examining the scattering
contours one can distinguish regions of chemical (non-magnetic) clustering
3.2 Experiments spanning Tf 63
Here is where the action is! The possibility of a sharp freezing transition
to a unique phase must be investigated through fine experimentation on
the canonical spin glasses. While many different techniques have been
1.0 I I 1 I ’ I
II) (b) r-4
0.0
0.0
9 0.6
$ 0.6
:- 0.4
$j 0.4
0.2
0- -
0 0.2 0.4 0.6 OB
0.6
4
s
F 0.4
‘2
Fig. 3.11 Magnetic cross-section for CuMn (25 at. %) along (105) at elevated
temperatures; from Cable et al. (1984).
64 Basic spin-glass phenomenon
employed to study the ‘phase transition’, only a few come to grips with the
dynamics (wide range of measurement times) and zero (or small) external-
field requirement. For, as we shall soon see, dynamics plays a most essential
role and a field of only a few hundred gauss wipe out the sharpness of the
transition. We made a choice of four experimental methods in the last
section, so let us now carry on with these. Our dc-susceptibility becomes
mainly ac-susceptibility and its frequency dependence is the key variable.
Specific heat and resistivity are extended to span Tf and certain types of
critical analysis are performed. We view different relaxation-time extractions
and interpretations of the neutron-scattering technique including the very
novel spin-echo method. And to gain some experimental variation on a
fine scale near T,, we introduce, at the risk of increased complexity, two
‘local-probes’ experiments: muon spin relaxation (l&R) which tests the
inter-atomic molecular fields both spatially and temporally, and the
Miissbauer effect which is a local nuclear technique for measuring the
hyperfine-field reflecting the magnetism of the electronic system.
3.2.1 Ac-susceptibility
T PK)
Fig. 3.12 Low-field susceptibility x(T) of AuFe for 1 I x 5 8 at. %. The data
were taken every t K in the region of the peak, and every 4 or 1 K elsewhere. The
scatter is of the order of the thickness of the lines. The open circles indicate isolated
points taken at higher temperatures. From Cannella and Mydosh (1972).
I 0 0 0 100 G I\
.61 ’ I I . .,. I
6 16 24 32 40
T PK)
Fig. 3.13 Susceptibility data for AuFe with x = 5 and 8 at. %, showing the curves
for zero field, and for various applied fields; from Cannella and Mydosh (1972).
CuMn 0.005
AuMn 0.0045
AgMn 0.006
P&n 0.013
NiMn 0.018
AuFe 0.010
(LaGd)Al, 0.06
(EuSr)S 0.06
W4W2 0.08
a-Co0.A120$i02 0.06
a-WOK%) (B203) 0.28
imaginary part x”(o), the absorption. They are usually related through a
relaxation time by a set of equations called the Casimir-du PrC equations
which we shall discuss later. For a spin glass there is a sudden onset of the
imaginary component near T,. We depict this behaviour in Fig. 3.15 for
(EuSr)S, an insulating spin glass. Since x” is so small it is difficult to
measure accurately and in a metal it will be distorted because of eddy
currents induced by the alternating field h. This limits x”-measurements to
about a thousand hertz even on powdered conducting samples. As with x’
the frequency shift of x” is downwards in temperature with decreasing
frequency. Now, however, the peak in x’(T) corresponds to the maximum
slope in x”(T), (dx”/dT),,,. The appearance of an imaginary component
means relaxation processes are affecting the measurement and by decoupling
the spins from the lattice they cause the absorption. Such effects are usually
not found at magnetic transitions except when hysteresis arises as in a
ferromagnet. So with an analysis of the complex ac susceptibility we have
a handle on the low-frequency or intermediate (1O-1-1O-5 s) relaxation
Temperature (K)
times. Remember that the neutron scattering tells us about the fast relaxation
processes (10e8 s and shorter).
Finally, with the newly developed SQUID techniques a highly sensitive
method exists for determining the magnetization in a very small applied
field. As H + 0 there must be a proportionality Xdc = M/H, and a quasi-
static (governed by the time of the measurement at a given T, e.g. = 100 s)
delineation of the susceptibility is gained. It is remarkable how many
decades of time or frequency may be probed using the various experimental
techniques, for, with the SQUID we have times available which can
approach lo5 s. There are two distinct ways to measure the susceptibility
with a small dc field. The first is to apply the field far above Tf and cool
the sample in a field to T 6 Tf all the while recording the magnetization.
This method we call field cooling (FC). Secondly, we can cool the sample
in zero field to T 4 Tf and at this low temperature apply the field. Then
we heat the sample while measuring M to T P Tf with the field constant.
Here we use the term zero-field cooling (ZFC). Figure 3.16 illustrates the
different temperature dependences (FC versus ZFC) for CuMn (1 and
2 at. %) with a field of 6 gauss. Notice what happens at Tf. The FC-x
becomes constant in value and to a great extent independent of time: if we
stop and wait at a given T(< T,), x remains unchanged. (Only at extreme
sensitivities and very long waiting times is there a drift in x.) Furthermore,
the process FC followed by field warming is reversible. If we cycle the
temperature back and forth (with H = const.) the FC-susceptibility traces
I ’ I I I I I
I-
I I 1 I I
O-3 IO 15 20 25
T (K)
Fig. 3.16 Field cooled [(a), (c)] and zero-field cooled [(b), (d)] magnetizations (x
= Ml6 gauss) for CuMn (1 and 2 at. %) as a function of temperature; from Nagata
et al. (1979).
70 Basic spin-glass phenomenon
the same path (see double-tipped arrows in Fig. 3.16). On the other hand,
the ZFC susceptibility is zero until the field is applied, in Fig. 3.16 at
= 5 K. Then with H on and T constant, the ZFC-x jumps to a value
comparable with that found from xac. However, now there is a slow, clear
drift upwards which continues over many decades in times. If we wait long
enough (and no one has yet achieved this - the average lifetime of a
graduate student is only about lo8 s), xzFc (t -+ m) = XFC at the given
T < Tf. So, when we proceed to increase the temperature at some reasonable
rate dT/dt, we arrive at the single-arrow curves in Fig. 3.16. These traces
are irreversible, XzFc is always drifting upwards. Therefore, the actual curve
or T-dependence is a function of dT/dt. Such does not occur with xac - it
is fully reversible. If we continue to increase the temperature above Tf and
then cool the sample backdown, there is first the reversible ‘paramagnetic’
regime above Tf and then the susceptibility will become the reversible Xrc
(double-tipped arrows) with subsequent cooling. In short, the dc field (albeit
small), if applied below Tf creates a metastable, irreversible state, and Tf
is nicely defined by the onset of these irreversibilities, i.e., difference
between FC and ZFC curves. More about the low-temperature effects in
the next section.
How can we better analyse and interpret the data near T,? Let us begin
with the frequency shift of Tf. First of all we could try the Arrhenius law
for thermal activation already used for a superparamagnet
T = T, exp [E&u T] (3.29)
which we can rewrite as
w = w, exp [ - E,/k, Tf] (3.30)
Here w is the driving frequency of our xa,-measurement and Tf its peak.
By plotting In (w/w,) versus l/Tf, we can from the slope and value of the
logarithm determine E, and w,. For the canonical spin glasses using the
data in Fig. 3.14, best-fitting to the Arrhenius law gives completely unphysical
values for w, and E,, e.g. o, = 1020° Hz and E, = 4400 K. This nonsense
is clearly due to the very small changes in Tf (maximum in xac), and it
distinguishes a spin glass from a superparamagnet where the Arrhenius law
does indeed hold. There is much more than simple energy-barrier blocking
and thermal activation in the transition of a spin glass.
A second method of analysis is ferreted out from the literature on real
glasses. It is the empirical law which describes the viscosity of supercooled
liquids, namely, the Vogel-Fulcher law. Written for use in describing T,
shifts this becomes
w = w, exp [ - E,lk,( Tf - TO)] (3.31)
where T,, is a new parameter (for real glasses it is called the ‘ideal glass’
temperature). With three fitting parameters (wO, E, and TO) the agreement
is naturally much better, except perhaps at low frequencies, using a
3.2 Experiments spanning Tf 71
more physical set of parameter values. Typically for CuMn (4.6 at. %)
WI = 1.6 x lo8 Hz, E, = 11.8 K and rO(< T, = 27.5 K) = 26.9 K. The
problem here is the precise physical meaning of the large T,,. Some attempts
have been made to relate it to the interaction strengths between the clusters
in a spin glass. Since a spin glass is not a non-interacting collection of
clusters something must be added to take into account the inter-cluster
couplings. To might also be related to the true critical temperature of an
underlying phase transition for which Tf is only a dynamic manifestation.
Thus far, a deeper understanding of To and its connection to the longer-
range interactions is lacking.
Finally, a third approach draws on the standard theory for dynamical
scaling near a phase transition at T,. The conventional result of dynamical
scaling relates the critical relaxation time r to the correlation length 5 as
r - 5’. Since 5 diverges with temperature as 5 - [T/(T - T=)]“, we have
the power-law divergence
(3.32)
7AV=To (3.33)
and
x” = WT (3.36)
10 lo2 10’
Frequency (Hz)
QiTk(T) W 7) (3.37)
and
(3.38)
0.2
X’ (emu/mot Eu)
Fig. 3.18 Argand diagrams for three different temperatures. The frequency of the
ac driving field h increases from right (10.9 Hz) to left (1969 Hz). The lines are
computer fits to the data points assuming a symmetric diagram (see text). From
Hiiser et al. (1983).
74 Basic spin-glass phenomenon
T(K)
J lo-'2 10-'010-8 10-6 10-4 1o-2 loo lo2 log r (9
Fig. 3.19 Schematic representation of the probability distribution for spin relaxation
times with its evolution as a function of temperature.
3.2 Experiments spanning Tf 75
(9 Xnl = 1 - JC (3.40)
x3
where x0 is the linear susceptibility M/H in the limit of H + 0. Since
we can expand M in odd powers of the field
M = xoH - u&,H)~ + a&, H)5 - . . . (3.41)
where the as are a function of temperature. Then
xnl = +a3(xoW2 - dxoH>” + * * * (3.42)
Hopefully, our power series rapidly converges so we are left with
only one or two leading terms.
(ii) Alternatively, by applying an ac driving field, h, at frequency w, we
perform a similar expansion, but now as a function of odd frequency
harmonics 30, 5w, . . . .
where
@;=x;h+%X;h3+iX;h5+. . .
0;=;X;h3+;x;hS+.. .
1 X;h5 + . . .
0; = 16 (3.46)
and similarly for the imaginary part 0”. If we assume that the driving
field is very small (h G 1 in our expansion), then
M(w) = xih cos WC+ x’;h sin wt +
(3.48)
After all this action around Tf for the susceptibility, the specific heat is a
big disappointment. Already from Fig. 3.5 there is little to be seen at Tf.
As a matter of fact without further analysis Tf does not even seem to exist
in a C, versus T plot.
What can we do further to reanalyse or blow-up the specific-heat
behaviour? One suggestion for extracting Tf was to make plots for the most
3.2 Experiments spanning Tf 77
1 , I I I
10-2 lo-'
(T-Tg)/Tg
a
.Q*
l ..*
.
.- l ..v.,, l
4 *
c
‘;
0.0 ‘.‘.I. “.I. ‘.“..“I. .“““’
0 IO 20 30
TEYF’ERATlRE
(K1
Fig. 3.21 Spin-glass specific heat divided by temperature plotted against tempera-
ture. The vertical bars intersecting the plots mark the estimate of Tf for each
composition. From Martin (1980).
(section 3.1.2) was that most of the entropy (- 80%) is already frozen out
in short-range order above Tf. So conceivably there is simply not enough
entropy remaining to produce a significant peak at Tf. Recall how our
cluster configuration model nicely spanned the C, data through Tf (Fig. 3.7)
whereas it could not deal with the susceptibility peak.
Such conjectures are trying to preserve a phase-transition explanation for
C,. However, we must live with the empirical fact that little happens in
the specific heat. Either the phase transition is unconventional or maybe
we should mimic the real glasses and their freezing transitions (fall out of
equilibrium) as a viable model for the spin glasses. Before we choose let’s
look at some other experiments.
3.2.3 Resistivity
For usual phase transitions the critical behaviour of the electrical resistivity
is reflected in its temperature coefficient or derivative d(Ap)/dT. Theory
has even related this derivative to the specific heat and its critical divergence
for certain metallic cases. So it is not surprising that little can be found in
d(Ap)/dT at Tf. Figure 3.22 illustrates the point by plotting the T-coefficient
d(Ap)ldT for some CuMn alloys as a function of T. The arrows represent
3.2 Experiments spanning Tf 79
1 I I I I I I I I I
I n -.---.- Cu+S.'lot%Mn :
-----Cu+2.7ot%Mn ------ Cu+63ot%Mn '
-----Cu+l.Got%Mn - Cu+CSat%Mn 1
- Cu+O.7ot%Mn I
I \ I ---__ I I I II I
lOi w 40 50 10 20 30 40
! T(K) T(K?"
Fig. 3.22 Temperature dependence of the temperature derivative of the impurity
resistivity d(Ap)dT (rL? cm/K) for the CuMn alloys. The arrows represent Tf as
determined from the susceptibility measurements. From Ford and Mydosh (1976).
Cu-3at % Mn
i=
i= q = 0.13 2
I II ,, II 1
0 2s 50 75 100
T(K)
Fig. 3.23 The ‘elastic’ cross-sectionsmeasured with two different energy resolutions
along with the quasistatic part of the total cross-section (as defined in the text) for
the CuMn (3 at. %) alloy at q = 0.13 A-l. The dots, 0, represent the ‘elastic’
cross-section measured with energy resolution of 1.5 PeV. The triangles, A,
represent the elastic cross-section measured with energy resolution of 230 FeV. The
quasistatic part of the total cross-section is represented by the squares, W. From
Murani (1981).
Fig. 3.24 The possible forms of the density of relaxation times N(T) per unit
logarithmic interval of T, for the CuMn (3 at. %) alloy at (i) T * Tsg(=T,) and
(ii) T = Tsg. The regions marked a, b and c correspond to the lengths a, b and c
in Fig. 3.23 representing the spectral weights in the various time zones. From
Murani (1981).
82 Basic spin-glass phenomenon
(3.49)
~0) = (3.50)
exp (- II) dr
5(O)
=‘s Ij(r) dr (3.52)
which is nothing but our old friend ST or equation (3.26) in the limits
q -+ 0 and T + QJ.If g(I) is a slowly varying function, then we can define
a correlation function at relaxation time 7 as
5~~) = ‘g :hg(r) dr = s, (4 = 0)
I
See equation (3.24). Therefore,
I;(O) - e(T) = S&q = 0) - SZ (4 = 0) = % (4 = 0) (3.54)
where
So we can collect these results and plot the measured I& against log-
3.2 Experiments spanning Tf 83
where we have assumed a phase transition and t(O) = 1. Note how the
correlation function jumps up at Tf and develops its long-time tail (right-
hand side of Fig. 3.25). Once more there is the same type of experimental
behaviour and phenomenological model. At least we have an empirical
picture of the freezing process that is unlike that of the standard phase
transitions which exhibit a rapid critical slowing down and disappearance
of the fluctuations, a small critical regime where all this happens, and a
becoming infinite correlation length (long-range order). Using these various
techniques we have traversed almost ten decades of time - something
necessary for the spin glasses. Yet needed are probes of the really long
times which characterize the frozen state (T + Tf). Section 3.3 will handle
this situation, as for now we persevere with two further experiments which
are useful for ‘around T,’ behaviour.
QMn (5at.%,Tg=27.4K)
u.+ will process about any local magnetic fields (internal or external) felt
at its rest site. After 2.2 us the muon decays by emitting a positron along
the instantaneous direction of its spin S. One measures the asymmetry in
the direction of the emitted positrons. Although various configurations of
fields and detectors are possible, we, for simplicity, treat one important
case for the spin glasses, namely, that of a weak (LF+R), or even no
(ZF-l&R), longitudinal field, H L, applied parallel to the incoming beam
and thus HL((pL. The difference between forward emitted positrons and
backward emission is related to a spin-relaxation function G,(t).
We assume for a spin glass that there exists a lorentzian distribution of
random internal fields which fluctuate as a markovian process ee”’ (v is the
rate of fluctuation). This model introduces two parameters: ‘a’ the half-
width at half-maximum of the lorentzian field distribution, and v which is
taken as the inverse average spin-correlation time reflecting the fluctuations
of the local field, v = l/7,. For v/a % 1 (slow modulations or quasi-static
local fields)
Hence, the LF-l&R experiments can determine the two important para-
meters, a as a spatial distribution of local fields, and 7, as a sort of spin-
relaxation time. An advantage of this technique is that it can be performed
in zero external field. We remark that the time window available in uSR
to study relaxation processes is = 10-10-10-6 s.
With the above oversimplified parametrization we obtain results for G,(t)
in the different limits. Then T= can be extracted at each temperature and
plotted versus l- TJT, as is shown in Fig. 3.26 for various alloy systems
and concentrations. This figure illustrates the rapid upturn in 7, as Tf is
approached from above followed by a saturation very close to Tf. The
effective correlation time appears to move through the l&R-time window
with a power law dependence. Best fitting of the data shows that
(3.59)
&Fe
Fig. 3.27 Normalized Miissbauer-effect splitting versus T/T, for a series of AuFe
alloys. From Violet and Borg (1966).
88 Basic spin-glass phenomenon
3.3.1 Magnetization
Let us use one of the many magnetization techniques and zero-field cool
(ZFC) the spin-glass sample to T < Tf. We begin with sweeping the applied
field slow enough that the measurement is effectively dc, however, the
sweep rate is sufficiently rapid to avoid large relaxation or creep effects
mentioned previously. A generic sketch is given in Fig. 3.28. The small
initial slope near the origin is comparable with our ac-susceptibility, limit
(H + 0) M(ZFC)/H = xac, and before discernible time dependences occur
a small finite H is required. These usually appear around the start of the
S-shaped form in the virgin M versus H curve. Thus, already in small field,
the magnetization slowly drifts upwards roughly following a log-time scale.
The inset of Fig. 3.28 illustrates the upward creep for increasing field and
a downward one for decreasing field by a thick line.
As we reach very large fields (= 40 T), now shown in Fig. 3.29, there is
a tendency towards saturation, but full saturation Ms,, = g&S (- 4 FB for
CuMn) is never attained even at the lowest temperatures and concentrations.
Notice also how the magnetization at constant field decreases with increasing
concentrations (Fig. 3.29). What does this all mean? Well we certainly do
not have a superparamagnet which is easy to polarize! Instead, not only
ZFC
T<<Tf
‘H
Fig. 3.28 Schematic of zero-field cooled hysteresis loop for T 6 Tf illustrating the
S-shape, the lack of saturation, and the isothermal remanent magnetization (IRM).
The time dependences are depicted in the inset.
90 Basic spin-glass phenomenon
,gIY 14 at %
0 10 20 30 40
B (V
Fig. 3.29 Magnetization per Mn-atom for the various concentrations of CuMn
studied at 4.2 K; from Smith et al. (1979).
the long-range RKKY interactions, but also the peculiar anisotropy associated
with CuMn play a major role and prevent the external field from fully
rotating the frozen moments into its direction. One could try to calculate
the various exchange interactions based upon a superposition of three
Brillouin functions fitted to a universal (concentration scaled) magnetization
curve. However, for CuMn too much antiferromagnetic coupling is required
for a reasonable fit. The results totally disagree with our cluster configuration
model’s estimates of the .Jis from section 3.1.1 and Fig. 3.4. Since the
present magnetization versus field behaviour is a low-temperature T G Tf
measurement, there must be a lzew aspect to the frozen state. And this is
the random anisotropy which is formed below Tf and is particularly strong
in CuMn. The external field has to overcome an array of local anisotropy
axes, randomly oriented, before the various clusters can point along the
field direction and fully align, i.e., saturate, the spins. It takes field energy
to rotate the cluster moment away from its anisotropy-pinned ‘frozen’
orientation. Such an energy barrier for rotation was already introduced for
a superparamagnet, but now for a frozen spin glass the barrier orientations
are randomly distributed and the barrier heights, also a distribution, are
temperature dependent disappearing above Tf. We shall delve more closely
into this anisotropy when we consider the torque and ESR experiments
(section 3.3.5).
The AuFe system is somewhat different. The magnetization at constant
field increases with increasing concentration and saturation, while not
completely reached, appears a bit closer, here Ms, = &J+$ = 3 PB, while
3.3 Low-temperature (T Q Ta experiments 91
cu Mn
cu Mn qsqs D/.
D/.
T = 1,35OK
+loo-
+loo
Fig. 3.30 Hysteresis for field-cooled CuMn with saturated remanence magnetization.
Different symbols are used for different runs. From Monod et al. (1979).
92 Basic spin-glass phenomenon
t H,,H=O H<O
l TRM l TRM l =TRM
(4 (b) (4
Fig. 3.31 Memory effect as represented by energy-barrier deformations for the
field-cooled magnetization of CuMn.
3.3 Low-temperature (T e Td experiments 93
equal to zero and the magnetization measured. Here saturation of the TRM
is quickly reached for low values of H. Figure 3.32 illustrates this behaviour
for a canonical spin glass AuFe (0.5 at. %). For the TRM an irreversible
susceptibility may be defined as Xir = TRM(H)IH. By adding xir to x =
u?$(M/H), called the reversible susceptibility, xFc may be indirectly obtained.
From this procedure XFC is found to be constant for T 5 Tf - the result
agrees nicely with the direct method of simply FC a spin glass in field H.
The second type of remanence is called isothermal remanent magnetization
(IRM) and is obtained by cooling the spin glass in zero field (ZFC),
followed by cycling an external field 0 + H + 0, and then measuring the
magnetization. A more difficult saturation of the IRM is observed for
sufficiently large H values. Figure 3.32 also shows the IRM(H) with its
slower approach to reaching the already saturated TRM.
We have frequently mentioned relaxation and time dependences of the
magnetization within the frozen state. Let us now examine these in some
detail. With the presently available, highly sensitive SQUID magnetometer,
it is quite easy to detect very small changes (10-100 ppm) in magnetization.
What is more difficult is to prepare the sample magnetically for a meaningful
and reproducible experiment. We treat here a few simple well-defined
examples, which have created a great deal of interest and introduced a
totally new concept, viz., the waiting or ageing time, t,.,.
Magnetization of a spin glass will react in time to any change of external
field while in the frozen state. There are various routes to prepare a spin
glass for systematic measurements of these long-time relaxational effects.
Since the behaviour is irreversible and complicated by ageing processes (see
Au- Fe 0.5 %
TA 4
t
3-
I
I
1 t
*
TIME
I I
I I
IL----____-----___---__ I _t
‘TIME
Fig. 3.33 (a) FC-TRM MR(t) measurement: solid line T-profile, dashed line ff-
profile; (b) ZFC or T-quench M(t) measurement: solid line T-profile; dashed line
H-profile.
3.3 Low-temperature (T 6 TJ experiments 95
(a)
(b)
0
loo 10’ lo2 lo3 lo4 t(sec)
Fig. 3.34 ZFC susceptibility (lIZ-Z)M(t) and the corresponding relaxation rate S(t)
at different wait times, tw = 102, 3 x 102, 103, 3 X 103, lo4 and 3 X lo4 set plotted
versus log t: (a) (l/H)M(t), 5% of (1IH)M Fc is indicated; and (b) S(t), 1% of
(l/H)k& is indicated. CuMn (10 at. %), T/T, = 0.91, Tf = 45.3 K. From Sandlund
et al. (1988).
96 Basic spin-glass phenomenon
always increases, has an inflection point, and is a strong function of t,. The
latter behaviour is due to the ageing processes of the spin glass below Tf
which are continuing even though H = 0. These are then reflected in the
different time dependences of M(t) after each f,. The lower plot gives the
relaxation rate S(t) = (l/H)aM/a In t. Here the inflection point of M(f)
corresponds to a maximum in S(f) which shifts to longer observation times
(t) with increasing t,. The peak in S(t) occurs at a time f which is
approximately equal to f,. For the small fields used in the experiments the
ageing process proceeds unaffected by the field. Also for small H (< 20 gauss)
there is a linear-response regime where the data scales as M(t)IH and the
principle of superposition applies for different field-cycling procedures. For
example, the isothermal remanent magnetization (IRM) may be measured
by ZFC, applying a field after a waiting time t,, returning the field to zero
at t; and then tracking IRM(f). The same quantity can be calculated by
applying a +H at t, and superimposing a -H at t&, and then subtracting
the two measured M(t)s. Needless to say the agreement for the decay of
IRM(t) from direct measurement and this subtraction is excellent.
Various functional forms have been proposed to describe the magnetization
as a function of observation time and waiting time. One of the most popular
relations is the stretched exponential
M(t) = MO exp [-(t/tr)l-n] (3.60)
where il4, and tP (the time constant) depend upon T and t,, while n is only
a function of T. If n = 0, we have the Debye, single time-constant,
exponential relaxation. On the other hand, for n = 1, M(t) is constant. So
the value of n will critically govern the exact relaxation rate from very
strong to none at all. In order to verify such dependences sensitive
experiments must be carried out over many decades of observation time
and waiting time for a well-defined H-T procedure. Some illustrative
measurements, which have attempted the above, are displayed in Fig. 3.35.
Once again we have CuMn (5 at. %) and the FC-cycle has been used to
create a TRM. In fact, M,(t) is tracked for different waiting times and
temperatures over four decades of observation time (l-lo4 s). This is
indicated by the bold solid line in the figure. Next an attempt is made to
accommodate the data with a stretched exponential using a two parameter
(tP and n) fit of the above equation. The results of the best-fit are denoted
by the fine solid lines. While the agreement is good at short times, clear
deviations (see Fig. 3.35) commence as the time interval is extended. In all
four cases the stretched exponential function is insufficient to match the
data over a broad range. Something more is needed and this has been
provided by the Uppsala group, Nordblad et al. (1986). Superimposed upon
the stretched-exponential contribution is a purely logarithmic-decay term
Mk=SHlnt (3.61)
where S is the relaxation rate in ‘dynamical equilibrium’ which only weakly
3.3 Low-temperature (T Q Ta experiments 97
t w =lOOO set
t w =lOO set
Fig. 3.35 Relaxation of the remanent magnetization (MR) in the time interval
10m4< t < 108. Thick full lines are experimental data. Dashed lines are the
relaxation according to power-law plus stretched-exponential dependences. Thin full
lines represent the stretched exponent alone: (a) t, = 103 sec., temperatures 21 K
(top) and 25 K (bottom; and (6) t, = 10r sec., temperatures 21 K (top) and 25 K
(bottom); from Nordblad et al. (1986).
depends upon the time and waiting time. With the addition of equation
(3.61) to equation (3.60) the fitting is greatly improved as illustrated by the
dashed line in Fig. 3.35. Therefore, the time decay of MR(t) is logarithmic
for t Q t, and t + t,,,, and the ageing process superimposes a stretched-
exponential term on the relaxation around t = t,,,. At this point a bit of
theoretical interpretation is required to distinguish the quasi-equilibrium
dynamics represented by the log(t) decay at very long or very short time
scales from the non-equilibrium dynamics causing an ageing-time dependent,
stretched-exponential behaviour spanning t = f,. We shall defer such
explanations until the theoretical chapter (section 5.9) when we will introduce
a domain-growth model.
98 Basic spin-glass phenomenon
Au - I- at.% Fe
TEMPERATURE (K)
Fig. 3.36 Plot of the spin-glass specific-heat data fro AuFe (1 at. %). Note that
the linear part extrapolates to a positive temperature intercept. From Martin (1980).
3.3 Low-temperature (T 4 Td experiments 99
I
O.lev AE
E(X)’
Fii. 3.37 Two-level energy-barrier construction and density of states at very low
energies AE = El - E2.
100 B&c spin-glass phenomenon
for the excess ‘amorphous’ specific heat or for the randomly frozen spin-
glass C,,,. It was this connection with the real glasses via the similar specific-
heat behaviours that caused the word ‘glass’ to be attached to the random
magnetic systems whether metallic or insulating. So an approximately linear
specific heat is the low temperature signature of a spin glass.
Since the specific heat is in principle a rather difficult measurement, it
has not been fully used to explore the dynamic of spin glasses by introducing
a frequency or time of measurement into the technique. A related experiment
can be performed to determine the energy flux (or heat flow) dQldt for
T 4 Tf as the external field is changed. Here the spin-glass sample is weakly
connected to a reservoir by a constant heat leak. For both FC and then
setting H = 0 or ZFC and then turning on a field H, heat flows out of the
sample with a long-time relaxation (similar to the magnetization). At very
low T and for a given AH, the energy flux dQ/dt 0: l/t and the energy
Q = (const) In t. One can picture this behaviour as a lowering of the spin-
glass energy by a change of field, either removal after FC or application
after ZFC. Such, alterations drive the system slowly towards a new
equilibrium state. A more complete interpretation may be gained from the
two-level tunnelling system with a distribution of both energy-barrier heights
P(V) as in our NCel superparamagnetic model and splittings of the double-
well (nonequal depths) P(AE) caused by the field change AH (see Fig. 3.37).
Note that the tunnelling processes are essential at these low temperatures
(T < T,), since thermal activation (or surmounting the barrier) is mainly
ineffective. Therefore, we maintain our nexus among relaxation, energy
barriers and tunnelling, all participating in the quest for equilibrium.
3.3.3 Resistivity
At low T the magnetic resistivity of a metallic spin glass, Ap, has a large
residual (T = 0) component - a combination of the disorder of the alloy
due to random A and B site occupancies, and the spin disorder of the
randomly frozen state. We call this contribution xAp, since these disorders
are roughly proportional to X, the concentration. As the temperature is
increased
Ap(T, x) = xAp, + A(x)Ts (3.62)
where A(x) decreases very slowly with increasing concentration. We illustrate
this behaviour, which seems rather general, in Fig. 3.38 for a series of
CuMn alloys.
Some evidence has been offered for the appearance of a limiting T2
dependence at the very lowest temperatures (T < 0.3 K). However, this
must quickly crossover to the Ti behaviour for a larger T-range approximating
Tf. The interpretation of these A(x) and T%dependences has been discussed
in terms of long-wavelength elementary excitations which are diffusive in
3.3 Low-temperature (T -+ TJ experiments 101
- T(K)
1.5 3.0 LO 5.0 6.0 7.0 6.0 9.0
0.3
0.0
0 5 IO 15 20 25
- 1% fK-%
Fig. 3.38 The tfmperature variation of the resistivity for 4.4, 6.3 and 9.5 at. % Mn
in Cu versus Ti. The lines through the data represent a fit to this temperature
dependence. From Mydosh and Ford (1974).
What does inelastic neutron scatter say about the above conjecture? Usually
this technique provides a powerful probe for the study of collective spin
excitations. If we recall our dynamic structure factor S&q, w) given by
equation (3.17), then we could look for propagating spin waves, namely,
two inelastic lines at o = +o,. Hence,
or
Im x(9, 0) = (0 - w,)’ + I-2 + (w + $2 + r*
where r is the inverse lifetime of the spin wave which has a typical
dispersion relation wq = Dq* (with D the spin-wave stiffness constant).
Much experimental effort has gone into searching for these excitations in
the metallic spin glasses - all in vain. There are no sharp inelastic lines to
102 Basic spin-glass phenomenon
Cu Mn (fee)
I -
hzz 1oA
I - t- -II-+
+-2oii ‘-I gz208 etc. _....
SR-SDW (incomm.)
Atomic-SRO
-‘j
-,) Ferromagnetic
clusters embedded
with in SR-SDW
Fig. 3.39 Schematic of linear SDW propagation, upper part: long-range for YGd
and lower part: short-range for CuMn. From Mydosh (1988).
Fig. 3.40 Rotation of the anisotropy triad during a torque experiment. A rigid
rotation of the spin system is induced from 0 = 0 to 0 = 7~when H is rotated in
the y-z plane from OH = 0 to Or., = 7~. When a subsequent rotation cpis induced
in the x-y plane, the total rotation angle of the triad is still 7~with respect to the
0 = 0 starting position. However, a non-controlled irreversibility P rotation of the
anisotropy triad around a brings back the rotation angle to n - cp. From Alloul
(1983).
where u is used to represent M,, and 0 is the angle of rotation of h4, with
respect to the z-axis. The field angle OH is somewhat larger due to the
‘pinning’ of it4, by the unidirection anisotropy Kt.
Figure 3.41 shows the measured torque r, versus the field-rotation OH
in curve (a), while the calculated torque derived from equations (3.66) to
(3.68) is also plotted as a dashed line. Notice how the two curves nicely
agree until about OH = 50” (or 0 = 40”). The measured torque is here
reversible and independent of time. This means a rigid rotation of the
entire spin system is occurring with M,. However, at larger angles small
rearrangements are taking place with the corresponding irreversibilities and
time dependences. Nevertheless, the deviations of the experiment from the
calculated curves are not too bad (see Fig. 3.41) for the particular system,
and the remanent magnetization turns over for a field rotation of 180” and
comes back into phase with H.
Now what happens, if we rotate the applied field back to its original +z-
axis, not via a return rotation about the x-axis, but instead via a rotation
+ about the y-axis? This results in a r,, which is measurable and shown in
Fig. 3.41 as curve d. Here there is a large difference in the r,, values
compared to r, at least for the first 20” of rotation. Therefore, a return
rotation about the y-axis after an x-axis rotation is non-identical to a direct
3.3 Low-temperature (T -6 TJ experiments 107
. . . . . . .
*e---w_
. l.
,’ \
I’ \
,’ ‘\
I \
#’ \
#’ \
\
#’
8’
,’ \ ‘\
a’P “,
%(deg)
L
Fig. 3.41 Torque measurements in a CuMn 20% sample at 1.5 K. The torque r,
and r, are measured after a 7~rotation of u in the y-z plane. These two responses
are quite different and point out the triad character of the anisotropy. r,, is found
to be very weak for small cp as the total anisotropy energy is then cpindependent.
From Fert and Hippert (1982).
rotation from the FC z-axis about the y-axis and back. For, we must attach
a threefold orthonormal coordinate system or triud to our iI4, (or a) as is
done in Fig. 3.40. Note the dissimilarity between an x-axis followed by a
y-axis rotation and a simple y-axis, back and forth, rotation. Although o
always points in the same direction after these rotations the two non-u triad
axes are oriented in opposite directions along the direct and return (after
X) y-axis rotations. This coordinate-system difference or triad property is
manifested in the two unlike values of rY measured around 180”.
Our external field controls the rotation of the remanent magnetization
M,, but we cannot control the rotations of the collective rigid spin system
around M, - slippage will occur at large angles and the spin system will
revert back to its direct-rotation and lower-energy form and r,, = r,. Since
the random freezing is isotropic, our spin system comprises all orientations
in the 3D space, i.e., is non-collinear or Heisenberg-like. This means the
random local anisotropy is also distributed over all directions as an ‘isotropic
anisotropy’. When we probe the macroscopic anisotropy via a FC-field, we
establish the ‘memory’ direction ti of our fixed triad independent of the
108 Basic spin-glass phenomenon
crystal axes. Moreover, we have also induced M, and its triad vector fi
which rotates with M, as we turn i&ii away from the FC orientation by
applying and rotating the external field I-I. With the proper choice of
system, temperature and fields, we can only hope that the randomly frozen
spin system remains rigid and moves bodily and in phase with i&f, which is
being controlled by H. This seems to be roughly true for the x-axis rotation
in Fig. 3.41, but, if followed by a y-axis rotation, the spin system readjusts
or inverts two of its triad axis and returns to the torque and lower energy
of the original x-axis rotation. All these contrasting processes are illustrated
in Fig. 3.40 using the triad coordinate that turns with M, (or a). Note the
striking difference with a collinear (Ising) spin system where all of the
randomly frozen spins point either ‘up’ or ‘down’ along a given unit-vector
axis. For this ‘vector model’, unidirectional anisotropy only, there is no
difference in the spin system after the various axes rotations. It remains
completely the same no matter which path is followed. Thus, we would
expect to see a dramatic difference in the rotational properties between a
Heisenberg and an Ising (collinear) spin glass.
Another powerful experimental technique which can be used to study the
consequences of the spin-glass anisotropy is electron spin resonance (ESR).
Here one detects the spin resonance frequency of the Mn atoms. It has
long been known (late 1950s) that Mn impurities in a noble metal satisfy
the condition for ESR and many interesting physical properties can be
derived from such experiments in the dilute limit. But now we wish to
investigate the co-operative behaviour of the frozen Mn spin system for
T < Tf- For this purpose we can apply a hydrodynamic approach and
construct an appropriate free energy. Hydrodynamic theory represents a
collective approximation at low frequency and long wave-length using
macroscopic variables, and as a result it neglects the microscopic details of
the systems under consideration. For a canonical spin glass we have a non-
collinear, isotropic, frozen, spin structure with its magnetization, M(= xH),
a possible remanent magnetization M, and the unidirectional (K,) and
uniaxial (K,) anisotropies. Accordingly, the free energy (FE) becomes
where the notation is similar to before. The triad coordinates must be used
with N the FC direction and il the rotated ikf, direction. One can check
this FE by minimizing it with respect to 0 and obtain the equilibrium
rotation angle 0, after field cooling I& and rotating the external field H.
Substituting M, = XH + M,ii, as shown in Fig. 3.42, we arrive at the
equilibrium condition with KZ = 0:
sin 0, = MH
__K sin (0, - 0,) (3.70)
1
Fig. 3.42 Equilibrium rotation angle 0, for a field-cooled spin glass followed by a
rotation of the field H to an angle OH.
sketches the various vectors and their equilibrium angles. These conditions
have been verified by several magnetization and torque measurements as
long as the frozen spin system remains rigid; however, often such is not
the case.
For ESR experiments more is needed. Equations of motion must be
derived for the dynamical variables M and ii. This is usually carried out by
forming Poisson brackets of these variables with the free energy. Let us
consider a special case, namely, M,ii is parallel to H which is parallel to
H,. Now three resonance frequencies occur because of the three orthonormal
triad vectors. For a collinear (Ising) spin system there would be only two
resonances. The solutions of the eigenvalue equations for the (3) frequencies
are
(3.72)
if H 4 KJx. The first ESR experiments, shown for this special case in
Fig. 3.43, gives good agreement for the o+ mode (linear slope = $ and
expected intercept) even with resonance fields as low as 300-400 gauss. The
predicted situation for the three modes develops according to Fig. 3.44,
where the longitudinal mode oL remains at a tied frequency (no dispersion),
thereby making it difficult to find. This mode was finally detected indirectly
110 Basic spin-glass phenomenon
3300 - .i
-
t . .
U-J
: 3200 - .* /
L9 .
h
/
‘47 - aH, + Hi
2
a = 0.51 f 0.02
Hi = 3023 i IO
3000 -
2900 I I I I 1
0 200 400 600 800 1000
Field for resonance, H, (gauss)-,
Fig. 3.43 ESR (spectrometer frequency versus field for resonance) for a ZFC spin
glass, the resonance corresponds to the w + mode according to the equation given
in the text. From Schultz et al. (1980).
Fig. 3.44 A plot of the three expected resonance modes as a function of H for
zero remanence.
accord with hydrodynamic theory were mostly gathered, provided the spin
system rotated bodily, that is, stayed rigid. At large angles without a proper
preparation the CuMn spin system usually broke apart and along with it
the hydrodynamic description. The crowning achievement of all this low-
temperature effort was to demonstrate the co-operative nature of the spin-
glass phase. Clearly a new and unique state exists in the frozen spin glasses,
so we can conclude that there must be a type of phase transition which
transforms the system from the paramagnetic state at high temperatures to
the collective entity at T 4 Tf. And this state cannot be a mere assortment
of superparamagnetic particles or domains.
I5 - kilogauss
1 _
0.5- \ Tg = 37.4 K
-L T
o- 0 .l,. . x
10 20 30 40 50 60
8 1 I . , I
0 10 2o 3o 4o 50
T(K)
Fig. 3.45 Inverse of the FC-susceptibility (HIM) for AgMn (10.6 at. %) as a
function of temperature for various magnetic fields (indicated on each curve in
gauss). Data were obtained by slow cooling in constant applied field. The onset of
the ‘plateau’ (marked by arrows) is taken arbitrarily as the point of the M(T) curve
departing by 3% from its low-temperature value thereby defining Tf. The resulting
boundary of the spin-glass phase H,(T) is shown as an inset. From Monod and
Bouchiat (1982).
represented by the arrows in Fig. 3.45. These not only shift downward in
T with increasing H, but become greatly smeared, an estimate of which is
given in the figure’s inset where an H-T phase diagram is attempted, albeit
with large error bars. At larger fields Tf is simply not well defined.
If we use other measurement techniques, first of all there are significant
differences in establishing an H(T,) line between the methods and analytic
procedures. And secondly, the time scale of the measurement enters and
gives different forms for H( Tf) depending on the experimental time window.
Hence, the dynamics of the transition are playing the more important role
3.5 Experimentalids (intuitive) picture 113
and a static H( Tf) phase diagram seems impossible to generate based upon
general experimental criterion. When the strong influence of dynamics is
coupled with the large smearing effects on a so-called critical temperature,
the whole meaning of a line of phase transitions created by a field becomes
moot.
Fig. 3.46 Computer simulation for a random lattice of points (spins) with the
contours and their overlap representing the different ranges of spin-spin interactions.
From Verbeek et al. (1980).
This then would indicate that an equilibrium state has been reached.
Experimentally, it is very difficult to know if Tfo has been attained, one
can always try for another decade of time, and see if there is a small
change in a measurable quantity. Most likely, there is and this then
makes the question of an equilibrium phase transition unresolvable. How
can we rid ourselves of these dynamical processes to determine the basic
properties of the underlying phase transition? At present an answer has
not yet appeared.
Alloul, H. (1983) Heidelberg colloquium on spin glasses, Vol. 192 Lecture Notes in
Physics (eds J. L. van Hemmen and I. Morgenstern), Springer, Berlin, 18.
Brodale, G. E., Fisher, R. A., Fogle, W. E., Philips, N. E. and van Curen, J.
(1983) J. Magn. Mugn. Muter., 31-34, 1331.
Cable, J. W., Werner, S. A., Felcher, G. P. and Wakabayshi, N. (1984) Phys.
Rev. B, 29, 1268.
Cannella, V. and Mydosh, 3. A. (1972) Phys. Rev. B. 6, 4220.
Fert, A. and Hippert, F. (1982) Phys. Rev. Lett., 49, 1508.
Ford, P. 3. and Mydosh, J. A. (1976) Phys. Rev. B, 14, 2057.
Htiser, D., Wenger, L. E., van Duyneveldt, A. J. and Mydosh, J. A. (1983) Phys.
Rev. B, 27, 3100.
Levy, L. P. (1988) Phys. Rev B, 38, 4983.
Martin, D. L. (1980) Phys. Rev. B, 21, 1902.
Meschede, 0.) Steglich, F., Felsch, W., Maletta, H. and Zinn, W. (1980) Phys.
Rev. Lett., 44, 102.
Monod, P., PrCjean, J. J. and Tissier, B. (1979) J. Applied Phys., 50, 7324.
Monod, P. and Bouchiat, H. (1982) J. Phys (Paris) Len., 43, 145.
Morgownik, A. F. J. and Mydosh, J. A. (1982) Physica, 107B, 305.
Morgownik, A. F. J. and Mydosh, J. A. (1981) Phys. Rev. B, 24, 5277.
Morgownik, A. F. 3. and Mydosh, J. A. (1983) Solid State Commun., 47, 321.
Mulder, C. A. M., van Duyneveldt, A. J. and Mydosh, J. A. (1981) Phys. Rev.
B, 23, 1384.
Murani, A. P. (1978 a) Phys. Rev. Lett., 41, 1406.
Murani, A. P. (1978 b) .I. Appl. Phys., 49, 1607.
Murani, A. P. (1981) J. Mugn. Magn. Mater., 22, 271.
Mydosh, J. A. and Ford, P. J. (1974) Phys. Lett., 49A, 189.
Mydosh, J. A. (1988) J. Magn. Magn. Mater., 73, 247.
Nagata, S., Keesom, P. H. and Harrison, H. R. (1979) Phys. Rev. B, 19, 1633.
Nordblad, P., Svedlindh, P., Lundgren, L. and Sandlund, L. (1986) Phys. Rev. B,
33, 645.
Sandlund, L., Svedlindh, P., Granberg, P., Nordblad, P. and Lundgren, L. (1988)
J. Appl. Phys., 64, 5616.
Schultz, S., Gullikson, E. M., Fredkin, D. R. and Tovar, M. (1980) Phys. Rev.
Lett., 45, 1508.
Smit, J. J., Nieuwenhuys, G. J. and de Jongh, L. J. (1979) Solid State Commun.,
31, 265.
Tholence, J. L. and Toumier, R. (1974) J. Phys. (Paris), 35, C4-229.
Uemura, Y. U., Yamazaki, T., Harshman, D. R., Senba, M. and Ansaldo, E. J.
(1985) Phys. Rev. B, 31, 546.
References 117
Verbeek, B. H., Nieuwenhuys, G. J., Mydosh, J. A., van Dijk, C. and Rainford,
R. D. (1980) Phys. Rev. B, 22, 5426.
Violet, C. E. and Borg, R. J. (1966) Phys. Rev., 149, 545.
Wenger, L. E. and Keesom, P. H. (1976) Phys. Rev. B, 13, 4053.
Systems of spin glasses
For the sake of generality and completeness we should try at this stage to
collect our various systems of spin glasses. The previous chapters have, in
their generic approach, dwelt on a few very specific materials, e.g. CuMn,
AuFe and to a lesser extent (EuSr)S. Yet the spin-glass phase is, as stated
previously, a very general phenomenon. Over 500 different systems have
been claimed to be spin glasses, i.e., exhibit at least some (even one) of
the experimental characteristics mentioned before for the canonical (RKKY)
spin glasses. After ferro- and antiferromagnetism, spin-glass freezing is the
third type of magnetic ‘order’ or, better put for the latter case, co-operative
magnetic state.
While it is not the purpose of this chapter to consider all 500-plus
materials, nevertheless, we wish to list the most important spin-glass
categories and to distinguish their class according to the strength of the
magnetic interactions which are present. Remember the conduction-electron-
mediated RKKY interaction of the transition-metal impurities have the
strongest coupling. Oppositely, superparamagnets (rock magnets or Co0
particles) have little or no coupling and are, therefore, not spin glasses.
We begin with the transition-metal-solute alloys, work our way through
the rare-earth binary or pseudo-binary alloys and add some of the vast
number of multi-component, amorphous magnetic alloys. Then we move
onto the semiconducting materials for which there is an enhanced superex-
change. Such a coupling may be sturdy enough to create a good spin glass
at a reasonable temperature. And finally, we reach the magnetic insulators
where the superexchange is weak and, thus, the concentration must be
increased to promote a possible spin-glass state. Or, if the coupling is
insufficient, we return to our (super) paramagnetic, at least, down to the
lowest available temperature of measurement (see section 2.10). Here the
blocking of individual spins or clusters at T, would take precedent over a
cooperative and collectively frozen ground state which might begin to form
at Tf < TB.
4.1 Transition-metal solutes 119
(a) ‘good’ moment systems, meaning that the Kondo temperature should
be less than = 1 K, so that no complications are encountered with
weakening of the local moments at low temperatures; and
(b) a favourable solubility such that at least 10 at. % of the 3d metal may
be dissolved in the noble-metal host.
cu xs xs GOOD XS+T XS XT
& xs xs GOOD XS xs xs
Au XT GOOD GOOD GOOD XS+T XT
in this special subset we should include AuCo for which the Co solubility
is poor, but not as bad as CuFe. The difficulty here is with the very large,
few hundred kelvin, Kondo (or spin-fluctuation) temperature. As before
TK is strongly dependent upon the local environment, i.e., the number of
Co nearest neighbours (nn). In order to reduce TK to less than 1 K, a Co
triplet is required and then modified spin-glass behaviour is observed. For
this system an effective magnetic concentration of Co-nn triplets should be
used instead of the actual atomic percent.
A few non-noble-metal host systems may be grouped with the above
collection. Examples of which include ZnMn and C&In, whose crystal
structure, being hexagonal, adds a crystalline single-site anisotropy to restrict
the orientations of the freezing spins. This breaks the isotropic distribution
of frozen-spin directions discussed in section 3.3.5 and creates Ising
(up/down) or x-y (spins confined to lie in a plane) types of spin glasses.
Nevertheless, for these systems the solubility of Mn in Zn or Cd is rather
limited, in the thousand ppm range. A final extreme example of a local-
environment spin glass is Feo.3Al,,7 where the very large Fe concentration
is necessary to make the Fe magnetic, since Fe in Al has a very high spin-
fluctuation temperature of some thousands of degrees. Note that in all of
the above systems the competing ferro-/antiferromagnetic RKKY interaction
is active and results in a strong and oscillating exchange coupling.
By using the previous criteria of good moments and high solubility, binary
combinations of two transition metals may be used to create a spin glass.
That is, if a giant-moment system does not sweep away the random freezing
and leave a dilute, but long-range ordered, ferromagnet. Table 4.2 offers
a collection of the various possibilities. We begin with the giant-moment
alloys whose formation requires an exchange-enhanced host (see section
2.3). This means that, firstly, there is an abundant density of itinerant (d-
Host Impurity: Cr Mn Fe co
MO XT SG SG XS+T simple
Rh XT SG XT XT
Pd XT GM+SG GM GM exchange
Pt XT SG GM SG enhanced
states near the Fermi energy N(Er) such that the Pauli susceptibility is
large: x,, = 2&N(&). And secondly, an intra-atomic exchange, 1, should
be present on the host sites. This results in an exchanged-enhanced
susceptibility x = x&l - IN(E,)]-1 where the factor [l - IIV(E,)-1 = 0 is
known as the Stoner enhancement factor. For Pd, 0 = 10, while for Pt,
0 = 3. Consequently, Pd or Pt hosts alone will produce the giant-moment
polarization. As mentioned previously (section 2.3) only in the very dilute
limit of Fe impurities (ppm) in Pd will a spin-glass state form at very low
temperatures (< 1 mK).
An especially interesting situation is with PdMn where for x < 3 at. %
Mn, the giant-moment ferromagnetism prevails. However, upon further
increasing the Mn concentration (X > 4 at. %), the probability of having
two Mn atoms as first or second nearest neighbours increases. This then
supplies the essential element of competing exchange for the appearance of
the spin-glass phase. Since Mn-nn couple antiparallel, they thereby
produce the antiferromagnetic coupling which mixes into the longer-range
ferromagnetic, giant-moment polarization. By making a ternary alloy
PdMnFe and utilizing different concentrations of Mn and Fe we can
independently vary the ratio of ferromagnetic to antiferromagnetic exchange.
This procedure has become very important in comparing with theory and
studying the re-entry spin glasses of section 2.9.
Cr impurities, when introduced into the Pd or Pt host, locally ‘blot-out’
the uniform exchange enhancement, but in this process the Cr loses its
magnetic moment and becomes a weak moment with a high TK. Only when
the Cr concentration is large enough to provide a suitable Cr local
environment does a stable (or good) Cr moment appear, now however, in
a non-enhanced Pd or Pt matrix. Accordingly, the RKKY-oscillating
interaction can occur leading to a spin-glass state at low temperature. Here
there is no giant-moment phase, if anything, the term ‘dwarf’ or ‘destroyed’
moment is more applicable. CrFe, a combination of two magnetic elements,
represents another exceptional case. Fe when diluted into the Cr-bee host
first eradicates the antiferromagnetic (itinerant) spin-density wave. Yet, the
Fe moments remain intact, and again, due to RKKY interaction a spin-
glass phase appears around 15 at. % Fe. At slightly lower concentrations,
there is a re-entry transition back to the spin-density-wave antiferromagne-
tism. At slightly higher Fe concentrations, a series of transitions occur:
paramagnetic * ferromagnetic + (re-entrant) spin glass.
At this point in our survey of binary, crystalline alloys we should also
mention the various Ni-solute systems. Either in noble-metal or transition-
metal matrix, Ni moments require a local environment of other Ni nearest
neighbours. For PdNi three or more nearest neighbours are needed to
produce a giant-moment ferromagnet. For CuNi at least seven or eight
nearest neighbours give a narrow spin-glass regime before ferromagnetism
takes over. The all-important concept, used time and time again for moment
formation, is the local-environment model that a particular number of weak-
122 Systems of spin glasses
We now return to our two most-popular canonical spin glasses, AuFe and
CuMn. There has been an enormous amount of effort put into studying
these materials and it seems appropriate to gather the various experimental
results in a temperature-concentration phase diagram. According to Larsen
(1978), different characteristic temperatures are defined in the following
way:
Figures 4.1 and 4.2 display the many data points collected over the years
for AuFe and CuMn, respectively. The labelling corresponds to the above
definitions while the solid lines represent Larsen’s theoretical calculation of
the freezing T&x). Here the model relates Tf to the root-mean-square
RKKY interaction energy. Exponential damping of the RKKY interaction
is included to take into account a finite mean-free path via a parameter
r(x). Some salient features of these data, which extend over five decades
in concentration, are listed below.
Au c
.- Irn
bO
0
01
0
0
/
Fig. 4.1 Collection of characteristic temperatures for AuFe alloys over the complete
concentration range; from Larsen (1978).
The important conclusion of all this data plotting and theoretical analysis
is that the RKKY-interaction is the driving force behind the spin-glass state
of these many transition-metal alloys. The coupling strength is large, its
spatial oscillations create the mixed couplings of the randomly distributed
124 Systems of spin glasses
""1
IO4 10-5 10.' 10-a 10‘) 10-l 1
c
Fig. 4.2 Collection of characteristic temperatures for CuMn alloys over the complete
concentration range; from Larsen (1978).
way as to which model or solution is the proper and correct one. However,
in this chapter we push forward with our spin-glass menagerie.
The same game can be played with magnetic rare-earth elements: dilute
them into a non-magnetic host metal and let the RKKY interaction
perform its coupling. Nevertheless, there are some advantages and certain
disadvantages of the rare earths when compared to the transition-metal
solutes. First the good news, for most rare earths, with the notable exception
of Ce, there is no Kondo effect, because the coupling between local moment
and conduction electrons is a positive (or ferromagnetic) one. The rare-
earth elements, which do exhibit a Kondo effect (antiparallel coupling),
e.g. Ce, usually have a very low TK.
And now the bad news. In many cases limited solubility is a problem
one cannot simply alloy over a large concentration range, e.g. LaGd exists
as a solid solution for only a few percent. A way of overcoming this
difficulty is to use a pseudo-binary intermetallic compound, e.g. La,-,Gd,Al,.
Here there are a vast number of such combinations, a few of which are
listed in Table 4.3. Further bad news is that the strength of the RKKY,
i.e., its coefficient J, in equation (1.2), becomes much weaker for rare-
earth impurities. This smaller polarization of the conduction electrons results
in much lower T,s even at increased concentration of impurities. A final
complication with the rare earths is the crystalline-electric-field splittings
which are rather low in energy = 100 K. This means that the effective
moment will change as a function of temperature and Schottky anomalies
will influence the various experimental behaviours. On the positive side,
the single-ion crystal field may be used to generate a crystalline anisotropy.
Such, in turn, can create Ising (up-down) and x-y (spins confined to a
plane) types of spin glasses which should have different critical phenomenon
LaGd, LaEu
Binary
ScGd, ScTb, YGd, YTb and PrTb
Lal-,Gd,Alz, Lal-,CeA&
La,-,Gd,I,, La3-,Ce,I,
Lal-,GdxBb, Lal-,Ce,B,
Pseudo-binary
Lal-,Ce,Ruz, Th,-,Gd,Ru,
Cel-,GdxRu2, Ce,-,Tb,Ru*
Thl-,NdxRhz
126 Systems of spin ghses
Pseudo-binary
r; lo-
t"
F
5-
0 A
0 0.5 1.0
x
work is needed to sort out these nuances and to arrive at the fine (or
significant?) distinctions between a competing or mixed-interaction spin
glass and an only antiferromagnetic one. As we shall soon see theory only
treats the mixed-interacting spin glasses.
Chemical compounds with one magnetic element have long been a testing
ground for magnetic critical phenomenon. The chemists can almost at will
produce a material which has the desired spatial (l-, 2- and 3D) and
order parameter (Ising, x-y and Heisenberg) dimensionalities. Exchange
interactions can be varied (ferro, mixed and antiferromagnetic) with
superexchange and sometimes dipolar mechanisms. In the 1960s and 1970s
such pure compounds dominated the subject of magnetic phase transitions.
For the chemist there is no great difficulty in randomly replacing a
magnetic element by a non-magnetic one. Thus, in principle, we can create
multitude spin glasses. However, what is to be gained by discovering yet
another spin glass? There must be something particularly interesting or
‘ideal’ about the given material. As a case in point let us consider the
mixed compound Fe,Mg,-,Clz. Pure FeCl, is a hexagonal-lattice layered
compound which has ferromagnetic a-b planes (with a triangular lattice)
that stack antiferromagnetically along the c-axis; for this reason we have a
long-range ordered antiferromagnet. A strong uniaxial anisotropy aligns the
moments along the c-axis so the material is Ising-like. The in-plane
interactions are ferromagnetic for nearest neighbours, but antiferromagnetic
for next-nearest neighbours (nnn) with J2/J1 = -0.13. In diluted Fe,Mg,_,Cl,
one expects for x I 0.4 a random site, 3D, Ising spin glass which should
evolve out of an ordered antiferromagnetic structure. Very interesting here
is the coexistence (at intermediate x = 0.45), proven by neutron-scattering
experiments, of an infinite antiferromagnetic network with a collection of
randomly frozen spins in the various a-b planes. At lower X, until two-
neighbour percolation cuts off the spin-glass phase, the usual competing-
interaction spin-glass behaviour is observed, but now in an Ising system,
i.e., predominantly along the c-axis are there magnetic responses (freezing).
Hence, this material is a good candidate for studying the dynamics of an
Ising random-site spin-glass.
Another example is the Fe,Zn,-,F,, also an antiferromagnetic Ising
system, which for x 2 0.5 represents a prototypical 3D random-exchange
Ising model (REIM). In the presence of a field applied along the Ising axis
it exhibits crossover phenomenon to random-field Ising-model (RFIM)
behaviour. The above two models (REIM and RFIM) represent unusual
critical phenomenon and crossovers of different types of phase transitions,
albeit both to a long-range-ordered state which will be discussed in Chapter
7. Now what happens when the concentration is reduced? Before attempting
132 Systems of spin glasses
We revisit this question and try to pose an answer with respect to the
different amounts of magnetic coupling between the randomly distributed
magnetic entities. Table 4.6 collects, in summary version, the results of the
Table 4.6 Examples of differently coupled random magnetic systems
TM I
II ,6
.u_
= z
2 RE I z
{
2
<
Amor. I ; I
2 Semi. cn
0
t;
2 Insul.
5
s Part.
z
0
Exchange or dipolar coupling
previous five subsections. A rough estimation is given for the range of the
exchange coupling in the different classes of materials which we have
encountered. The interactions are basically RKKY versus superexchange.
We list the material entries according to the two types of magnetic elements,
3d and 4f, that go into the RKKY random alloys. Here the 3d-impurities
are noticeably stronger than the 4f ones. The amorphous magnetic metals
follow with their similar expanse of interaction strengths. Then come the
not-so-dilute magnetic semiconductors with a more narrow range comparable
to the rare earths. The insulators appear next, occasionally the superexchange
can be rather large, but it is too often only antiferromagnetic. Finally, we
have the particles or glasses, which possess little or no coupling energy and
thus can be ignored with respect to their spin-glass properties.
So once again a spin glass is a random, mixed-interacting, with sufficient
strength, magnetic system characterized by a random, yet co-operative,
freezing at a well-defined Tf below which a highly irreversible, metastable
frozen state occurs without the usual long-range magnetic order. As this
chapter has demonstrated, there are myriad systems which satisfy this
definition. And we have been calling such systems the good spin glasses
where the special case of canonical denotes the infinite-range RKKY noble-
metal alloys upon which so much experimentation has been performed. The
word ‘ideal’ we reserve for those materials which closely conform to a
specific theoretical model as we shall enumerate in Chapters 5 and 6.
Nevertheless, there are other materials, discussed previously, which lie in
the grey area, and when further progress is gained through more experimen-
tation they may possibly be shifted into the good spin-glass category.
REFERENCES
ground states. The cause of the instability of the SK model lies in the
replica-symmetric order parameter and a replica-symmetry-breaking (RSB)
scheme must be incorporated into the theory. At this point, an unconven-
tional, yet mean field, order parameter arises with fundamentally new and
subtle physics. We try to dwell on the pictorial representation of RSB and
its physical meaning. Afterwards, we move onto the dynamics of the mean-
field model and the time-dependent interpretation of the order parameter.
We hope these seven sections will present an introductory overview of the
mean-field theory (only a first approximation) for the spin-glass transition.
Remember the theory took more than five years to evolve and at least a
thousand publications to elucidate.
The remaining sections of the chapter summarize two more recent
treatments of spin glasses, namely the droplet model which is diametrically
opposed to the RSB standpoint, and the fractal-cluster model which closely
resembles our experimentalist’s interpretation, yet can quantitatively derive
the results. Then, since all of the above models and calculations have
usually relied on the Ising simplicity, we briefly consider non-Ising spin
glasses and what kinds of new effects can be expected. Finally, the all-
important computer simulations of spin glasses are described and the various
results collected for comparison (in Chapter 6) with experiment.
In the old days (1960s) before the name ‘spin glass’ was coined, a number
of theoretical ideas were afloat helping to explain random magnetic systems.
Four of these early approaches are now recapitulated, particularly because
we have already utilized them in the previous chapters. So we begin this
theory chapter with a bit of history by collecting the four notions and
briefly outlining their physical content.
The concept of concentration and range scaling for the RKKY interaction
was exploited by Blandin (1961) to derive some universal properties of the
noble-metal alloys. All thermodynamic quantities follow from the partition
function
Z= Trexp(-X/&T) (5.1)
where for an RKKY system, recall (1.2), the Hamiltonian can be
approximated with leading term and external field H
Jo COS (2kFrij)
(2kFrij)3 si ’ % - @BHxsi (5.2)
- not the lattice,so defined that its volume always contains a constant
number of impurity spins, xR%. If we divide this invarient by the constant
volume V, determined by the number of conduction electrons N, we can
write
xR%x(k&)”= const or (kFRc)3 m i
vo- 37r2N
using
(5.4)
Therefore, the first (RKKY) term in our Hamiltonian (5.2) also scales with
the inverse concentration, and after dividing by x it becomes an invarient
with respect to (rq/R=). Since all terms scale with x-l, the partition function
may be expressed in scaling form as
H )I
zzzT””
X
P’=-kn~lnZ=-k~,ln[,(~,~)]=-knx2~(~,~) (5.6)
where Q is the scaling function of the virial expansion. Fc2) and Fc3) may
be calculated in certain limits and, accordingly, various prognostications are
acquired for the specific heat, magnetization and susceptibility. These are
only valid at very low concentrations and rather high temperatures for
diverse values of the applied field. By studying sufficiently dilute magnetic
alloys, a number of these predictions have been verified. For example, the
magnetization should approach saturation in a large field as
M=M,,,(l-y) (5.8)
138 Models and theories
where R, is used to limit the range of correlated spins, and, due to the
previous scaling arguments, varies like (x))*; therefore, contains a constant
number of impurities. The final distribution P(H) is a convolution of
reaction-field distribution with the cavity-field distribution and results in a
broadened, more lorentzian-like shape with a significant value at H = 0.
For the Heisenberg case all 3D orientations are possible and we must use
a 3D random walk. But now the probability density P(H) (note the vector
H), which is constant near H = 0, should be integrated over the appropriate
volume element 47rH2 dH to obtain the scalar molecular-field distribution
P(H). This gives P(H) + 0, as H + 0, and the full distribution is shown
in Fig. 5.1 a. Here P(H) begins as H2 and reaches symmetric maxima at
+H,. Thus, most of the moments are aligned in finite field (H,I. Nevertheless,
for the Heisenberg case there will be a transverse response to a field H
and this creates elementary excitations (localized magnons) of the random
system. Employing numerical simulations of randomly distributed spins
coupled via the RKKY interaction, the density of these excitations at
various energies may be estimated. This N(o) is shown in Fig. 5.1 b from
the computer results. Note that now a finite density of excitations is found
as w + 0. Such an N(O) # 0 requirement is essential for the agreement of
5.2 The Edwards-Anderson model 139
H
Fig. 5.1 (a) Expected shape for the internal (molecular) field distribution P(H)
for a Heisenberg model in d = 3; note the initial Hz dependence. (b) Density of
elementary excitations from the computer simulation for an RKKY interacting
x = 0.3 at. % fee lattice. From Walker and Walstedt (1977).
How can we describe the sudden random freezing of a spin glass? In 1975
Edwards and Anderson (EA) proposed the following picture. Each spin Si
140 Models and theories
The model system has quenched disorder, i.e., the impurity degrees of
freedom are rigidly frozen, meaning there is no change in the randomness
‘.2 The Edwards-Anderson model 141
)f the spin sites (sample structural disorder is frozen in), only the spin
orientations can vary. For such a system one must average In 2 over the
listribution P(J& which is a difficult task. However, it is relatively easier
o average the partition function raised to some power Z” where IZ is an
nteger. Here the so-called replica trick can be employed
(Z”), = Tr exp
CS3 [
- i X{x, Sy}IRBT
a=1 1 (5.18)
W)c = c
{SPJ
I_‘:[(rl)
I,
dJij P(Jij) exp SPSF
II (5.20)
The integral over Jij can be performed for the gaussian P(J,) by completing
he square
rhis four-spin product in the exponential arises from the addition and
ubtraction of a term
(5.22)
o complete the square. Now the free energy may be obtained in the mean-
142 Models and theories
(5.23)
and
q(T+ Tf) = - ;
theorem as
(5.25)
where all i # j terms have to be set equal to zero (no long- or short-range
correlations). Since ((S&)o = 1 and ((SJc)o = q we obtain
Substituting the limiting results for q(T) and since H = 0, we write for the
ac-susceptibility
xa,(Ts
(&wBY
TJ = 3k,Tf o(Tf - T)2
and
(5.28)
-Y
0’
0.5
Fig. 5.2 (a) susceptibility and (b) specific heat versus reduced temperature calculated
from the EA model for two spin values; from Fischer (1975).
where a mean, J& has been included for the possibility of ferromagnetism
in the gaussian function. Infinite-range interactions require the scaling of
the variance and the mean according to A’ = A/N+ and JA = JOIN, so that
both our new A and JO are intensive quantities. Thus
Repeating the ‘replica trick’, i.e., calculate (Z”), instead of (ln Z),, as with
the EA calculations and after a lot of mathematics, which nobody likes to
show, SK arrived at a rather complicated expression for (Z”),, the
configurational-averaged, replica partition function. Substituting this
expression into the replica equation, we obtain
frdy(““‘, (%Trdx’..’
B
5.3 Sherrington-Kirkpatrick model 145
q=&/exp($]tanh’[;z+$]dz (5.35)
m=&/exp($]tanh[sz+g]dz (5.36)
Hence for given ratios of JO/A, q(T) and m(T) may be calculated and a
magnetic phase diagram is thereby established. Figure 5.3 shows this T
1.25 - Para
Ferro -
0.25 -
Fig. 5.3 Magnetic phase diagram for Ising spins interacting via an infinite-ranged
gaussian distribution of exchange forces with variance A and mean J,; from
Sherrington and Kirkpatrick (1975).
146 Models and theories
versus J,/A plot for Ising spins interacting via an infinite-ranged gaussian
distribution of exchange forces centred at Jo with width A. Note the three
possibilities for phase transitions which are predicted from the SK model:
(i) paramagnetic * spin glass; (ii) paramagnetic + ferromagnetic; and
(iii) double (or re-entry) transitions paramagnetic + ferromagnetic + spin
glass.
The differential (or ac-) susceptibility may be computed from the q(T)
function using
[l - q(T)1 (5.37)
X(T) = kBT - JO[l - q(T)]
By including the applied field H from our original hamiltonian in all the
SK calculations (something we omitted for simplicity), we can also obtain
the field dependence of x(T). Figure 5.4 exhibits the susceptibility behaviour
for J,lA = 0 and 0.5 with and without a field H = 0.1 A. Once again,
conforming to experiment, we have a cusp in the ac-x which becomes
rounded and shifted downward in a dc field. But, when the specific heat is
calculated according to the same energy-derivative procedure used in the
EA model, there is also a cusp in the predicted C,(T) at Tf. For
T > T,, Cm = NkBA2/2(kBT)Z, hence a tail in C, 0: l/T2 persists to higher
temperatures. (Remember from the virial expansion treatment of the specific
heat at high temperatures (section 5.1) C, CCl/T.) The SK result is in stark
contrast to the usual mean-field-theory conclusion for a pure system where
C, = 0 for T > Tf. The leading term for the spin-glass specific heat at low
temperatures is proportional to T. However, the entropy S, when determined
O.OO
I
0.00 1.00 1.50 2.00
kT
rheoretically the first warning that something was wrong with the SK
solution came from the negative, low-temperature entropy. Other difficulties
)ecame apparent with the free energy which turns out to be a maximum
vith respect to q for the solutions q = 0, T > Tf and q # 0, T < Tf.
tioreover, the q = 0 solution, if analytically continued below Tf has the
ower free energy than the spin-glass state (q # 0). These results are
jpposite to what is expected for a conventional second-order phase transition.
Finally, after a few years, de Almeida and Thouless (1978) performed a
letailed analysis of the SK solution and showed it to be unstable at low
emperatures both in the spin-glass and ferromagnetic phases. According to
he SK solution the spin-glass susceptibility is really negative. This is not
mly in contradiction with experiment, but also with the correlation-function
#usceptibility which is positive definite. In the presence of an applied field
H # 0) the instability line of the SK-solution for the spin-glass phase
:xtends all the way from Tf down to 0. Such behaviour is shown in Fig. 5.5
vhere we plot the H-T line (also called the AT line) that gives the stability
imits of the SK solution. The functional form is
Tf - T,,(H) = 3 f H :
(5.38)
A 4
O( a 1
‘or Tf = TAT(H) and an Ising spin glass. This becomes exponential in -H2
IS T + 0. An experimental consequence is that below the stability line
rreversibilities in the magnetic properties should appear. As we know from
3hapter 3 these irreversibilities are indeed a decisive characteristic of the
rozen spin-glass state.
The reason for the instability has to do with treating all the replicas as
ndistinguishable. Remember SK set q = qup in (5.33) - a replica symmetric
,olution, and such an assumption leads to an invalid solution of the mean-
ield EA model. Despite these theoretical difficulties, the SK model offers
148 Models and theories
I I
1.00
0.00
0.00 0.20 0.60 1 .oo
kT/A
Fig. 5.5 H-T phase diagram (or AT line) illustrating the stability limits of the SK
solution for the case of J, = 0; from de Almeida and Thouless (1978).
Let us recall what the usual mean-field equations for a magnet (ferro or
antiferro) are
5.5 The TAP approach 149
where the thermal averaging of the hyperbolic tangent has been replaced
by the thermal average of Sj, namely, (f(Sj))T = f((Sj)T). And such is the
standard mean-field approximation. However, for a spin glass we must
subtract from the ‘effective field’ given above a reaction term. This reaction
field represents the influence of the ith spin on the polarization of the
neighbouring spins at site j. A spin (SJ, produces a ‘field’ J,(SJT at site j
and induces a moment Jij(S&Xjj at that site (where xii is the local
susceptibility). The induced moment in turn produces a reaction field back
at site i:
The physical meaning of F -rAp in (5.43) is the following. The first term
represents the energy of the frozen spin glass, the second is the correlation
150 Models and theories
energy of the fluctuations which are reduced from the paramagnetic result
by 11 - (S&l f or each spin; and the last term gives the entropy of N Ising
spins possessing the mean values (S&e
At T 4 T,, TAP find that the order parameter behaves like
9(T) = 1 - or ; * (5.45)
0
which is different from the linear SK dependence 4 = 1 - a,T/A derived
from (5.35). Near Tf there is a saddle-point behaviour defining the accessible
region q0 1 q = 1 - (T/T,), below which divergences occur in the FTAP.
Thus, a linear temperature dependence of q occurs just below Tp According
to these q(T) variations we can calculate the low-T susceptibility as
(Xi)T = aiT/A which goes to zero as T + 0 (recalling Xsk remains finite as
T + 0 in agreement with experiment). The entropy S may also be
determined: S,,, = -aFTAPIaT, and in the low-T limit
fL,(T+ 0) = (5.46)
Since we know where the trouble lies let’s try to overcome the replica
symmetry by introducing a scheme that breaks it. After a number of
i.6 How to break the replica symmetry 151
Ittempts, the correct one was finally proposed by Parisi in 1979. His
lrocedure for breaking the symmetry of the order-parameter matrix was as
~0110~s. Consider an n x 12 replica symmetric matrix q0 (for illustrative
Imposes we take n = 8). We then divide the (n x n) matrix of elements
I,, into constant blocks [(n/ml) x (n/ml)] of size ml x ml. Along the
diagonal we introduce sub-matrix ql. The off-diagonal blocks are left
unchanged with elements qO. This process of dividing 8 x 8 matrices into
4 blocks of 4 X 4 matrices is portrayed in Fig. 5.6 and is only the first of
nany such steps. The division is again repeated along the diagonal:
n1/rn2 X ml/m2 (or 4) new blocks of m2 X m2 (2 X 2) sub-matrices are
:reated with order parameter q2 for the new diagonal ones. And so on (see
Fig. 5.6) the procedure is reiterated with smaller and diagonal blocks each
with its own order parameter qk. (In our 8 X 8 matrix shown in Fig. 5.6,
:he next iteration results in all zeros along the diagonals.) Finally, for a
nuch larger matrix the process is terminated after R such iterations with
:he smallest diagonal block of size mR X m& Throughout this construction,
:he successive sizes of the blocks are
n~mI~m2~~~~rnR~l (5.47)
which in the limit as n + 0 becomes ‘turned around’
OSmlIm2-~ -ImRI1 (5.48)
3y taking R, the number of iterations, to be very large, we obtain a
:ontinuous variation and so
$$+, BLOCKS
n x n OF $lZE m, x ml
8x8 4 BLOCKS
OF SIZE 4 X 4
Fig. 5.6 Replica symmetry breaking (RSB) scheme for qae with two levels of
breaking (n = 8, ml = 4; m2 = 2).
152 Models and theories
(5.52)
I 12345678 I
Fig. 5.7 Tree representation of Parisi’s RSB scheme. To find qms trace back along
the branches of the tree from cxand from p until they join: qae = q1 is the value
of q at this point.
Fig. 5.7, both of which are a direct result of ultrametricity. And this new
‘space’, called ultrametric, is a natural consequence of the replica-symmetry
breaking required to remove the defects of the SK solution
Returning briefly to experiment, let us try and see how this highly
sophisticated mathematical theory makes contact with what we have
154 Models and theories
measured on real spin glasses. The key information after the RSB has been
accomplished is contained in our now-continuous Parisi order parameter
q(x). This is obtained in the interval 0 to 1 via the matrix construction and
limit taking procedures
(5.54)
q-
q P
t
(L(.
% %I q
Fig. 5.9
156 Models and theories
1 x 9
”
%A
II) (I)
Fig. 5.9 Continued Behaviour of the order-parameter function q(x) and the
probability distribution of overlaps P(q) for various cases. (a) and (b) just below
T, in zero field: qM a T, - T and the initial slope of q(x) is roughly independent
of temperature. (c) and (d) at low-temperatyre T Q T, and in zero field: qM = 1.
(e) and (f) for small magnetic fields: q,,, 0: Hs. (g) and (h) at larger magnetic fields
near the AT line: (qw - qm) is proportional to the distance from this line. And (i)
and (i) in the replica-symmetry region T > Tf at non-zero magnetic field. Note the
value of qEA vanishes when H + 0. From Mezard et al. (1987).
xFc, not the ZFC or ac-x. Further (computer) computations show that
U(T = 0) reaches a negative value while S(T = 0) = 0, so we have removed
via RSB the unphysical negative entropy. Finally, the stability of the Parisi
solution has been carefully analysed and has been found to be at least
marginally stable.
In order to avoid the subtleties of the RSB approach we digress for a
paragraph and use two results of RSB in combination with the stable SK
solution (above the instability line) which we project into the spin-glass
phase. This is called the ‘projection hypothesis’ which begins with valid SK
mean-field theory. The two features of RSB are that x(T) = constant in
the spin-glass phase and that linear response is destroyed at low T.
Accordingly, we can assume for T < T,:
W, H> = W)
MT, H) = M(H)
0, W = dT) (5.57)
5.7 Physical meaning of RSB 157
I
RSB
F
0 1
T/T f
Fig. 5.10 Difference in the temperature dependence of the susceptibility as
calculated from the RSB-solution of the SK model and the SK-solution without
RSB.
3T2
HZ T2+2T:+
T2-T; ‘-*
1
forT’Tf (5.58)
M=; IHI...I
f[l--a+ for T = Tf (5.59)
smearing, and others clashing as with the specific heat. But much more is
needed in the interpretation of RSB.
Is there a simple physical picture for the replica symmetry breaking
(RSB) and the resulting ultrametricity? Well let’s attempt to sketch one.
During the freezing of a spin glass, the various spins can align their
directions in many different ways such that all these arrangements have
similar values of the free energy. Hence, a frozen spin glass possesses a
multi-degenerate ground state. The configurations which have lower free
energies correspond to pure equilibrium states. Other configurations with
higher free energies correspond to the metastable states. How do we then
characterize the ensemble of equilibrium state? Answer: use a weighting
probability w, for the state (Y to appear in the ensemble
w-p LF
i kBT p)
where F, is the free energy of equilibrium state 0~. Another question is
how does one equilibrium state (Y differ from a second one B. Answer:
define a ‘distance’ d& between these states as the difference in their average
spin values square
(5.63)
Here the EA order parameter q uA represents states which have the same
overlap with themselves, i.e., qaa = qpp = qnA. Thus
k$3 = %,A - qa& (5.65)
So we have arrived at a simple relation between distance and overlap via
(5.65).
Now we must consider the probability distribution P(q) of the various
overlaps among the pure equilibrium states
Let us examine a most simple example via this formalism, namely, the
ferromagnetic Ising model. In zero magnetic field at T < T,, P(q) contains
two delta functions at q = + m2 where m is the spontaneous magnetization.
If an up (or +) external field is applied, then only one delta function at
q = +m2 remains. For the former case there are two equilibrium states;
for the latter only one. Now we return to the spin glasses. At T > Tf one
delta function at q = 0 exists (the paramagnetic state). However, as T is
lowered below Tf in zero field, a delta function survives but along with it
is a smooth region of probabilities in between 0 and qw - see Fig. 5.9.
Here, when P(q) is not an array of discrete delta functions, we say that
the replica symmetry is broken. This usually occurs in a very smooth way
as the temperature is reduced through Tf. Smeared out transitions?
Consequently, there are a significant number (not just one or two) of pure
equilibrium states which compose the spin-glass ground state. And the
weight and overlap of these states generate the continuous function P(q).
Figure 5.11 gives a schematic, 1D picture of the free-energy landscape
that exists in multi-dimensional configurational space. The lowest-lying
minima represent the pure equilibrium states. At higher energies are the
many metastable states. Upon cooling, a spin glass may become stuck in
one of these states. If the system cannot explore all of phase space, then
we call it non-ergodic. RSB leads naturally to ultrametricity which in turn
is pictured by the multi-valley free-energy landscape. And when the barriers
between the valleys become infinite, ergodicity is lost. Note that the time
it takes to go from one valley to another is the exponential of the height
of the lowest saddle point between these two valleys. So it is not surprising
CONFIGURATION SPACE
Fig. 5.11 Multi-valley ‘landscape’ for the free energy according to the RSB scheme.
160 Models and theories
that there are many different relaxation times present, and we should now
move on to examine the dynamics of spin glasses within the SK and RSB
models. However, before we do, we must mention a final consequence of
the RSB. This is the lack of self-averaging.
Usually in a random system we can average over the extensive quantities
by dividing it into a large number of macroscopic subsystems, each containing
a different set of random variables {x}. If we then average over the
subsystems, we expect to obtain a result very similar to performing a
complete configurational average over all choices of {x}. Taking the
magnetization per spin as our extensive quantity, self-averaging may be
written as
M(x) - (M)c 3 0 for [N-spins] + uJ (5.67)
where the chosen set {x} occurs with reasonable probability. In other
words, one large random system gives the same result for the magnetization
as a configurational average over all choices of {x}. Clearly, when we
measure the magnetization on various samples of CuMn having the same
concentration, there are no differences in the experimental values, even
though each sample has a different set {x}. The configurational-averaged
magnetization is expressed as
where the summation runs over all the spin sites for a single large sample.
Although self-averaging is intuitive and seems almost obvious, it does
not work when a phase transition occurs which makes the system depend
crucially on the boundary conditions. For our spin glass the RSB solution
creates the existence of many degenerate ground states and this causes the
lack of self-averaging. Such quantities as (r& - q2 (overlaps) and
(PJ(ql)PJ(q& - P(ql)P(q2) (weights) do not equal zero as RSB calculations
have shown. Here the weights of the thermodynamic states are the important
factor. Extensive quantities depending on these weights will not be self-
averaging. This means different bond configurations will give different
results even for N + ~4. It turns out in RSB theory that the magnetization
is self-averaging, but the susceptibility is not:
5.8 Dynamics of the mean-field model 161
Since the SK-model considers Ising spins that have no natural intrinsic
dynamics, we must introduce time dependences artificially via Glauber
dynamics or the soft-spin gaussian-noise term. Near a second-order phase
transition time-dependent correlations take the form predicted by dynamical-
scaling theory. The autocorrelation function is defined by
(5.71)
where the thermal average ( )* can be replaced by an average over
sufficiently long observation times. Dynamical scaling predicts that the
decay of q(f) is governed by a characteristic time T which (power-law)
diverges at Tf. Accordingly, we can write
with the proper scaling function Q+(x) a exp (-x). Thus, for T > Tf and
ast+w
162 Models and theories
4(t)Oc
(T+t $exp[-t(T)‘] (5.76)
Tf
The scaling function reduces to q(t) a t-i at Tf. For H # 0 one can begin
with the dynamical-scaling form and define a critical temperature according
to [T - TAT(WYTATW). I-I owever, both the scaling functions and the
exponents are non-universal and vary along the AT-line. This suggests a
marginality of the transition in a field. By examining the dynamical
susceptibility near the AT line, theory predicts the frequency dependence
of x”(o) a o for T %=Tf crossing over to x”(o) a & at Tf.
Below the AT-line, instabilities also occur in the dynamics of the SK
model, so a new and different approach is required here. And Sompolinsky
(1981) has provided the method with a stable solution entirely within a
dynamical framework. We now write our continuous order parameter as
q(x) = ((Si(O)Si(Tx)),>, 0s X5 1 (5.77)
where (.)= and (.)o represent thermal and configurational averages,
respectively, and T~ represent a time in the broad continuous spectrum of
relaxation times. Note that T,.. % T,~, if x2 > x1. So as x + 0 we obtain the
very long relaxation times necessary to equilibrate the system. This gives
the statistical-mechanics order parameter
q = q(0) = lim lim q(t) (5.78)
N-SW f-vu
which Sompolinsky assumed goes to zero in zero field, i.e., all correlations
have decayed away. On the other hand, for x + 1 we have the short times
scales characteristic of the EA (or confined to one valley of Fig. 5.11) order
parameter
(5.79)
The time scales necessary for (5.84) to be valid may be far beyond our
txperimental reach of maximum lo6 s. Nevertheless, we can measure a
ield-cooled susceptibility which although not the true equilibrium one, does
ndeed mimic it in a field. The FC-field seems to drive the spin glass
owards its long-time equilibrium, and thus, XrC f constant at T 5 Tf. For
he intermediate cases we must return to the local dynamical susceptibility
neasured at frequency w, = l/t, according to (5.80) and explicitly calculate
w.
For our practical purposes the Parisi and Sompolinsky results may be
llaced within the model of RSB and benefit from its consequences.
rherefore, we can return to our rugged landscape of Fig. 5.11 and use it
o describe the spin-glass dynamics. Two main dynamical processes can be
Iistinguished in the frozen state. Firstly, confined to a single deep valley
tre the non-exponential-relaxation processes within one phase, i.e., the
ransversing of small bumps composing the floor of the ‘equilibrium’ valley.
n contradistinction, there are the hopping processes between the equilibrium
chases or from one deep valley to another. These appear on much longer
ime scales and they are estimated to diverge with the size of the system
IS log 7 - IV:. Now according to RSB, hierarchical distributions of time
scalesmanifest themselves and the hopping has taken place in an ultrametric
ipace.
164 Models and theories
Fig. 5.12 Schematic picture of the droplet model. A droplet of length scale L
(containing site j) has all its spins reversed (global spin flip) creating ground state
r. Outside the droplet the spins are aligned as in ground state r. The surface of
the droplet is fractal. From Fisher and Huse (1988).
5.9 Droplet model 165
When the autocorrelation function C(t) is derived, the above thinking leads
to an extremely slow logarithmic decay of temporal correlations
C(t) - (ln t)+‘* (5.86)
Further calculations within the droplet model give a ‘l/f-noise’ spectrum
proportional to l/olln oJ1+@‘JI) and real and imaginary parts of the
susceptibility
(5.89)
where t, is the total age of the system after the quench and f, is a
microscopic unit of time (to = 1 is the usual basis). This growth of domains
is the fundamental process of the long-time, non-equilibrium dynamics of
the spin glasses below Tf. A corresponding ‘quench’ can be performed by
turning off (H = 0) an infinite magnetic field at T G Tf. The magnetization
is expected to decay with the formation of domains according to
(5.90)
Let us treat the processes which occur in the two limiting cases: In t 4 In t,
and In t % In t,. For the first case we have
(5.92)
5.10 Fractal cluster model 167
where now ta = t,,, + t. For all practical purposes Rta will be constant in
this ‘early epochs’ regime, and since L,, or experiment, probes length scales
which are much smaller, we use the term quasi-equilibrium dynamics. In
other words, our experimental scale is so small that it does not see the
domain walls, only the pure states I and r. So to a first approximation the
relaxation processes are characteristic of equilibrium. And the magnetization
should decay as
-- H
m(t) (In t)W
with 8 and JI the equilibrium exponents. Figure 5.13 sketches the various
functions R’J’, L+ and the expected derivative of m(t) with respect to In t.
For the second case: In t % In t, (‘late epochs’), and therefore L, = R,
since t = ta = t + t,. Now the experiments are clearly seeing the domain
walls and thus the non-equilibrium dynamics are detected with
40
decay.
There are two crossover regimes of interest. The first is at In t = In t,
or where L, + R,, i.e., a transformation from quasi-equilibrium to non-
equilibrium dynamics. Here the measured magnetization decay shows an
‘S-shape which, when plotted as dm(t)ld(ln t), reflects a peak at t = t,; see
Fig. 5.13. The second crossover occurs in the limit of infinite time which
of course cannot be reached by experiment. At this point L and R have
reached macroscopic length scales so the number of domains and their walls
become insignificant. True equilibrium has finally been attained and the
magnetization and its slope have decayed to zero. In general, based on this
picture of the domains and their growth, we have a simple and qualitative
model for interpreting the long-time and ageing behaviours of the spin
glasses presented in section 3.3.1.
A scaling theory, which makes particularly close contact with our experimen-
talist’s interpretation (Chapter 3), is the fractal-cluster model of Malozemoff
and Barbara (1985). They proposed a scaling theory of the spin glasses by
considering clusters of correlation length 5 which diverge as [(T- Tf)/7’r]-“.
These coherent regions have a cluster size sE on which all relevant physical
quantities depend and whose volume is SD, where D is the fructuZ dimension
because the clusters are expected to be highly irregular and branched. The
clusters have a magnetization mg - si or s, if the given cluster consists of
168 Models and theories
w
InO, 1
MACROSCOPIC
/ ------- LENGTHS
/’
04
0 I 0 EOUILIBRIUM
to tw In(t)
Fig. 5.13 Growth of domain size R and experimental-probing-length scale L versus
logarithm of time (+ is the barrier exponent); (a) t, = t+t,,, the total age of the
system; (b) t is the time of measurement, tw, the waiting time, and to, a microscopic
time; and (c) relaxation rate aM/a In t versus In t. From Lundgren (1988).
(5.95)
(5.98)
S
= exp [ - (S/S&~] (5.102)
f(k)St
with T 5 Tf and sufficiently large clusters. In the long-time limit and
ignoring algebraic prefactors, the integration of the numerator gives
where n = x/(x + u) and t1 - se. This is our old friend, the stretched
exponential, which has been used to describe the slow relaxation of real
glasses and also certain regimes of spin-glass relaxation. If we repeat the
above calculations for M(t), but at T = T,, we can use the percolation-
theory result right at the percolation threshold: n,(T,) = s-‘l. In this case
the integral for M(t) becomes
D
M(t) a exp [- (t/t&-“] ds a : 6r (5.104)
0
Up until now we have been mainly considering Ising spin glasses. We will
now relax this restriction and briefly deal with m-component vector spins
Si, such that Si - Si = tn. The spin-glass order parameter is now a tensor in
spin space which for H = J, = 0 reduces to a simple scalar, i.e., the system
will be isotropic in spin space. And here one can repeat the standard
procedures of RSB and obtained modified solutions and properties for the
static and dynamical behaviours.
A more interesting case is the isotropic spin glass in non-zero field with
J, = 0. Now M = ((S$)c # 0 and qll = ((S~l’)T(S~l)),), # 0 for the
parallel (to the field) magnetization and spin-glass order parameters,
respectively. In addition, there are various transverse order parameters
which allow the possibility for transverse freezing
Fig. 5.14 Sketch of field-temperature phase diagram for the mean-field model with
vector spins: (a) freezing of transverse spin components with weak irreversibilities
(GT line); and (b) onset of strong irreversibilities with RSB (AT line).
172 Models and theories
the (GT) transverse freezing ifiT, 0: HZ, followed by the (AT) longitudinal
freezing 6Tf 0: H% RSB occurs at both lines and the transverse spin-glass
susceptibility is expected to diverge as [T- T&H)]-l for T + T&(H). If
we additionally let J, (the mean of the gaussian P(J,) distribution) be non-
zero, Gabay and Toulouse predict the more complicated ‘re-entry’ T-J,
phase diagram shown in Fig. 5.15. Here there is a set of transitions:
paramagnetic to ferromagnetic to the first mixed phase Mr where ferromag-
netic and spin-glass (frozen transverse components) phases coexist. Finally
Mi + M2 signifies another transition where a crossover occurs to a new
mixed state with strong irreversibilities as opposed to the weak irreversibilities
in Mi.
Continuing with the vector spin glasses we next discuss systems with a
uniaxial anisotropy. Here the hamiltonian becomes from equation (1.4)
Para
Ferro
/
Spin glass
0 1
Jo/A
Fig. 5.15 Phase diagram for vector spins interacting with infinite-ranged gaussian
distribution (JO is the mean and A the width or variance). M1 and M2 represent
mixed magnetic phases of the re-entry transition. From Gabay and Toulouse (1981).
.ll Non-king spin glasses 173
<SB theory setting J, = 0. The results are shown in Fig. 5.16. ‘P’ represents
he paramagnetic phase at high temperature, ‘T’ the transverse phase with
fll = 0 and ~7~ # 0, ‘L’ the longitudinal phase with qll # 0 and q1 = 0,
.nd a mixed phase, ‘LT’ is also present, where both order parameters are
lot equal to zero. A variety of different phase transitions is predicted with
wo possible types of entry into the mixed LT phase. Experiments, mainly
sing crystalline anisotropy as the source of D, seem to exhibit the L or T
ransitions, however, values of D are not accessible in real materials to test
or the LT phase.
We conclude this section by mentioning the theoretical conjecture that
he Dzyaloshinskii-Moriya (DM) interaction (see section 1.4) and its
orresponding random anisotropy in an isotropic (or Heisenberg) spin glass
an be converted by an applied field into an Ising spin glass. The small
leld reorganizes the random DM directions to make a component of them
loint along the field. This then establishes an Ising-like (longitudinal) frozen
lignment and changes the universality class from Heisenberg to Ising. With
urther increasing field the behaviour transforms to a new regime which
xhibits both a transverse freezing with RSB at higher temperatures followed
bya crossover to longitudinal freezing (strong irreversibility) as T is lowered.
he two H-T lines are similar to the GT and AT lines - see Fig. 5.14 -
nd note the different H(T) dependences. Experimental searches for these
arious lines and crossovers have had to rely on a precise determination of
he freezing temperature over a wide range of magnetic fields. As we have
‘lg. 5.16 Phase diagram of an infinite-range vector spin glass with uniaxial
nisotropy D; from Roberts and Bray (1982) and Cragg and Sherrington (1982).
174 Models and theories
Spin glasses are a natural choice for numerical simulations and the Monte
Carlo method lends itself readily for computational studies. These computer
experiments serve to span the gap between the measurements done on non-
ideal materials and the analytic calculations performed via approximate and
oversimplified theoretical models. But first what is the Monte Carlo (MC)
method as applied to a spin glass?
We pick a lattice in arbitrary dimension usually 2 or 3, but depending
on computer power we can go even higher, e.g. d = 5. Sizes of the lattice
or number of spins can reach 64d using fast special purpose machines
especially built for simulations of random spin systems. One then selects
the number of spin components: m = 1 Ising, m = 2 x-y, and m = 3
Heisenberg. Finally, we need a probable distribution P(J,) for the random-
bond configurations. This is customarily taken as the EA or SK gaussians
or simply +J. Although a more realistic model, the canonical RKKY-
random site ‘CuMn’ situation, has also been MC simulated (see Chapter
8). Once we have decided upon the desired spin-glass ‘system’ via the above
choices and with a given initial configuration of energy U,,, the game starts
by flipping one spin. If the energy change U1 - U, < 0, the spin flip is
accepted and we repeat the process for a second spin giving a U,. However,
if the energy change U1 - U, > 0, then we calculate the exponential
B = exp [-(UI- U,)Ik,T] and compare the result with a random number
R between 0 and 1. If B > R, then the spin flip is accepted (Metropolis
acceptance criterion) and we go on to a U,. If B < R, the spin flip is
rejected and we return to our original initial condition U,, and start all over
again by flipping another spin. Each time we attempt a spin flip, we set
the time scale with MC step/spin (MCS). This for a good equilibrium
ensemble average can reach lo7 or 108 MCS. For better averaging we
should repeat the calculations with 30 to 100 realizations {Jij} of the random-
bond configuration. In addition, the initial condition should be checked for
not influencing the results.
Such recursive procedures give a nearly exact calculation of the partition
function Z{Jij} averaged over the various sets of {Jij}. MC simulations have
been employed from almost the beginning of the spin-glass problem. The
early computations (late 1970s) focused upon 2D lattices with Ising spins
(Glauber dynamics) and a gaussian distribution of near-neighbour bonds.
5.12 Computer simulations 175
(5.108)
(y and u are the usual critical exponents associated with a possible spin-
glass transition). Again using MC simulations these quantities may be
numerically computed. And in a similar manner just about any quantity of
physical or theoretical interest could be calculated for the various models
of spin glasses.
Rather than delve into the many such MC experiments we limit ourselves
to discussing the main and accepted (by the MC experts) results. Remember
these numerical simulations are a middle route between theory and
experiment, and they are most useful as a test or proving grounds for one
176 Models and theories
or the other. Hopefully, as is the case in simpler problems, all three should
agree.
In huge simulations on a specially built processor Ogielski and Morgenstern
(1985) considered the Ising, +J, 3D spin glass. The static properties showed
a sping-glass ordering transition at Tf = 1.175J with critical exponents for
the correlation length u = 1.3 and for the non-linear susceptibility y = 2.9.
Working within the mean-field RSB approach, Bhatt and Young (1985)
compared order-parameter distributions P(q) computed in different ways.
A finite-size scaling function was derived from the moments of P(q) which
should be independent of size at Tf and then splay out again at lower
temperatures. The various size curves do indeed cross at T = 1.755 for
d = 4. In the case of d = 3 they come together at T z 1.21, but stay
merged as T is reduced below Tf. This indicates that d = 3 is very close to
the lower critical dimension dl below which T, = 0. Additional scaling fits
yield y = 3.0.
In another series of computer simulations called domain-wall renormaliza-
tion-group techniques, Bray and Moore (1984) and McMillan (1984) further
studied the dimensionality of the spin-glass transition. In this method one
computes the variation of the ground-state energy 6E, by changing the
boundary conditions from periodic to antiperiodic in one direction. For a
spin glass (6E.J = 0, but (6E2,); and (ISE,I) are not zero. So if the energy
change decreases with system size, then there is no order because at some
length scale SE, = kBT, and this will disorder the system. The results
indicate order for d = 3, but not for d = 2.
Naturally, these techniques can be repeated for isotropic (Heisenberg)
vector spin glasses with short-range *J or gaussian bond distributions in 2,
3 and 4 dimensions. Even the long-range RKKY interaction may be included
in the simulations. The results suggest that Tf = 0 for d = 2 and 3 isotropic
spin glasses and that dl = 4 (above which Tf should be finite). Also the
long-range RKKY interacting spin glasses are at, rather than below, their
lower critical dimension.
MC simulations (Ogielski, 1985) have also been used to determine the
spin-glass dynamics via the behaviour of q(t), the dynamic correlation
function defined above. Figure 5.17 exhibits how q(t) decays in time (MCS)
at various temperatures spanning Tf. Note the system size is 323 = 32768
spins and about 5 x lo7 MCS have been performed. Once one knows the
q(t) dependence various analytic fits can be made. For example, above Tf
the functional form is
q(r) = C,t-” exp [ - (wt)a] (5.113)
where both exponents x and B depend on the temperature. Above the
Curie temperature of the non-random 3D Ising model (T, = 4.55) x = $
and B = 1. While from T, down to Tf (the Griffiths phase) j3 is between
0.4 and 1 indicating that the time decay is a stretched exponential or
Kohlrausch function. Below Tf it changes to power law
5.12 Computer simulations 177
2.
IO'
IO0 J-L&l_ I1 III
10-I IO0
(T/T,,-I)
Fig. 5.18 Power-law fits for the relaxation time T versus (T- 7’,)-‘” from the
Monte-Carlo simulations for lattice size 643 and T 2 1.30; from Ogielski (1985).
Statisical mechanics
Phase transitions
Amit, D. J. (1984) Field Theory, the Renormalization Group and Critical Phenomena
McGraw-Hill, New York.
References
Now that we have toured the main spin-glass experiments, listed the various
materials and scanned the highlights of the theory, how close can we relate
theory to experiment on the most ideal systems? We do not wish to make
a detailed comparison of every aspect, but chiefly to show which models
can be mimicked by experiment and whether their predictions are valid
especially regarding the freezing transition. Remember we have our
experimentalist’s rendition of the basic phenomenon: growing and freezing
clusters with a broad and shifting distribution of relaxation times. Theory,
if it is correct, can take us much further with a quantitative description of
experiment and the predictions of new effects. Once we gain a sufficient
rapport between the two, experimentation and theory, the problem is solved
and the physics is said to be understood. However, for the spin glasses the
full solution has not yet been attained. There is no complete quantitative
comparison or detailed set of predictions directly relating to experiment.
Great progress has indeed been made and is continuing with the various
models and calculations, particularly the computer simulations. On the
other hand, the majority of the measurements have been performed on
systems which are too complicated and greatly differ from the simplified
theoretical models. So how can you expect to compare experiments on a
complex material with an idealized theory?
In this chapter we adopt the point of view that real spin-glass systems
must be made or chosen to resemble intimately the theoretical model. And
these special materials we call the ideal spin glass. We distinguish an ideal
spin glass from a canonical (RKKY) or good spin glass - these generic
terms were frequently used in the previous chapters - as follows. An ideal
spin glass completely conforms to the theoretical model under consideration.
In contrast, a canonical or good spin glass exhibits the main experimental
features of a spin glass exposed in Chapter 3. There is no question about
the latter being a spin glass with a sharp freezing transition. However,
complications are present due to its universality class, deviations from
randomness, departures from pure RKKY behaviour, higher spin values,
orbital moments, undefined anisotropy, etc. Such aberrations add to the
182 Ideal spin glasses: comparisons with theory
The many canonical or good spin glasses, which exist in nature and possess
the prerequisite experimental criteria, are not at all ‘ideal’, i.e., closely
conform to the basic theoretical model (see below). For example, most of
the natural spin glasses are Heisenberg systems with site randomness. Very
many, especially the metallic alloys, are severely complicated with local
chemical clustering that leads to atomic, and in turn, to magnetic short-
range order. While the RKKY interaction is long range, it is greatly
modified by the magnetic clusters and may even taken on different distance
dependences for real metals. Another possibility is that a truncated spin
density wave will form. Molecular-field theory is inappropriate to treat
these effects.
Other spin glasses do not have fully localized moments and are intricated
with spin fluctuations or the Kondo effect. Still more materials are unknown
regarding their exchange interactions and, in the case of metals, are
arbitrarily defined as proper RKKY systems. Nevertheless, despite these
difficulties, such systems as (canonical) CuMn and AuFe or (good) (EuSn)S
and a-GdAl, are often assumed to be ideal and compared in great
detail to Ising-model, random bond, mean-field theory. Insulating mixed
compounds also have their difficulties with only antiferromagnetic exchanges
or complicated crystal structures and Ising systems are usually hard to find.
We now propose to create an artificially structured spin glass which will
fully satisfy the theoretical models and their computer simulations. Figure
6.1 shows a 2D projection of our ‘ideal’ spin glass. The ‘material’, which
also could be 3D, is an insulating Ising system with good local moments
(S = !J occupying regular lattice sites of a simple crystal structure. In Fig.
6.1 the magnetic sub-lattice is square and the spins can point only up or
down (Ising). An interpenetrating sub-lattice is available for non-magnetic
elements. Here we randomly distribute two types of ligands which control
the superexchange coupling. The exact ligand configuration between two of
6.1 What is an ‘ideal’ spin glass? 183
The crystal structures of both RbzCuFd and Rb2CoF, are the K,NiF,-type,
meaning a body-centred tetragonal sub-lattice of magnetic Cu or Co atoms.
Since the spins are separated by such a large distance along the long or c-
axis, the compound is magnetically quasi-2D. In the decoupled (u-b) planes
a simple square lattice of strongly interacting moments is formed. Rb,CoF,
is an archetypal Ising (up/down spins along the c-axis) antiferromagnet with
T, = 103 K. Rb,CuF, possesses a small x-y anisotropy which causes the
spin to lie ferromagnetically aligned in the u-b plane, Tc = 6 K. The
magnetic interactions are superexchanged via the F-ligands and fall off very
rapidly such that the next nearest-neighbour coupling is only 1% of the
nearest neighbour one. Thus a nearest-neighbour exchange is fully sufficient
to describe the magnetic behaviour of this system.
Now instead of mixing ligands to create random bonds as in Fig. 6.1, we
substitute Cu for Co which effectively mimics bond randomness. The reasons
for this are as follows - recall section 1.35. The Cu2+ ion has two types
of ground state orbitals whose lobes alternate along the x and y axes. As
nearest neighbours two Cu ions are ferromagnetically coupled and their
lobes are always orthogonal. But a Cu lobe pointing towards a no-lobe Co
neighbour (remember the square-lattice symmetry) generates an antiparallel
alignment. For a Cu lobe perpendicular to its Co neighbour, a parallel spin
pair results. Finally, Co-Co nearest neighbours are always ferromagnetic.
A sketch of the lattice with its ‘random’ and competing bonds is given in
Fig. 6.2. Notice we have two ferro ( t t ) and two antiferromagnetic ( t 4 )
interactions. These can be varied by changing the concentration x, and for
a particular x = 0.29 the resulting P(J) was already shown as part of the
introduction to spin glasses in Fig. 1.8. In essence we are now repeating
what has been previously stated in section 1.3.5 for a random-bond spin
glass.
Fig. 6.2 Representative distribution of ferromagnetic (F): Cu-Cu and Cu-Co and
antiferromagnetic (AF): Co-Co and Cu-Co exchange interactions in RbZCul-$ZoxFd.
Also indicated are the in-layer d-orbital lobes of Cuz+. From Dekker (1988).
6.2 Rb2Cu1&oxFq, an ideal 2d spin glass 185
The Co, although in the minority, plays a very interesting role. Because
it experiences a tetragonal crystal field, Co*+-ions have only their lowest
Kramers doublet occupied. Thus its effective spin is i, exactly the same as
the Cu. Furthermore, the single-ion anisotropy of the Co*+ is such as to
create a good Ising, uniaxial-oriented system. Last but not least, when
single crystals are grown, and this procedure is not without difficulty, they
form without short-range chemical order. So there is a true distribution of
random bonds unhampered by any atomic clustering. Consequently, the
Rb2Cu1-$oxF4 compounds represent an Ising, spin i, random bond, short-
range (-+J), square-lattice spin glass. This set of adjectives closely conforms
to our ideal sketch in Fig. 6.1 with the exception of the symmetric +J
distribution. For the real case there are 2+Js and 2-Js, and, while the
relative probabilities Z’(J) (see Fig. 1.8) may be adjusted by varying x, the
asymmetric magnitude of the four Js is an aberration from our ideahzation.
The next step is to optimize the x-value and to demonstrate that the
system is not long-range ordered but indeed a random freezing occurs at
some temperature. In the present 2D case we expect Tf = 0. Experiment
establishes the spin-glass regime to exist for 0.18 < x < 0.40, with a good
working concentration x = 0.22. The usual measurements of ac- and dc-
susceptibility, magnetization and specific heat are carried out to prove the
absence of long-range order, i.e., a featureless specific heat, and the
beginning of a random freezing via the time (frequency) and field (FC
versus ZFC) dependences of the other quantities.
Once the spin-glass properties have been firmly established and a T-x
phase diagram mapped out, we assume a phase transition transpires at
some T,. Since we are being guided by theory which would predict a critical
temperature T,, we have changed our notation from Tf to T,. The static
critical behaviour can be studied through the non-linear susceptibility xnl.
Here a dc method in different applied fields (vibrating sample magnetometer)
was used to determine x,,i taking advantage of the magnetization expansion
indicated in equation (3.41).
(6.1)
0.7 I I I I I I Ii I I I I I I I I I I I
~1.67 K
v
~2.30 K
h -
0.6 -*2.75 K 0
03.05 K A
l 3.35 K .
-03.65 K v
2 0.5 A .
l 3.95 K 0
x 04.25 K p
.- VA .
s 0.4 0
v d
z
PA . .
e 0
:: 0.3
t VA
v b .
. 0
. 1
4
I A l
. I
b 0.2
.-iz
f
2 0.1
0.0
III III I I I I II I I I I 11 I
-0.1
0 50 100 150 200
Magnetic Field (G)
Fig. 6.3 Non-linear susceptibility xnl = l-(M/x,H) versus the internal field for a
selection of temperatures. Solid lines are fits to determine the a-coefficients. From
Dekker et al. (1988).
6.2 RbZCul--xCo,Fq, an ideal 2d spin glass 187
where the scaling function G(x) goes to c,x2 in the limit x + 0 and to 1
- c,x-“~ in the limit x + 03, with A = 1 + ‘2 (-y+B). By plotting x,T-P
versus H/TA at a variety of temperatures, excellent overlap is obtained -
10
Y- 6
A-
5
C.
“x 4
0 1 2 3
T, 09
Fig. 6.4 Results for x2 test (best-fit for T,), y, and 2y + l3 versus T,, obtained
from fitting the non-linear susceptibility with various values of ‘T,‘. Solid lines are
guides to the eye. From Dekker et al. (1988).
188 Ideal spin glasses: comparisons with theory
10”
T,=OK
7
A = 3.2 , /e?= 0.0
yx 10-l if
l- l
$ ‘4.10 K
,A 0 3.95 K
4 =3.80 K
a* ~3.65 K
- 4v
,s’ ~3.35 K
03.50
4
l 3.20 K
v I I 1111111 I 11111111 I I llll~
lo-'
10-l loo 10' lo*
xoH/TA-'
Fig. 6.5 T, = 0 scaling plot of the non-linear susceptibility data. External fields
range from 1.0 G to 12.5 kG. From Dekker et al. (1988).
6.2 RbzCul-$o,F4, an ideal 2d spin glass 189
I 1 I / , , ,
(a)
0-
012345678 012345678
Temperature (K) Temperature (K)
Fig. 6.6 (a) In-phase linear susceptibility ~‘(0, T) for Rb&u0.,&00.218F4 versus
the temperature. Data were taken at frequencies 0/27~ ranging from 0.3 Hz to 50
kHz, and with the driving field along the c axis. Solid lines are guides to the eye.
(b) Same as (a), but out-of-phase linear susceptibility x”(o, T). From Dekker et al.
(1989).
190 Ideal spin glasses: comparisons with theory
(6.7)
3 4 5 6 7
Temperature (K)
Fig. 6.7 Relaxation time T, versus the temperature. Solid line is a fit to the
activated-dynamics equation. From Dekker et al. (1989).
6.2 RbzCul-,CoxF4, an ideal 2d spin ghs 191
where B is the barrier height (IJJis its exponent) and 5 m t-” for a zero
temperature critical point. Thus, according to theory T, diverges much faster
with temperature than a power-law divergence. Returning to Fig. 6.7, we
have displayed the fit to the above equation by the solid lines with +I.J = 2.2,
70 = 2 x lo-l3 s and B = 11 K - a physical set of values. Other fitting
procedures were attempted, e.g. power laws with T, = 0 or T, as a
parameter. But in all cases poor x2-tests and unphysical values of the
parameters resulted. The best agreement with experiment is clearly via the
activated dynamics of the droplet model with critical exponent IJJU= 2.2.
One can also use the Cole-Cole analysis to generate the form of the
relaxation-time distribution g(~). Figure 6.8 exhibits the evolution of g(~)
as a function of temperature. Notice the many decades shift with T especially
in the long-time tails. These enormous displacements and broadening are
all occurring far above T, = 0, at a few kelvin. Once again the dynamics
is dominant and such huge changes within the measurement window between
5.8 K and 3.4 K could easily and incorrectly be taken for finite temperature
phase transition.
Alternative methods exist to derive T, from the spin-correlation function
and by use of the fluctuation-dissipation theorem:
do = &w&)> (6.8)
and
0.06
0.05
0.04
0.03
0.02
0.01
0.00
“Log(T) (-r in s)
Fig. 6.8 Distribution of relaxation times g(T) according to the Cole-Cole analysis
for a selection of temperatures. From Dekker et al. (1989).
192 Ideal spin glasses: comparisons with theory
(6.10)
and
Edwards-Anderson
SG model for d = 2
complete description for this ideal spin glass and a most favourable
experimental-theoretical agreement. Certainly more work is needed to
reduce the error bars to justify that the four real Js of Rb2Cuc7sCo,,22F4
are definitely in the same universality class as the simple ?J model. And
finally upon viewing Fig. 6.7 again, the warning is clear. Far above a
possible T, the relaxation shifts outside the realm of feasible experiment.
Here the spin glass is out of equilibrium and we must live with these
disturbing non-equilibrium effects.
With this in mind, let us move onto the available 3D materials where
the question of a finite T, is paramount and, if so does T, = T,? And what
about the behaviour below T,, once a phase transition has been established?
Keep in mind we are using T, to denote the critical temperature of a phase
transition as predicted by theory. In contrast we reserve Tf for the freezing
temperature as is clearly characterized by experiment. The latter may or
may not be related to the former.
shifts the exponents upwards and T,, which is taken as a fitting parameter,
downwards. Furthermore, the ‘best’ exponent values obtained to date are
anomalous, i.e., much too large when compared to those observed in
normal phase transitions. These points when taken collectively, indicate
that a 3D spin glass is close to or just above its lower critical dimension.
In addition, we have all the complications of non-equilibrium setting in as
we approach T,, so that even if there is a phase transition, we could not
measure its equilibrium properties.
Rather than belabour the above obstacles (or shall we say challenges),
let us consider the present situation regarding a possible ideal, or model-
conforming (see Table 6.1 and dilate the dimensionality from 2 to 3), 3D
spin glass. To date our ‘best choice’ lies with the Fe,,,Mn,,STi03 mixed
compound, although strictly speaking it is not as ideal in 3D as the
Rb&uCoF, is in 2D. For the former compound we have a random
substitution of Fe and Mn spins situated in a hexagonal lattice; see Fig.
6.9. In the basal plane the Fe-Fe nearest neighbour superexchange is
ferromagnetic, while the Mn-Mn one is antiferromagnetic. However, it has
recently been verified that the Fe-Mn coupling is also antiparallel. Between
the planes, the interplanar coupling is always weakly antiferromagnetic as
shown in Fig. 6.9, and this gives the system its 3D character. So for a
50-50, Fe-Mn composition, there are three -Js with large weight and one
+J creating the short-range interactions. This represents a most asymmetric
distribution P( +J) of exchanges along with a large amount of frustration
due to the ‘missing-centre’ hexagonal (see Fig. 6.9) crystal structure.
Moreover, the Fe and Mn ions have different values of spin, both not 4,
which makes the system seem random-site-like due to the distinction
between Fe and Mn ions. In spite of these difficulties a strong unaxial
anisotropy keeps the spins aligned (up/down) along the c (or hexagonal)
axis, so Fe,,SMn,,,TiO, is indeed a 3D Ising short-range spin glass. And
this conclusion has naturally been substantiated by the usual measurements
of ac- and dc-susceptibility, FC versus ZFC magnetization, time dependences,
etc. These experiments have nicely mapped out the (T-x) phase diagram.
Equal amounts of Fe and Mn lies right in the middle of the spin-glass
regime.
Let us proceed with the scaling analysis of this quasi-ideal 3D spin glass.
By defining the non-linear susceptibility in a similar manner as in section
6.2, we expect the coefficient a3 to diverge as (T-T&Y. One then plots
xnl versus Hz (the small applied dc field) and in the limit of H,, + 0, i.e.,
the initial slope a3 is acquired at the various temperatures. Then a3 may
be plotted on a log-log scale against t = (T- T,)/T, for various choices of
T,. Figure 6.10 shows such dependences where xn, m u3, and Tg = T,. Note
there are two parameters here T, and y. A ‘best fit’ is curve B, T, = 20.70 K
and y = 4.0. The reduced temperature interval is very small because,
according to FC-magnetization, the system loses equilibrium below = 22.0
196 Ideal spin glasses: comparisons with theory
C-axis
0: F&&f+ FeTio3 MnTi03
t
(4 (b)
(cl
Fig. 6.9 (a) Crystalline structure of ilmenite-type compounds in the hexagonal
lattice. (b) Magnetic structure of FeTi03 and MnTi03. (c) Arrangement of magnetic
ions in the hexagonal c-layer. From Ito et al. (1988).
(6.13)
where F(x) a x for small x and F(x + m) a x1/8 with 6 another critical
exponent. This scaling function is displayed in Fig. 6.11 as a log-log plot.
The data ‘collapsing’ is obtained using the previously determined values for
T, and y. l.3 was the only remaining parameter and a value of 0.54 was
optimized from the scaling overlap. It should be noted, however, the
6.3 Ideal 30 spin glasses 197
B C
. .
.\ .
TBW)
A 21.00
B 20.70
C 20.50
Fig. 6.10 Log&-x2) versus loglo for three different values of Tg = T,. The best
fit, indicated by a solid line, is obtained for Tg = 20.70 K, yielding y = 4.0. From
Gunnarsson et al. (1991).
-3
3 4 5 6 7 8 9
log IO[ H 0’ / ++-’ (G2)]
Fig. 6.11 LoglO(<~) versus logIO(H~ltP+~). The points represent the data
collapsing obtained using Tg = 20.70 K, y = 4.0, and p = 0.54. From Gunnarsson
et al. (1991).
superposition of data point is not very good for lower magnitudes of the
coordinates.
Via ac-susceptibility experiments the dynamical behaviour can be analysed
using the conventional divergence of T, = T,(T- T,)-‘“. Here for
Fe0.5Mn0.,Ti03 eight decades of frequency are available from x’(T) and
198 Ideal spin glasses: comparisons with theory
(6.14)
Since we expect llzv and plzv to be much less than one, the above equation
considerably weakens the frequency dependences. Now a linear scaling plot
of x”T/wP’~” versus (T- T,)/w~‘~” can be made and this permits a much
better judgement of the scaling quality. Such refined analyses result in
T, = 20.5 K, zv = 11.5 and l3 = 0.58. Significant here is that as always
when a more sophisticated scaling analysis is made T, is reduced and the
critical exponents are increased.
As a final comparison we use the static form of this improved scaling
analysis to treat the non-linear susceptibility for Fe0.5Mn,,STi03. With the
same rationale as above in mind, we write
(6.15)
where the scaling function d(x) + 1 for x - 0 and G;(x) = X-Y for small
Ho. As before scaling plots are constructed by plotting on linear scales x,,,/
H iP’r+P versus (T- T,) TJH :‘+a. Figure 6.12 displays the scaling analysis
for a particular choice of parameters. The inset is an enlargement of the
scaled data around the upturn. Unfortunately, it is impossible to find a
uniquely optimum set of parameters (here T,, y and p) that gives a perfect
collapse of the data. Various sets of the three parameters produce equally
good ‘collapsibility’. T, can vary between 20.5 and 21.0, y between 3.6 and
4.3 and p = 0.5 to 0.7.
Now that we have summarized the latest attempts at scaling the most
ideal 3D spin glass, we should collect the results and relate them to the
theoretical predictions from the Monte-Carlo simulations of a 3D Ising
short-range spin glass (see section 5.12). Tables 6.3 and 6.4 list the salient
features and results. Regarding the ‘idealization’, notice in Table 6.3 the
inconsistency with respect to the different spin values and the distribution
of interactions. The unlike crystal structures should not be of great
significance, except that the hexagonal lattice will increase the amount of
frustration and strongly reduce the magnitude of superexchange coupling
-J4 along the c-axis.
Table 6.4 requires a close examination. Here we make a direct comparison
6.3 Ideal 30 spin glasses 199
0.3
0.06
_ t Tg=20.70K
0.2
- '. ;=I=,",",
*
I
0.1 t
i
Fig. 6.12 x~JH~@‘(~+@) versus t/Hz(Y+a). Th e points represent the data collapsing
obtained using Tg = 20.70 K, y = 4.0, and l3 = 0.54. The insert figure displays a
magnified part of the plot. From Gunnarsson et al. (1991).
Fel-,Mn,TiOs EA Model
Table 6.4 Comparison of critical exponents obtained for Fe0.5Mn0.5Ti03 with those
of the 3D king +J Monte-Carlo simulations
P(T). Around T,, where the simulations may not even be valid, l3 = 0.5.
So this is a deficiency of the MC technique, namely that l3 cannot be
accurately determined. Returning to the real experiments only 3 exponents
can be directly evaluated via the diverse scaling analyses of the measurements:
y, the susceptibility exponent, p, and ZY, a product of the dynamical
exponent with the correlation-length exponent. These are shown with the
appropriate error bars in Table 6.4 In order to continue the comparison,
different scaling relations (equations connecting the various critical
exponents) must be used to calculate 6 (the magnetic field exponent at T,),
cx, v, T (the correlation function exponent), and z. With this set of exponents
we should be able to completely characterize a second-order phase transition.
In general, the agreement is at best fair. Particularly disturbing are the too
large values of y and v for our real (ideal) 3D spin glass; also the inability
of fixing T, to within 1% weakens the analysis. While it seems that there
is a phase transition, it is not of the conventional type. Either the lower
critical dimension is close to 3, e.g. as some theories predict 2.7, so that a
marginal phase transition occurs, or the non-equilibrium behaviour which
is certainly present hinders a meaningful characterization of the critical
phenomenon. At this time we cannot proceed and resolve these difficulties.
One may study Table 6.4 and draw his or her own conclusions about the
nature of the spin-glass freezing. To be sure we must await future experiments
and theory we can offer a definite and physical settlement. Perhaps such
will be possible in the second edition of this book.
Never before has so much effort been put into the study of a possible phase
transition as with the spin glasses. The previous sections have only scratched
the surface of huge amounts of work done on the problem. For we have
5.4 Future possibilities 201
theory of spin glasses and despite the past successes new approaches, not
so heavily based on second-order phase transitions, and models from other
areas should be tried ensuring future progress not just in spin glass, but
more extensively with random and glassy systems in general.
The previous six chapters have dwelt upon the spin-glass phenomenon with
emphasis on the word ‘spin’. Except for a few cursory remarks pertaining
to real glasses, we have remained fully on the track of magnetic effects and
spin models. Naturally, magnetism does have its advantages since we can
easily couple into such systems with a magnetic field, the conjugate field
of the problem, and there exists a vast range of sensitive experimental
techniques to measure this coupling under various conditions. However,
now that we have outlined the basic spin-glass behaviour and theory, and
hopefully have gained an understanding of the rudimentary physics, the
obvious question arises: are there non-magnetic analogues? Answer: yes,
many! So let’s at this point take a look at some of these and venture
beyond the spin and even beyond the glass into life sciences.
We begin with the electrical analogues where electric-dipole moments
giving rise to a polarization P replace the spin moment causing the
magnetization M. Higher-order moments are also mentioned in the form
of electrical quadrupoles. By substitution of a moment bearing species for
a non-moment or a different-interacting moment, we introduce the elements
of randomness and competition. Thus, we can create electric dipolar and
quadrupolar glasses.
Our analogues are continued with various superconducting phases. First,
there are granular superconductors which behave like a network of randomly
coupled tunnel junctions, Then, the high-T, oxide superconductors have
been modelled on the spin glasses, initially because of their ceramic nature
and later due to the unusual field penetration (H,, < N < H,,) creating a
flux-line ‘lattice’ which becomes glassy.
In order to contrast the spin-glass phenomenon with a random magnetic
system which does exhibit a good second-order phase transition, we briefly
discuss the random-field Ising model. Here we focus upon the critical
behaviour and the H-T phase diagram whose critical line must be traversed
in a certain way in order to observe the phase transition.
Finally, we conclude the chapter with three non-physical analogues to
the spin glasses. These ‘spin-offs’ are combinatorial optimization, popularly
known as the travelling-salesman problem, neural networks or how our
7.1 Electric dipolar and quadrupolar glasses 205
FH
25 30 35 40 T (IO
10 20 30 40
T (K)
Fig. 7.2 Non-linear dielectric susceptibility E,,(T) and critical coefficient a(T) in
K1-,Na,TaO,. Solid curves: best fits to power law. Dotted curve: extrapolation of
a(T) to show criticality. From Maglione et al. (1986).
7.1 Electric dipolar and quadrupolar glasses 207
-wu 160
Fmq. Hz
0 106
. 3260 1 (a)
0 10640 fI
t l 33700
I I 1 I 1 I 1 I
0 10 20 30
TEMPERATURE (K)
Fig. 7.3 (a) The real part of the dielectric constant ELfor Rb,-,(NH&H,PO,. The
lines are guides to the eye. (b) The imaginary part of the dielectric constants E:
fitted via the lines to a broad distribution of relaxation times. From Courtens (1984).
208 Spin-glass analogues
Fig. 7.4 Sketch of the various molecules in a (1, 0, 0) plane for 50%
(KCN),(KBr),-,. The arrows on the cyanide molecules indicate the random freezing
of the dipoles. From Sethna and Chow (1985).
. :
.v .. :
1.2
2 0.9
g
d 0.6
2
w” 0.3
0.0
0 40 60 120
T 09
Fig. 7.5 ~1 (=e’) and Ed (=E”) versus temperature at several frequencies for
(KBr),.,(KCN),.,: o = 6~10~s~’ (A), 6~10~s~’ (m), 60-l (O), and 12s-’ (V).
From Birge et al. (1984).
210 Spin-glass analogues
which vanishes in the ground state of the solid phase. As a result of this
difference, o-HZ molecules are non-spherical, while p-H, molecules have
perfect spherical symmetry. Therefore, the asymmetry of the o-HZ offers
the possibility for various types of quadrupolar orientational order (dipole
moments are negligible) whereas the no-moment p-H2 can be used as an
inert host. Figure 7.6(a) illustrates the situation for the long-range-ordered
phase of pure o-Hz. A planar projection of the fee-like crystal structure is
shown with the novel type of periodic long-range orientational ordering at
low temperatures. More important for us is the mixed crystal where sufficient
amounts p-H2 have been substituted for o-HZ to form the mixed phase.
This is displayed in Fig. 7.6(b) where the p-Hz molecules are indicated by
the ‘dotted’ spheres and there are various shapes and orientations for the
o-HZ molecules. The latter is caused by a probability distribution for
different molecular shapes which occurs in the mixed phase. Here molecules
can be orientated preferentially along an axis (‘cigars’) or preferentially in
a plane (‘discuses’) with a continuum of intermediate shapes. At high
enough temperatures these dissimilar-shaped quadrupoles can freely rotate,
Fig. 7.6 Comparison of the orientational ordering for (a) a plane section of the
ordered structure and (b) the quadrupolar glass phase; from Suilivan (1983).
‘.l Electric dipolar and quadrupolar glasses 211
Jut as the temperature is lower they seem to gradually freeze out into a
,andom static alignment. Such glassy state is sketched in Fig. 7.6(b).
The main experimental technique used to measure these effects has been
VMR on the protons; resonance-line shapes and nuclear-relaxation times
ue determined as a function of temperature and concentration. Based upon
he results a phase diagram (T-x) is proposed in Fig. 7.7 for the mixed
lo-H,), (p-H&, system. Notice at the very low temperatures (below
1.3 K) a frozen (orientational glass) phase appears to form for a restricted
.ange of X. The coupling between the o-H, molecules is the weak electric
luadrupole-quadrupole interaction. Its conjugate field is an electric-field
Iradient which would be required for the measurement of the quadrupolar
susceptibility. However, it is quite difficult experimentally to produce a
sufficiently large electric-field gradient in order to perform accurate
susceptibility measurements. Consequently, most of the information per-
aining to o-HZ/p-H2 mixed crystals has been obtained from NMR which
should be complemented by other techniques. The generally agreed
nterpretation of the NMR results is a gradual freezing of quadrupole
noments with a distribution of relaxation times which seems to continuously
:volve as the temperature is lowered. The quadrupoles appear frozen when
he time of measurement is smaller than the average relaxation time,
vhereas no freezing is observed for experiments with longer observation
imes. As above we have a spin-glass analogue, but without any indications
)f a sharp and critical phase transition which seems to be generally missing
n these ‘electrical’ systems.
I I I I
- wqx (P-$1,-)(
I
0 0.2 0.4 0.6 0.8 1 .O
ortho-concentration
i‘ig. 7.7 Phase diagram for solid solutions of o-HZ and p-H*. For large concentrations
)f the orientationally anisotropic component (o-H,) the orientationally ordered
chasehas face-centred-cubic structure, the disordered one is hexagonal close packed.
Ihe orientational glass phase (denoted by SG) forms a narrow low-temperature
wedge. From Sullivan et al. (1978).
212 Spin-glass analogues
X= -CJijcOs($i-+j-Aii) (7.2)
{ii}
where
dl (7.3)
A is the vector potential and the integration is along a path connecting the
centres of the ith and jth grains. Jij varies between 0 and J without reversing
sign, but it falls off with distance between the grains as m l/r, for a
Josephson coupling or a exp(-rij) for the proximity effect. A random *
(or ferro/antiferro) interaction is caused by the variable A, as part of the
argument of the cosine in equation (7.2). Let us position the applied
magnetic field along the z-axis, B = BZ and use the gauge for the vector
potential, A = Bxf. Then
where x,Yi and XjYj are the centre coordinates of the ith and jth grains,
7.2 Superconducting glass phases 213
(H>O)
It_
N
(P’PJ
SC
is.
new here is the prediction of a glassy phase just beyond pc for sufficiently
large fields. Glassy means a penetration of the magnetic field into the
superconductor via a completely random alignment of frozen-in supercurrents
among the various grains. This leads to long-range fluctuations in the
magnetic-field profile which decay as a power law with distance. So we
have a highly inhomogeneous field distribution within the superconductor
which yet remains to be detected. Unfortunately, the superconducting order
parameter is experimentally inaccessible. However, the differences between
FC and ZFC of the magnetization and the time dependences (dynamics) of
the glassy state should be measurable characteristics.
Recently there has been renewed interest in the superconducting/glass
concept due to the high-Tc oxide superconductors and their inability to
carry large currents in a field. Firstly, the ceramic form of these materials
is a natural granular superconductor. Thus, the model of the preceding
paragraphs may be directly applied. And, indeed, the experiments just
mentioned have borne out the spin-glass-like behaviour. Secondly, for the
oxide superconductors there are an enormous number of microscopic
defects, which, when combined with the unusually small superconducting
coherence length and large anisotropy, make even bulk singZe crystals a
candidate for glassy behaviour. The name ‘vortex glass’ has been given to
describe the random penetration of magnetic flux in such single-crystal
materials.
It is instructive to reconsider our superconducting order parameter in
order to elaborate our magnetic analogue. Recall that JI is a complex
quantity with two components: amplitude and phase. This corresponds to
the order parameter of a two-components (x-y) magnetic system. Returning
to Fig. 7.8 in the Meissner phase (H < Z&r), + is uniform as in a ferromagnetic
(uniform magnetization). In the Abrikosov phase (H,, < H < Z-Z,,) a
periodic flux-line lattice is formed without any disorder in which +(r) = $(r
+ R) where IR( = a, =(h/2eH). However, if some quenched disorder exists,
as it does in the high-T, materials through the assortment of defects, the
flux-line lattice will be destroyed on long-length scales. Figure 7.9 illustrates
the difference: long-range ordered Abrikosov phase versus short-range
ordered domains of flux lines and finally complete disorder (glass). It is
expected that the latter case is more vulnerable to displacements from the
Lorentz force, I x II, if the superconductor carries a current 1. When such
motions of the previously-rigid flux-line lattice occur, a resistance appears
and the superconductivity will effectively disappear.
As explained before it is the magnetic field via its random vector potential
A + A,, which arises in our hamiltonian that creates the ‘competing
interactions’ and thereby frustrations. Such are the basic ingredients for
forming an analogue spin glass. In a type-II superconductor a vortex glass
(see Fig. 7.9) may, therefore, exist and, since the flux-line lattice is rigidly
frozen, have no resistance at least for small currents. Although our initial
discussions have dwelt upon granular superconductors, now with the advent
7.2 Superconducting glass phases 215
l l l l l l l l l
l l 0 l l l l l l
l l l l l l l l l
l l l l l l l l l
a l eeeeee l
l l l l l l l / aOe
l l l l l l l l l
l l l l l l l l l
l l l l l l l l l
l l l l l l l l l
l l l l ;e l l l l
l l l l /e l l l l
I
l l l l ie l l l l
l l l l /e l l l l
- - ----------I
l l l l
/e l l l l
b l l l ‘/e l l l l
l l l l ’ /e l l l l
~.l l
-.--------4 l l i l l l l l
l l l r--e---c--3-------
e
l l l l
je l l l l
l l l l I
l l l l l
l e l l l e l l
l
l l l
l l l
l e l
l l l l eee l l
l *e l e l l l l
l l l e l
l
l e l
l l l
’ l l l l
l l
l l l l l
l e l
l l l
l
l l l l l l l
l
l e
l
l l
l l e
l eeeee l l m
Fig. 7.9 Superconducting flux-line lattice for H,, < H < &: (u) long-range
ordered, Abrikosov triangular lattice; (b) short-range ordered lattice with domains;
and (c) vortex glass. For the magnetic field into the page the supercurrents circulate
clockwise around each dot.
216 Spin-glass analogues
of the high-T, oxide materials, we can apply the model to a distinct phase
of bulk, single-crystal superconductors. We continue by drawing a direct
analogue with our spin glasses. The hamiltonian, if the A, values are
constrained to the range between 0 and 7~, is similar to the EA x-y +-J
spin glass. Based upon our previous review of the theory, especially the
computer simulations, this universality class is not expected to have a phase
transition in d = 3. For a superconductor a more realistic model would be
to let the A, vary from 0 to 27~. Such a model is called a ‘gauge glass’.
This means that the frustration is a serious disorder which is invariant under
the gauge transformations Si + -Si and J, -+ -Jij. The disorder cannot be
‘gauged away’ using the word gauge in analogy with the invariant
transformations of electro-dynamics. A gauge glass does not have reflection
& + -& as one of its symmetries because the magnetic field breaks the
time reversal invariance. And this lack of symmetry distinguishes it from
the x-y (0 to n) model. So present theoretical thinking is that the gauge
glass belongs to a different universality class and may very well have a
phase transition in 3D. Extensive Monte-Carlo computations indicate that
the gauge-glass model in 3D has behaviour comparable to an Ising spin
glass (d = 3) which should have a finite T, rather than to an x-y spin glass
which should not.
Experiment has tried to determine the various properties of the flux-line
lattice, especially to establish the existence of an intermediate phase with
zero resistance corresponding to the vortex glass (see Fig. 7.9). Initial
indications seem to be favourable for its formation accompanied by a sort
of phase transition and there is experimental agreement with other effects
predicted by the model. Since a flux-line lattice is a mesoscopic phenomenon
with larger than atomic dimensions (typical lattice constants are of order
microns), it can be directly observed via scanning tunnelling microscopy or
decoration techniques. To find such a glassy structure as sketched in Fig.
7.9 would be a challenging experiment and complete confirmation of the
model, as would the observation of the melting behaviour at higher fields
and temperature.
Before continuing with our spin-glass analogues and our transversal into
areas far different from our spin systems, let us pause to examine another
type of random magnet which is not a spin glass. The model, first proposed
by theorists (Imry and Ma, 1975), is called the random-field Ising model
(RFIM). Here an Ising ferromagnet of various dimensions is placed in a
site-random magnetic field. The model hamiltonian becomes
where the spins are all z-components, Jij is a constant (J), representing
positive ferromagnetic interactions between nearest-neighbour spins, and
the Hi are independent, local, random variables with
[HilAV = 0 (mean) [HflAv = Hg (variance) (7.7)
Using the following simple argument Imry and Ma concluded that such
RFIM systems would exhibit a phase transition for dimension d > 2 (d, = 2,
the lower critical dimension). The net-energy cost for the formation of an
oppositely-oriented domain of volume Rd out of the ferromagnetic state is
E(R) = JRd-’ - HRRdf2 (7.8)
The first term represents the domain-wall energy where the spin bonds are
broken along a surface R d - l . The ferromagnet in the presence of random
fields has a mean-square energy of HRR 2 d. For a particular choice of an
oppositely-oriented domain, this energy (after taking the square root) can
lower the total energy of the domain formation by a more favourable
alignment with the random fields. And thus we arrive at the second term
which is of opposite sign to the first. For HR * J the total energy is positive
for d > 2 (domains cannot form), but for d < 2 it becomes negative for
sufficiently large R. This means that domains are energetically favourable
and will be created thereby destroying the long-range order phase and its
phase transition. The d = 2 case is unresolved, hence giving rise to the
lower critical dimension.
Therefore, in 3D a special phase transition and critical phenomenon are
expected for the RFIM but not in 2D. And the H-T mean-field phase
diagram for this transition should be as sketched in Fig. 7.10. However, in
order to verify such behaviour we need a real system to measure. How do
ZFC
--*_
* +..,
.
T
TN TN
Fig. 7.10 Zero-field-cooling (ZFC) and field-cooling (FC) procedures for the RFIM.
The dotted line indicate the AF phase boundary in the H versus T phase diagram.
The arrows show the direction of the field-temperature cycling in each procedure.
P represents paramagnet and AF the RFIM-state which develops at sufficient field.
TN is the REIM Neel temperature (H = 0).
218 Spin-glass analogues
Paramagnetic
i 1 FC ! / RFIM ,
Random exchange
‘,\ /’ REIM
TN
Temperature
Fig. 7.11 Schematic H-T phase diagram for the diluted AF, d = 3, RFIM system
at a fixed concentration, following a ZFC procedure. New RFIM critical behaviour
is observed for T,; < T < TA. All boundaries, both static and dynamic, obey REIM
to RFIM crossover scaling HU+, with Q, = 1.42. T*(H) indicates the onset of the
dynamic rounding of the critical behaviour associated with the extreme slowing
down as T + T,(H). It is only weakly time (frequency) dependent, as indicated by
the characteristic times associated with particular measurements. Outside of the
shaded region, the system is in equilibrium on all accessible time scales. In an FC
route (inset), the system falls off equilibrium at T,,(H), preventing the correlation
length 5 from diverging further for T < T,,(H) and ultimately leading to a frozen-
in (non-equilibrium) domain state below T,(H). From Jaccarino and King (1990).
line labelled T,,(H) in the ellipsoidal inset of Fig. 7.11. Now the system
falls out of equilibrium for T < T,,(H), thereby preventing the correlation
length 5 from further diverging and ultimately leading to a non-equilibrium
frozen-in domain state which forms below T,(H).
In 2D RFIM specific-heat and neutron-scattering measurements of
RbzCoXMg,-xF4 display a systematic rounding of the critical behaviour,
independent of the cooling procedure. There is no hysteresis in the vicinity
of the ‘destroyed’ phase transition at ‘T,(H)‘. From these data it was
concluded that d, 2 2. By utilizing the field scaling of the non-equilibrium
properties well below ‘T,(H)‘, the crossover exponent + is determined to
be 1.74, indistinguishable from the pure Ising value of 1.75. Figure 7.12
gives the schematic H-T phase diagram for the diluted antiferromagnetic
d = 2 case. Since hysteresis only occurs below ‘T,(H)‘, the FC and ZFC
procedures yield identical results in the surrounding ‘T,(H)’ region which
represents an equilibrium domain state, i.e., no long-range order or 5
remains finite. When hysteresis finally begins (see ellipsoidal inset of Fig.
7.12), it is to a non-equilibrium, metastable, frozen state that depends on
220 Spin-glass analogues
d=2
Domain state Paramagnetic
‘N
TEMPERATURE
Fig. 7.12 Schematic H-T phase diagram for the diluted AF, d = 2 = dl RFIM
system at a fixed concentration. The destroyed transition ‘Z’,(H)’ broadens and
shifts with HZ+ crossover scaling; it separates a domain state below ‘T,(H)’ from a
paramagnetic one above. Here + = 1.75. A freezing (metastability) boundary T,(H)
exists well below ‘T,(H)’ and also scales as Hz’+; its position depends on the
measurement time scale 7. Notice that destruction of the phase transition is an
equilibrium event. Unlike the d = 3 case, non-equilibrium behaviour at d = 2 has
nothing to do with extreme critical slowing down. From Jaccarino and King (1990).
the time scale of the measurement as indicated in the figure. Here 7’r(~) is
the time-dependent freezing temperature which tracks this loss of equilibrium.
Note the distinction between this freezing (non-equilibrium) behaviour and
the extreme critical slowing down in equilibrium of the rounded or destroyed
phase transition at ‘T,(H)‘.
Returning to the 3D RFIM systems, we briefly examine the dynamics of
the phase transition. Theory (Villian, 1985; Fisher, 1986) predicts an
activated-dynamics model where now the relaxation time becomes
CO
1
7
- = exp (ke) = exp 1 (7.9)
70 L[ T- T,(H)]“e
here we have used the standard result for the RFIM correlation length
5= 50 (7.10)
should be compared to our previous usage in the 2D, Ising spin glass
(Tc = 0) of section 6.2. According to the RFIM theory, the free energy
and the associated thermodynamic quantities vary logarithmically with time
or frequency rather than linearly. For example, the frequency-dependent
Free energy scales as
(7.11)
x(w, T) = cy-” x
ln(o/o,)
____
Se [ 1
Hence, there exists a dynamical rounding temperature which limits the
(7.12)
TN
)
I I-
> t*(w) cf ln- w ev
WO
since dynamical effects (rounding) should start appearing for ln(w/w,)/[e = 1.
(7.13)
Fig. 7.13 x:(w) versus (t(, the reduced temperature, at three frequencies after
background subtraction, see inset. The open and closed symbols refer to T < T,(H)
and T < T,(H), respectively. Rounding of the transition due to the concentration
radient occurs only within the shaded region, i.e. ItI < 2 x 10d4. The expected In
f t( critical behaviour of X;(W) is used to determine a dynamic rounding temperature
t*(w) as indicated. From King et al. (1986).
two parameters with 8 = 1.0 and o, = lo7 rad/s. Since Y has also been
found from neutron scattering to be equal to one, the product 8v = 1.0
and hence the relaxation form is a simple Vogel-Fulcher type:
7 = r,exp (7.16)
between two chips placed one above the other. The number of signals that
pass between the ith and jth circuits is denoted by Kiie Next we attach a
two-value variable Si to each circuit i, such that Si = +l for the upper chip
and -1 for the lower one. The number of signals crossing from one chip
to the other is
C (A-Kq/2)s,sj (7.18)
(ii)
should be minimized. Here A (a constant) takes into account the imbalance
in the number of circuits on the two chips and in the number of boundary
crossings. The above quantity to be minimized resembles an Ising, spin-
glass hamiltonian where the ‘interactions’ are the short-range ferromagnetic-
like Kij and the long-range antiferromagnetic-like A. Due to the competing
interactions and the intrinsic randomness, frustration results, and thus, we
have all the basic ingredients of a spin glass.
In order to secure the analogy of combinatorial optimization with the
spin glasses let us re-examine the SK-model in the former language. An
instance of this problem is a sample consisting of a given number of spins
N and a set of couplings Jij where 1 < i < j < N. The domain of possible
instances is defined by a probability distribution which reproduces the
choice of couplings. In the SK-model each Jij is taken as an independent
random variable with a gaussian distribution: exp [-(N/2).3. A configuration
is characterized by the values of the N Ising spins, Si = +l. And finally
the cost function is the energy given by the hamiltonian
%e= - ~ JijSiSj (7.19)
Iti<j<N
E(r) = E, - C, sr (7.21)
( 1
The conclusion is that for the latter four models the slow logarithmic
226 Spin-glass analogues
+pk (7.22)
(ii) J-N
This is the quantity to be optimized by finding the shortest length lmin. Als
are equivalent to AEs and the ‘temperature’ equivalent generates the
acceptance criterion for a given configuration. Initially, at high temperature,
the algorithm accepts deteriorations in the cost function with a high
probability. However, in the course of the algorithm’s running time the
probability is gradually decreased to zero. This is accomplished by a control
parameter or temperature c which is a function of t, the number of MCS,
whereby c + 0 at the end of the computer run. Notice that we are still far
away from lmin with the finite rs used in the algorithm. So we must
extrapolate to r (or l/in r) + 0 as done in Fig. 7.14. The extrapolated
results of fmin agree nicely with the ‘exact’ value for N = 100 and the
‘expected’ value for N = 400.
1.0
0.61 ’ 0.2
0 0.1 ’ 0.3
’ 0.4
’ 0.5
’ 0.6
’ 0.7
’ 0.6
’ 0.9’ ’
1.0
’
1.1 . 2
-(l/lnr) l 10
Fig. 7.14 Optimal tour length 1versus - (ln r)-l for the N-city ‘travelling salesman’
problem for N = 100 and 400. From Grest et al. (1987).
7.4 Other ‘spin-offs’ 227
Here the Jij need not be symmetric, yet both + and - (competing ferro-
and antiferromagnetic) pseudo-exchange interactions do exist since the
neuron synopses can be ‘excitatory’ or ‘inhibitory’. Also the various couplings
Jij may change with time, i.e., the system can learn in addition to store.
The set of all the spin values represents the collective state of the system
and such a situation closely resembles our SK, Ising hamiltonian in a
random external field. The solution of this hamiltonian, based upon our
previous spin-glass experience is expected to be represented by the multi-
valley, phase-space landscape of the many nearly-degenerate ground states.
In the present case these local energy minima of the collective system
correspond to patterns to be recognized or memories to be recalled or
other types of mental behaviour. It is the set of neural connections that
fixes the number location and meaning of the valleys. A given external
stimulus (input) determines the initial state of which neurons are firing and
which are not. Afterwards the system via the particular set of couplings
evolves to a special low valley (memorized pattern) that represents the class
to which the input stimulus belongs. Thus, the memorized patterns are
retrievable by their association with a given input pattern or addressable
by the content of the input pattern. A ‘basin of attraction’ surrounds the
lowest point in a valley such that if the system finds itself anywhere near
the valley it will be attracted to the lowest point - a non-ergodic process.
Again stimuli from the external world generate the selection of a special
solution or ground state. Note that with N-neurons there are 2N different
patterns for the network.
The choice of the Jij usually takes the form
where the 67 are quenched random variables assuming the values +l and
-1 with prescribed probabilities. p corresponds to the p-learned patterns
of N-bit words which are fixed by the p sets of { @} . From a straightforward
stability analysis, the stored patterns would only be stable at T = 0 K for
negligibly small values of p/N. However, we must recompute the stability
criterion using an upper limit of retrieval error of learned patterns. This
greatly increases the maximum number of patterns able to be stored in an
N-neuron network. As a first step towards the calculation of retrieval error,
we define a so-called Hamming distance d between a pure learned pattern
Sk and the pattern S’ as
The Hamming distance represents the overlap between the two patterns
7.4 Other ‘spin-offs’ 229
It is exactly the same concept resulting from RSB and ultrametricity. The
average number of errors is given by the number of spins which do not
align with the learned or embedded pattern, and mathematically this
becomes
N, = :(1-q) (7.33)
Recent calculations have shown that for NJN = 1%) p/N = 0.14 and for
smaller error limits, the ratio p/N goes rapidly to zero. In fact, if reasonable
retrieval errors are permitted, the maximum storage capacity of learned
patterns may be increased to pmax = O.l4N, a rather high percent.
Relaxation of the neutron or spin system from the initial stimulus or state
to the final steady-state pattern or valley state is governed by the dynamics.
Any model of associated memory requires a well-defined dynamical process.
The Hopfield model assumes a single spin-flip Glauber dynamics which was
discussed in section 5.12 on Monte-Carlo spin-glass simulations. Here, after
every updating of a spin, it is the new configuration which is used to update
the next one. This is nothing but the heat-bath version of the Monte-Carlo
process required for the Ising (up/down) spin glasses. The other model,
called Little dynamics, updates all N-spins independently of each other.
Now at each time step all the spins simultaneously check their states against
the corresponding local fields. Hence, such an evolution of parallel
relaxations is called synchronous whereas the Hopfield ones are asynchronous
or series relaxations. Different forms are derived for the firing probability
of the neuron (or spin) at site i and t+Aht given the configuration {Si} at
time t. But it must be remembered that the dynamics are part of the model
and not a fundamental stochastic process. Various neuron-like electronic
circuits with the above properties have been built to study collective and
complicated computations and this is just the beginning of an exciting new
field.
We conclude our brief treatment of neural networks with a simple
illustration of associative memory whereby pieces of the stored information
serve to retrieve the entire memory. Consider the six-neuron network of
Fig. 7.15 and let the neurons have the + 1 or - 1 features corresponding to
the description of a person according to name, height, age, weight, hair
and eyes. See Table (a) as inset to Fig. 7.15. We wish to store three sets
of characteristics for three different people as memories A, B and C. See
Table (b) as inset to Fig. 7.15. Table (6) determines the wiring diagram of
the 6-neuron network. For example, the link between neurons (1) and (2)
must be for A + (+l)(+l), for B 3 (+1)(-l), and for C + (+l) (+l).
230 Spin-glass analogues
1 2 3 4 5 6
1 2 3 4 5 6
A +1 +1 +1 -1 -1 -1
I
B +1 -1 +1 +1 -1 +1
C +1 +1 -1 +1 -1 -1
Fig. 7.15 Associative memory with six nodes, or ‘neurons’, is linked by excitatorv
(solid line) and inhibitory (broken line) connections. The number of lines in eat
link indicates the strength of the connection; each solid line represents a connectio
strength of +l and each broken line represents a strength of -1. Each node migl
denote a characteristic of a person, as is shown in the Table (a). Suppose one wan
to store three memories, or sets of characteristics as given in Table (b). The node
that are supposed to be in the +l state are given an excitatory link to the othc
+1 nodes and an inhibitory link to the -1 nodes. To store information about a
three memories one simply adds up the connections. From Tank and Hopfiel
(1987).
7.4 Other ‘spin-offs’ 231
References
If you have come this far you must either be an expert on the spin glasses
or slightly disappointed and still confused. In order to improve and solidify
the first and to alleviate the second possibilities we offer this concluding
chapter. It will not be a succinct recapitulation of the preceding seven
chapters. Hopefully after a rereading they will become sufficiently clear
and to the point. The phenomenon and physics of ‘spin glass’ have been
explained via an experimental approach - the purpose of this monograph.
In this final chapter we wish to reiterate briefly, using other words, the
basic and key concept of the spin glasses, namely, symmetry breaking
resulting in the rough multi-valley landscape of the system’s energy in phase
space. Then, we proceed to one of the latest (1992) developments in the
experimental area that of mesoscopic or nanostructured spin glasses and
their special fluctuations. And afterwards we take a last quick look at the
theory: what does it say about dimensional crossover (3D + 2D) and how
does the freezing temperature depend on the length scale? This is followed
by some of the most recent computer simulations. Here there are appearing
some novel and altered interpretations. These final considerations will
clearly illustrate to us the ongoing stream of progress in a relatively dormant
field of physics. At last we end the book with a few compositional thoughts
about our journey through the spin glasses and where to find what.
UJ
T<Tc
1
-tvl- +M
Fig. 8.1 Free energy as a function of the magnetization (order parameter) for a
ferromagnetic above and below its Curie temperature Tc.
(a>
(b)
-Mj/M
A = qEA - 4 (8.2)
Notice that A + 0 when we return to our single equilibrium phase of Fig.
8.1 (T < T, with a small +H), even though one half of the phase space
has been removed.
As stated previously the key to the spin glass and its many analogues is
this peculiar and undeterminable multi-layer landscape - a unique but
necessary conception. This is an important bit of new physics with significant
consequences that can be tracked experimentally. For example, if we
measure the ac-susceptibility we are effectively probing (IuA and its short-
time quasi-equilibrium state. However, if we try to ‘mimic’ the very long-
time behaviour by field cooling, we obtain a measure of q which gives the
average equilibrium susceptibility. Can you remember this discussion in
section 5.8? More important is not to forget the picture and underlying
concepts of the many valleys, so you should look once again at Figs. 8.1
and 8.2 for here is the heart of the spin-glass problem.
Very recently (1990s) the study of spin glasses has turned to samples with
reduced size. This means one or more dimension has been constrained to
less than c. 1000 8, lengths. The structure may be a very thin film (2D), a
narrow ‘wire’ or strip (1D) with both reduced thickness and width, or a
‘clump’ of the material usually called a dot (OD). (The latter has not yet
been tried for a spin glass.) Such samples are fabricated using the modern
sub-micron and nanostructure techniques currently available. The word
applied to describe these samples and their associated physics is mesoscopy.
238 The end (for now)
At the present time it is a very popular and active area especially with
semiconductors where its consequences and benefits are obvious.
The unique effects of mesoscopic physics (which lies between the very
small atomic scale, = 1 A, microscopy and the very large bulk behaviours,
2 few l.r,rn, macroscopy) are caused by the smallest sample length L being
comparable to one or more of critical length scales of the given problem.
Thus, for a mesoscopic conductor, L could be the same order of magnitude
sas
(i) the elastic mean free path e = vg, where vF is the Fermi velocity
and T is an elastic scattering time;
(ii) the inelastic mean free path [in which depends on the inverse
temperature to some power and includes now magnetic inelastic
scattering processes, e.g. spin flip;
(iii) the phase-coherence length Q, which permits ‘time-reversal’ interference
between the electron states.
For the latter we can write in the diffusive-scattering regime &, = (Dr+);
where D is the diffusion constant and T+ is related to the inelastic scattering
processes. A typical metal has 4 = 50 nm and 4, = 1000 nm at 1 K, hence,
L could easily be made much less than e, by sub-micron fabrication and
going to yet lower temperatures. Such a condition (L < Q,) eliminates a
complete cancellation of the different quantum interference terms (remember
the electron is a wave) between the various diffusive (4 4 L) paths. This
in turn produces finite fluctuations in the conductivity 6G, which depend
sensitively (in both magnitude and sign) on the specific placement of the
scattering centres or ‘impurities’. These direct interferences between multiply-
scattered electron waves give the so-called universal conductance fluctuations
(UCF) of amplitude
(8.4)
Here the 4, e, and L have been previously defined, and kF and TF are the
Fermi momentum and temperature. Note that the freezing temperature Tf
enters this equation. If k&T,IT, 2 P/e& then we have the saturation
regime
1o-6
0 20 40 60 80
T(K)
Fig. 8.3 Electrical resistance ‘noise-parameter’, a(f, T), for various concentrations
of CuMn; from Israeloff et al. (1989).
-1
“0 -2
C
x
-3
5
a -4
-5 -2.5
-6 -3.0
0 15 45 60
;i” (kG)
Fig. 8.4 Upper three curves magneto-resistance as a function of field for a CuMn
(0.1 at. %) film (420 8, thick by 2 Frn wide). Lower two curves same as above but
for a CuMn (0.1 at. %) wire (420 A thick by 900 A wide by 2 km long). From de
Vegvar, Levy and Fulton (1991).
However, once the wire is heated above Tr and retooled a new fluctuation
spectrum will appear in the magneto-conductance related to the new
microscopic orientations of frozen spin. It is estimated that within the above
wire dimensions there are lo5 Mn spin and 25% of these must reorient to
produce of 6G = e2h. The question here is what does reorientation mean,
certainly not a spin flip, perhaps a small collective distortion would be
sufficient. But this is just the beginning of the new and exciting field of
mesoscopic spin glasses.
So what has been happening with the theory? In this subsection we mention
one ‘older’ (late 1980s) and one brand-new (1992) pieces of progress. The
former has to do with the droplet model and its approach towards answering
the question of dimensional crossover: 3D + 2D and this influence on T,.
Certainly it is related to the previous topic of mesoscopic or restricted-
dimension effects. The second concerns some very recent numerical
simulations carried out in improved ways or on more realistic models which
indicate that specific revisions are perhaps necessary in the long-accepted
theoretical conclusions.
242 The end (for now)
0.20
0.00
0.0 0.5 1.0 1.5 2.0 2.5
AH (kG)
Fig. 8.5 Aharonov-Bohm conductance oscillations (AG versus AH) for a CuMn
$.l;t. %) square ring 0.4 pm sides at 23 mK; from de Vegvar, Levy and Fulton
The first question is how does Tf change as we reduce one of the sample
dimensions beginning with a bulk 3D system. For example, we can decrease
the thickness of the sample until the 2D limit is reached, all the while
tracking Tf. The simplest theoretical approach is to use finite-size scaling
which introduces the new variable y = L/c(t). Here L is the sample
dimension that we are changing and t[t = (T- T,)/T,] is the correlation
length for an equilibrium phase transition. For the spin-glass case we have
a reduction in T,(L) at a given constant value of y. Thus, setting
T,(L = 03) = T, we obtain
T,(L) - T,(L=~) cc -$
T,(L=m) (8.6)
with a being a combination of the other four (3r3, f13, I& and v2) critical
exponents.
The above equations suggest the ways in which the experimentalists
should plot their data. The different critical exponents may be estimated
from MC simulations or the measurements on ideal spin glasses - see
Chapter 6. And as mentioned in section 6.4 a number of groups are quite
active in preparing thinner and thinner films, usually of the canonical
(RKKY) spin glasses, in order to confront the theory. Figure 8.6 illustrates
one such ‘multilayer’ attempt. Here various Mn alloys of Cu or Ag, which
are kept separated from each other by a sufficiently thick interlayer of the
pure host metal (Cu or Ag), have their thicknesses reduced. T,(L) is taken
0.8- r’
.a
,600
0.67 12: ’ cuo.a~oo.,,/cu
’ cuo.oo~no.,,/Cu
0.4- ,9J A Cu0.0dno.dA8
0 Cu0,0dn0.0&u
* cuo.“~o..o,/cu
0 A80.01~n0.m./Cu
’ A8o.oo~oo.oo/cu
A A8o.dndA8
I .,.14
102 1
spins interacts with all the others (N-l) spins according to a constant
coupling strength J independent of the distance separating the spins. Only
the sign of J changes randomly. Here dimensionality does not play a role
and it is the size of the system given by the N spins that counts. Large-
scale (to N = 4096), long-time Monte-Carlo procedures were used and
averaged over various sets of bond configurations, i.e., sample averaging.
When the number of bond configurations averaged over was large enough,
the results confirmed Tf = J (Bhatt and Young) as predicted by the SK
continuous infinite-range model and reinforced by RSB theory. However,
for insufficient number of samples, the generated (non-linear) susceptibility
values were widely distributed and highly skewed. Thus the conclusion of
Campbell that Tf = 0.75 was biased because of the large fluctuations or
“rare samples” present in his simulations. Such effects can only be gotten
rid of by proper sample averaging in these finite size samples. So you see
there is still life, danger and contention in the spin glasses and certainly
other paragraphs of this book will require future revision as they continue
to evolve.
The ‘more-realistic’ model (Matsubara and Iguchi, 1992) is a random
alloy of RKKY-coupled, Heisenberg spins and the numerical simulation is
a hybrid MC spin-dynamic method. Here the model hamiltonian
X = -C J(r) SiSj
ij
is used with J(r) being given by equation (1.2). The classical Heisenberg
spins ISi\ = 1, are randomly distributed on the sites of an (L)3 fee lattice.
Parameters are chosen to represent CuMn with, for example, 5 at. % Mn.
The spin-glass susceptibility is calculated as
(8.10)
+ L= 6
* L= fJ
...& L=1’)
80k o Lz1.2
ae o L=16
t. ;..:.\ : 0 L=20
E....
Il...g
-. +...+ :::*::: *::::,8:zax,...,..
Before we stop (for now) let us briefly look back at the composition of
this book. If you want to learn, put most simply, what a spin glass is and
8.4 A final word 247
meaning are given via a few highly helpful sketches. Hopefully the above
acronyms are still familiar to the reader. The all-important dynamics enters
our discussion as we wind up with the mean-field model. It is then further
onwards into the more recent theoretical issues of the droplet and fractal-
cluster models. Finally, we switch from Ising (as with the preceding topics)
to non-Ising vector spin glasses and conclude Chapter 5 with the computer
simulations to which we returned briefly in Chapter 8.
In order to confront the experiment of Chapter 3 with the theory of
Chapter 5, we require ideal spin glasses. In Chapter 6 we first explain what
is an ideal spin glass, and then consider a 2D example and a 3D less ideal
one for detailed comparisons. A demanding (perhaps overly so) examination
is carried out to evaluate the correspondence of experimental results to
theoretical predictions. The emphasis here lies with the nature of the phase
transition and its critical properties. Chapter 6 closes with the conclusions
and some future possibilities.
At long last in Chapter 7 we turn our attention to non-spin analogues.
This area has become a vast out-growth of the original magnetic alloys.
Yet the ideas and notions are very similar only without the word ‘spin’. A
few examples are electric dipolar and quadrupolar glasses, and supercon-
ducting glass phases. For the sake of comparison, we then briefly revert
back to a spin system, the random-field Ising model, whose phase transition
is well understood. Finally and succinctly, we introduce three non-
physical analogues, namely, combinatorial optimization, neural networks
and biological evolution. Books can and are being written about these
topical subjects. However, we wish to show the reader that they owe a
great deal to the spin glasses for their recent development. For they all
stem from the basic spin-glass hamiltonian and possess the same jagged
phase space.
In this final chapter, which for obvious reasons is called ‘The end (for
now)‘, we have attempted to place the spin glasses within the notions of
symmetry breaking and phase space. The unique feature here is the multi-
valley landscape of the spin-glass phase space and it is the key to
understanding much of the unusual behaviour. In order to prove that there
is still life and new physics in the spin glasses we discuss the nascent subject
of mesoscopic spin glasses to which significant experimental interest is
turning. Also from the theory side we related how this reduced geometry
(or nanostructure) should influence the freezing temperature. Since the
theoretical efforts are continuing to generate progress, we have selected
some very recent and undigested Monte-Carlo numerical simulations and
tried to outline their provocative conclusions. And last but not least, here
we are attempting to summarize the book’s composition by giving the
reader some tips on where what is and how to find it in a ‘final word’.
REFERENCES 249
References