You are on page 1of 824

Handbook of Magnetic Materials,

Volume 3
North-Holland Publishing Company, 1982

Edited by: E.P Wohlfarth


ISBN: 978-0-444-86378-2

by kmno4
PREFACE

This H a n d b o o k on the Properties of Magnetically Ordered Substances, Ferro-


magnetic Materials, is intended as a comprehensive work of reference and
textbook at the same time. As such it aims to encompass the achievements both of
earlier compilations of tables and of earlier monographs. In fact, one aim of those
who have helped to prepare this work has been to produce a worthy successor to
Bozorth's classical and monumental book on Ferromagnetism, published some 30
years ago. This older book contained a mass of information, some of which is still
valuable and which has been used very widely as a work of reference. It also
contained in its text a remarkably broad coverage of the scientific and tech-
nological background.
One man can no longer prepare a work of this nature and the only possibility
was to produce several edited volumes containing review articles. The authors of
these articles were intended to be those who are still active in research and
development and sufficiently devoted to their calling and to their fellow scientists
and technologists to be prepared to engage in the heavy tasks facing them. The
reader and user of the H a n d b o o k will have to judge as to the success of the choice
made.
Each author had before him the task of producing a description of material
properties in graphical and tabular form in a broad background of discussion of
the physics, chemistry, metallurgy, structure and, to a lesser extent, engineering
aspects of these properties. In this way, it was hoped to produce the required
combined comprehensive work of reference and textbook. The success of the
work will be judged perhaps more on the former than on the latter aspect.
Ferromagnetic materials are used in remarkably many technological fields, but
those engaged on research and development in this fascinating subject often feel
themselves as if in strife for superiority against an opposition based on other
physical phenomena such as semiconductivity. Let the present H a n d b o o k be a
suitable and effective weapon in this strife!
The publication of Volumes 1 and 2 took place in 1980 and produced entirely
satisfactory results. Many of the articles have already been widely quoted in the
scientific literature as giving authoritative accounts of the modern status of the
vi PREFACE

subject. One book reviewer paid us the compliment of calling the work a
champion although with the proviso that the remaining two volumes be published
within a reasonable time. The present Volume 3 goes halfway towards this event
and contains articles on a variety of subjects. There is a certain degree o f
coherence in the topics treated here but this i s not ideal due to the somewhat
random arrival of articles. The same will be the case for the remaining Volume 4
as such, although this will then complete the work so as to finally produce a fully
coherent account of all aspects of this subject.
Three of the authors of Volume 3 are members of the Philips Research
Laboratories, Eindhoven and, as already noted in the Preface to Volumes 1 and 2,
this organization has been of immense help in making this enterprise possible.
The North-Holland Publishing Company has continued to bring its profes-
sionalism to bear on this project and Dr. W. Montgomery, in particular, has been
of the greatest help with Volume 3. Finally, I would like to thank all the authors
of Volume 3 for their co-operation, with the profoundest hope that those of
Volume 4 will shortly do likewise!

E.P. Wohlfarth
Imperial College
TABLE OF CONTENTS

Preface . . . . . . . . . . . . . . . . . . . . . . v

T a b l e of C o n t e n t s . . . . . . . . . . . . . . . . . . vii

List of C o n t r i b u t o r s . . . . . . . . . . . . . . . . . . ix

1. M a g n e t i s m a n d M a g n e t i c M a t e r i a l s : H i s t o r i c a l D e v e l o p m e n t s a n d
P r e s e n t R o l e in I n d u s t r y a n d T e c h n o l o g y
U. E N Z . . . . . . . . . . . . . . . . . . . . . 1
2. P e r m a n e n t M a g n e t s ; T h e o r y
H. Z I J L S T R A . . . . . . . . . . . . . . . . . . 37
3. T h e S t r u c t u r e a n d P r o p e r t i e s of A l n i c o P e r m a n e n t M a g n e t A l l o y s
R.A. McCURRIE . . . . . . . . . . . . . . . . . . 107
4. O x i d e S p i n e l s
S. K R U P I C K A a n d P. N O V A K . . . . . . . . . . . . 189
5. F u n d a m e n t a l P r o p e r t i e s of H e x a g o n a l F e r r i t e s with M a g n e t o p l u m b i t e
Structure
H. K O J I M A . . . . . . . . . . . . . . . . . . . 305
6. P r o p e r t i e s of F e r r o x p l a n a - T y p e H e x a g o n a l F e r r i t e s
M. S U G I M O T O . . . . . . . . . . . . . . . . . . 393
7. H a r d F e r r i t e s a n d P l a s t o f e r r i t e s
H. S T J i d ~ L E I N . . . . . . . . . . . . . . . . . . 441
8. S u l p h o s p i n e l s
R.P. V A N S T A P E L E . . . . . . . . . . . . . . . . . 603
9. T r a n s p o r t P r o p e r t i e s of F e r r o m a g n e t s
I.A. CAMPBELL and A. FERT . . . . . . . . . . . . 747

Author Index . . . . . . . . . . . . . . . . . . . . 805

Subject Index . . . . . . . . . . . . . . . . . . . . 833

Materials Index . . . . . . . . . . . . . . . . . . . 845

vii
chapter 1

MAGNETISM AND MAGNETIC


MATERIALS" HISTORICAL
DEVELOPMENTS AND PRESENT
ROLE IN INDUSTRY AND
TECHNOLOGY

U. ENZ
Philips Research Laboratories
Eindhoven
The Netherlands

Ferromagnetic Materials, Vol. 3


Edited by E.P. Wohlfarth
© North-Holland Publishing Company, 1982
CONTENTS

Introduction 3
1. F r o m l o d e s t o n e to f e r r i t e : a s u r v e y of t h e h i s t o r y of m a g n e t i s m . . . . . . . . . 3
2. T h e r o l e of m a g n e t i s m in p r e s e n t - d a y t e c h n o l o g y a n d i n d u s t r y . . . . . . . . . . 6
3. D e v e l o p m e n t o f s o m e classes of m a g n e t i c m a t e r i a l s . . . . . . . . . . . . . . 10
3.1. I r o n - s i l i c o n a l l o y s . . . . . . . . . . . . . . . . . . . . . . . . 10
3.2. F e r r i t e s . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.3. G a r n e t s . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.4. P e r m a n e n t m a g n e t s . . . . . . . . . . . . . . . . . . . . . . . 24
4. T r e n d s in m a g n e t i s m r e s e a r c h a n d t e c h n o l o g y . . . . . . . . . . . . . . . . 30
4.1. M a g n e t i s m r e s e a r c h b e t w e e n p h y s i c s , c h e m i s t r y a n d e l e c t r o n i c s . . . . . . . . 30
4.2. T r e n d s in a p p l i e d m a g n e t i s m . . . . . . . . . . . . . . . . . . . . 31
4.3. O u t l o o k a n d a c k n o w l e d g e m e n t . . . . . . . . . . . . . . . . . : . 34
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
Introduction

In this contribution we attempt to trace a few main developments of the history of


magnetism and to give an account of the present role of ferromagnetic materials
in industry and technology. The treatment of a subject as broad as the present one
must necessarily be limited and incomplete; nevertheless, we may give an im-
pression how the large body of knowledge on magnetism accumulated in the past,
and how important it is at present. The first section gives a short sketch of some
early historical developments and inventions. Such a flash back to history may be
useful to place the m o d e r n activities and achievements in a wider context. The
next section deals with the role of magnetism and magnetic materials in modern
technology, especially in the context of power generation and distribution, tele-
communication and data storage. Some statistical figures on the economic
significance of magnetic materials are included. The third-section gives a some-
what more detailed account of the development lines~i~f a few selected classes of
materials, whereas in the last section an attempt is made to indicate the trends in
applied magnetism.

1. F r o m lodestone to ferrite: a survey of the history of m a g n e t i s m

The notion of magnetism dates back to the Ancient World, where magnets were
known in the form of lodestone, consisting of the ore magnetite. The name of the
ore, and hence that of the whole science of magnetism, is said to be derived from
the G r e e k province of Magnesia in Thessaly, where magnetite was found as a
natural mineral. It seems very likely that the early observers were fascinated by
the attractive and repulsive force between lodestones. Thales of Miletus (624-547
BC) reports that the interaction at a distance between magnets was known before
800 BC. Another, probably m o r e apocryphal account is due to Pliny the Elder,
who ascribes the n a m e magnet to its discoverer, the shepherd Magnes "the nails
of whose shoes and the tip of whose staff stuck fast in a magnetic field while he
pastored his flocks". From such modest beginnings grew the science of mag-
netism, which may be represented as a tree on whose growing trunk new shoots
and branches continuously appeared (see fig. 1). The trunk represents the mag-
4 u. ENZ

netic materials such as metals, alloys or oxides, because history shows that the use
and study of materials have been the main sources of discoveries and progress. T h e
new branches which developed in the course of time f o r m e d scientific fields in
themselves. A brief account of s o m e of these d e v e l o p m e n t s is given in the
following pages (Encyclopedia Britannica: Magnetism; see also, Mattis (1965)).
Magnets f o u n d their first application in compasses, which were m a d e from a
lodestone b u o y a n t on a disc of cork. W e k n o w that the compass was used by
Vikings and, of course, by Columbus, but the art of navigation guided by
compasses m a y be much older. T h e invention is p r o b a b l y of Italian or A r a b i c
origin. T h e earliest extant E u r o p e a n reference to the compass is attributed to the
English scholar A l e x a n d e r N e c k a m (died 1217). T h e influence of this simple
device was far-reaching in every respect: it m a d e it possible to navigate on the
high seas. T h e principle of the compass has r e m a i n e d unchanged, the device is still
in full use. T h e invention of the compass is characteristic of m a n y later develop-
ments in magnetism: seemingly marginal effects turned out to be very i m p o r t a n t
and to have had a t r e m e n d o u s impact on later technological developments.
A milestone in the history of magnetism was William Gilbert's De Magnete,

electro- high neutron- critical materials spin structures leo-


magnetic fields diffr, phenom metals semi- amorphous micromagnetism mac netism
radiation , magn. M6ssbauer para- alloys cond. mogn. J spin I domains stellar
I inductio~moment ;NI~R ;mogn. ;sp;e'Ig° ne i I gl~ss I I bb,es
~' magnetism

ro~ F?Curi~
I Jalloys ~-
Maxwell I L oxides ~ ~
Faraday,
I

A~ere,1 8 0 O r ~
ert' tedr2[lmeognete
1269 Peregrinus de Maricourt J
1200 Neckarn describes compass J
J 800 b ~ r l d : magnesian stones
Fig. 1. Development of the modern branches of magnetism from a common root. A few names and
dates are indicated to mark some of the most crucial moments in this development. The modern fields
of magnetism, ranging from basic entities like magnetic fields and particles to more complex
ensembles, emanate in quite a straightforward way from a few basic branches. A central position is
reserved to the various classes of materials, reflecting the central position of materials in magnetism
research.
MAGNETISM AND MAGNETIC MATERIALS 5

Magneticisque Corporibus, et de Magno Magnete Tellure (1600, "Concerning


Magnetism, Magnetic Bodies and the Great Magnet Earth") which summarized
all the available knowledge of magnetism up to that time, notably that of Petrus
Peregrinus de Maricour (1269). In addition Gilbert describes his own experiments:
he measured the direction of the magnetic field and its strength around spheres of
magnetite with the aid of small compass needles. For this purpose he introduced
notions like magnetic poles and lines of force. Gilbert found that the distribution
of the magnetic field on the surface of his sphere or terella ("microworld") was
much like that of the earth as a whole and concluded that the earth is a giant
magnet with its two magnetic poles situated in regions near the geographical
poles. This observation made him the founder of geomagnetism. Gilbert's work
not only strongly influenced the later development of magnetism, but also
contributed to the development of the idea of universal gravitation: it was
believed, for some period of time before Newton, that the planets were held in
their orbits by magnetic forces in some form or other. Gilbert also discovered that
lodestone, when heated to bright red heat, loses its magnetic properties, but
regains them on cooling. In this way he anticipated the existence of the Curie
temperature. For more than two centuries after Gilbert little progress was made
in the understanding of magnetism, and its origin remained a mystery.
The early nineteenth century marked the beginning of a series of major
contributions. Hans Christian Orsted discovered in 1820 that an electric current
flowing in a wire affected a nearby magnet. Andr6 Marie Amp6re established
quantitative laws of the magnetic force between electric currents and demon-
strated the equivalence of the field of a bar magnet and that of a current-carrying
coil. Michael Faraday discovered magnetic induction in 1831, his most celebrated
achievement, and introduced the concept of the magnetic field as an independent
physical entity. Guided by his feeling for symmetry and harmony he suspected
that an influence of magnetic fields on electric conduction should exist as a
counterpart to Orsted's magnetic action of currents. After a long period of
unsuccessful experiments with static fields and stationary magnets, he discovered
the induction effects of changing fields and moving magnets. This line of in-
vestigation culminated in Maxwell's equations, establishing the synthesis between
electric and magnetic fields.
Progress in the understanding of the microscopic origin of magnetism was
initiated by Amp6re, who suggested that internal electric currents circulating on a
molecular scale were responsible for the magnetic moment of a ferromagnetic
material. Amp6re's hypothesis enabled Wilhelm Eduard Weber to explain how a
substance may be in an unmagnetized state when the molecular magnets point in
random directions, and how they are oriented by the action of an external field.
This idea also explained the occurrence of saturation of a magnetic material, a
state reached when all elementary magnets are oriented parallel to the applied
field. This line of thinking led to the studies of Pierre Curie and Paul Langevin on
paramagnetic substances, and also to the work of Pierre Weiss (1907) on ferro-
magnetic materials. Pierre Curie described the paramagnetic substances as an en-
6 u. ENZ

semble of uncoupled elementary magnetic dipoles subjected to thermal agita-


tion, the orienting action of the external field being counteracted by the thermal
agitation. Such a description was also applied successfully to ferromagnetic
materials at temperatures higher than the Curie temperature. The modern dis-
cipline of critical phenomena is, for the time being, the end point of this branch.
Weiss, in turn, postulated the existence of a hypothetical internal magnetic field
of great strength in ferromagnets, resulting in a spontaneous magnetization even
in the absence of an external field. Amongst his other contributions is the notion
of a magnetic domain, a small saturated region inside a ferromagnet, and the
notion of domain walls. Weiss's work can be viewed as the starting point of the
branch leading to the modern disciplines of micromagnetism and domain theory,
and also as the point of departure of N6el's work on interactions, leading to the
fields of ferrimagnetism and antiferromagnetism, including the actual disciplines
of spin structures and spin glasses. A branch of its own, the study of the magnetic
aspects of particles is perhaps a less obvious offshoot from the common source,
but it nevertheless forms a very important part of magnetism. Indeed fields like
electron spin resonance and nuclear spin resonance, M6ssbauer spectroscopy and
structure analysis by neutron diffraction are at the same time indispensable tools
and important disciplines of magnetism.
The modern disciplines of magnetism as represented at the top of fig. 1 range
from the fundamental entities like fields and particles on the left to more complex
systems on the right. Critical phenomena, magnetic phase diagrams and spin
structures in various materials including spin glasses are important fields of modern
research. The classical discipline of domains, domain walls and micromagnetism is
still being actively studied, and has even received renewed attention stimulated by
the modern investigations on bubbles. The various materials appear in the centre
of the three, thus confirming their central role in magnetism. The few materials
that are named represent just a very small fraction of the magnetic materials
known at the present time. The study of the magnetic properties of materials is
the subject of the present handbook, and the present article is intended to give a
general introduction to the remaining chapters of this work.

2. The role of magnetism in present-day technology and industry

Having outlined the early developments of magnetism as well as the subsequent


accumulation of knowledge on magnetic phenomena and materials, we now turn
to the description of the role of magnetism in present-day technology and
industry. Magnetic materials occupy a key position in many essential areas of
interest to society. The most important of these, which depend in an essential way
on magnetic materials, are the generation and distribution of electrical power, the
storage and processing of information, and of course communication in all its
forms, including telephony, radio and television. Apart from these major fields,
many other industrial machines and devices, including motors for numerous
applications, depend on magnetic materials or magnetic forces. Figure 2 gives a
MAGNETISM AND MAGNETIC MATERIALS 7

Industrally and economicallyrelwant fields of


application of matlneticmaterial=
Function Physicaleffect, Material classes
important
parameters
power generators, high induction
power transformers material,
magnetic induction, silicon-iron sheet,
oriented sheet
recording heads,
medium frequency
transformers,
inductors, Permalloysheet,
particle accelerator amorphousmetals,
ferrites

electric motors, I ~ [ magnetomotive


small motors with permanentmagnets,
permanentmagnets ~ior~teS~linduction
Ticonal,
Ferroxdure~
SmCo-magnets
latching devices,
levitation (trains)

television,radio, self-induction,
resonantband pass ferrites
high Q
filters

unidirectional ferromagnetic garnet crystals,


microwavedevices resonance, hexagonal
gyrators low damping ferrites

storageof digital squarehysteresis I square-loop


information in loop, fast ferrites, plated
cores, plated wires switching time wires, permalloy

massstorage l smallparticlesof
in magnetic remanenceand 3'- Fe203, CrO2, Fe
tapes, magnetic coerciveforce of
discs and small particles
floppy discs and thin magnetic
films thin metallic films,
Co-based

information storage 1 high Faradayand amorphous


in magneto-optic Kerr effect films, MnBi,
stores GdFe

stability and
compact digital mobility of monocrystalline
massstorage bubbles garnet films

Fig. 2. Industrially and economically relevant fields of application of magnetic materials. T h e central
column lists the basic physical effects together with the important parameters. T h e left-hand column
gives the useful functions and the right-hand one the classes of materials most commonly used to fulfil
these functions. The arrows indicate the various interrelations.
8 u. ENZ

survey of the use of magnetic materials in various areas of application. The


left-hand column shows the type of application or the function realized with the
aid of magnetism. The central column lists the various physical effects and the
parameters most relevant to a specific application. The right-hand column displays
the classes of materials used to accomplish the various functions. The variety of
physical effects, applications and materials is impressive, even in this necessarily
incomplete survey. Most of the applications are based on magnetic induction,
magnetomotive forces or the specific properties of the hysteresis loop such as
squareness or coercive field, but other effects like ferromagnetic resonance or the
Faraday effect also find their applications. The useful materials are in general
restricted to those classes which are ferromagnetic or ferrimagnetic at room
temperature and which possess a sizable saturation magnetization. Although
magnetic materials are indispensable for the listed and for many other ap-
plications, their role remains to some extent hidden because the ultimate practical
function is not associated with magnetism. The public at large is often unaware of
the role of magnetism in everyday life. Occasionally one gets the impression that
the same holds even for some professionals!
The economic impact of magnetism is very considerable. Jacobs (1969) esti-
mates the total value of processed magnetic material produced in the United
States in 1967 at about 680 million dollars, which represents about 0.1% of the
American gross national product. In 1976 the corresponding amount was 2140
million dollars (Luborsky et al. 1978, Snyderman 1977). These figures apply to
magnetic materials as such and do not take into account that magnetic materials
are nearly always components of more elaborate products such as electric motors,
transformers, loudspeakers, microwave isolators or computer memories. Jacobs
estimates that for such products a "multiplication factor" of the order of 15 is
appropriate, so that the total economic impact of magnetism was of the order of
1.5% of the American gross national product in 1969. This figure has not changed
much since, and probably applies to other economies as well. In table 1 some
economic data on magnetic materials are given, split up into various classes of
materials to be discussed below.
The largest quantities of magnetic material enter the field of electric power
generation and distribution, a field which is historically the major application of
magnetism. For this function the aim is to reach the highest possible saturation
magnetization and the lowest possible total loss, properties which are best met in
iron-based alloys such as silicon-iron sheet and grain-oriented sheet. More
recently amorphous metals have become competitive for some specific ap-
plications in this field. Although power generation and distribution have now
achieved the status of well-established and mature technologies, progress towards
reduced losses is still going on. As a result the amount of power handled by a
transformer of constant size has increased continuously, resulting in the last 40
years in a tenfold increase in power-handling performance.
A second area of great and rapidly increasing importance is that of magnetic
materials for information storage and processing. The amount of material in-
volved in these fields is much smaller, but their economical significance is larger
MAGNETISM AND MAGNETIC MATERIALS

TABLE 1
Annual magnetic materials market in millions of dollars (not corrected for inflation)

196%1968 1976-1977 1979-1980

US World (17 US World 07 US World 07

Electrical steel 180(27 4400)


Magnetic recording
tapes 180(27 45007 1300(3/
Magnetic discs
and drums 100(~) 900(37 3000(37
Soft ferrites
(communication,
entertainment and
professional) 110(2) 130 535(4) 148 600(4)
Square-loop ferrites 55(2) 70(67 57(7)
Permanent magnets 55(27 1 5 0 (3,6) 170(5) 960(5)

(1/Western world including Japan. (2)Jacobs (1969). (3)Luborski et al. (1978). (4)De Bruyn and Verlinde
(1980). (s)Hornsveld (1980). (6)Snijderman (1977). (7)Electronics, 3, Jan. 1980.

than that of electrical steels. T h e leading e c o n o m i c position was taken over by


information storage about ten years ago. A m o n g s t information-storing materials,
magnetic tapes and discs are the most i m p o r t a n t groups. T a p e and disc techniques
are both based on the same physical principle, the association of a bit of
information with the direction of the magnetization in a small area of the
material. T h e merits of this storage principle are simplicity, p e r m a n e n c e (or
non-volatility) of information, and a high information density per unit surface of
the (film-like) material. T h e progress of these techniques has been entirely
directed towards higher information densities and thus towards lower prices per
bit of information, and further progress in this direction is expected. Square-loop
ferrites or bubble d o m a i n m e m o r i e s are also based on the association of in-
f o r m a t i o n with the direction of the magnetization in the material, in the f o r m e r
case in a small sintered core, in the latter case in a single domain, a bubble, which
is able to m o v e in a single-crystal film. Magnetic cores have been the basic m e t h o d
of information storage in c o m p u t e r s for m o r e than 25 years, but have now
r e a c h e d a level of saturation, which is i m p o s e d by the limitations of handling ever
smaller (and faster) cores. T h e bubble d o m a i n memories, on the o t h e r hand, are
in full d e v e l o p m e n t and m a y find a p e r m a n e n t position in the hierarchy of
information storage. N u m e r o u s other magnetic information storage principles
have been p r o p o s e d and realized, including thin permalloy films and permalloy-
plated wire, but their total e c o n o m i c impact has r e m a i n e d small.
C o m m u n i c a t i o n is the third i m p o r t a n t field of application of magnetic materials.
This field includes telephony, radio and television broadcasting and receivers, and
radar, all of which techniques use m e d i u m to very high frequencies. D u e to their
10 u. ENZ

high resistivity and consequently low eddy-current losses, ferrites are the best
suited materials for communication. In telephony the medium frequency channel
filters are based on resonant circuits using high-O ferrite cores, one of the first
applications of ferrites and the standard technique till now. Every broadcast
receiver contains many ferrite parts, such as inductors, deflection units, line
transformers and antenna rods.

3. Development of some classes of magnetic materials

In this section the development of a few selected classes of magnetic materials is


described. These case histories relate to iron-silicon alloys, ferrites, garnets and
permanent magnets. The emphasis will be on the chronology of events, the
improvement of performance and the technical relevance.

3.1. Iron-silicon alloys

Electrical grade steel is the magnetic material produced in the largest quantities of
all; hundreds of thousands of tons are annually needed by the electrical industry.
The bulk of this material is used for the generation and distribution of electrical
energy and for motors. High magnetic induction and low losses are of prime
importance in these applications. Alloys of iron and silicon meet the above
conditions well and therefore take a prominent position with the product category
of electrical steels. Here we confine ourselves to some historical remarks and to
give some recent figures on silicon-iron alloys.
The starting point of the development of silicon-iron alloys of suitable quality is
marked by the work of Barrett et al. (1900). These authors found that the addition
of about 3% of silicon in iron increased the electrical resistance and reduced the
coercive force as compared to unalloyed iron. The slight reduction in saturation
magnetization due to silicon was far outweighted by the improvement in the other
properties. Some years later, in 1903, the industrial production of these alloys
started in Germany and in the U n i t e d States. The promotion and subsequent
improvement of this technique is, to a considerable extent, the merit of Gumlich
and Goerens (1912). The material was used in the form of hot rolled poly-
crystalline sheets having random grain orientation. Due to the higher permeability
and the reduced hysteresis and eddy-current losses, iron-silicon sheet replaced the
conventional materials within a few years, in spite of the initial difficulties of
production and the higher price, and was used in this form for about three
decades.
However, these random oriented materials were still imperfect because satura-
tion was only reached by applying magnetic fields well above the coercive field,
which limits the useful maximum induction to about 10 kG. The hysteresis loops
of single crystals or of well oriented samples, on the other hand, are nearly
rectangular. Fields slightly higher than the coercive field are therefore sufficient to
drive the core close to saturation. The useful maximum induction is thus higher,
MAGNETISM AND MAGNETIC MATERIALS ll

reaching 15 to 1 7 k G (1.5 to 1.7Wb/m 2) (fig. 3). An ideal transformer would


consist of single crystal sheets oriented such that a closed rectangular flux path
along [100] directions, the preferential directions, results. Progress towards this
goal was achieved by Goss (1935) who showed that grain oriented sheets can be
obtained by certain cold rolling and annealing procedures. This so called Goss
texture is characterized by crystallites having their [110] planes oriented parallel
to the plane of the sheet, with common [100] direction in this plane. The magnetic
properties of such materials are characterized by coercive fields around 0.1 Oe
and maximum permeabilities up to 70.000 (fig. 3). Data obtained by Williams and
Shockley (1949) on a oriented single crystal frame of high purily 3.85% silicon
iron with limbs parallel to [100] directions are included in fig. 3 for comparison.
The coercive field of this single crystal was as low as 0.028 Oe (2.2 A/m) and the
maximum permeability exceeded 10 6. Grain orientation contributed to a further
decrease of the magnetic core losses. Loss figures dropped to about 0.6 Watt per
kg at 60 cycles and an induction of 10 kG (1 Wb/m 2) for commercial grade
oriented silicon-iron sheet. An additional advantage was that maximum induction
up to 17 kG (1.7 Wb/m 2) became practical.
Another more recent step towards higher quality is related to even more
perfect crystalline orientation combined with the introduction of a controlled
tensile stress in the sheet (Taguchi et al. 1974). The composition of the material
remained unchanged i.e. the silicon content is still around 3%. The tensile stress is
introduced by a surface coating consisting of a glass film and an inorganic film
applied on both surfaces of the sheet. This procedure leads, by magnetoelastic

Magnetization (kG)
18
Br.=17 kG
~kO

t l
'°r i/ill'
I I I I

-0.2 0 0.2 0.L 0.6


Magnetic field (Oe)
Fig. 3. Static hysteresis loops of grain oriented silicon-iron sheet (Goss texture). For comparison the
loop of a single crystal (broken line, Williams and Shockley 1949) is shown (after Tehble and Craik
1969).
12 U. E N Z

interactions, to improved properties of the hysteresis loop and thus to still lower
magnetic losses. The improvement of the quality of electrical grade steel since
1880 is shown in fig. 4. In a time span of 100 years the core loss decreased from
8 W/kg to about 0.4 W/kg for an induction of 10 k G and a frequency of 60 Hz.
The innovations described above clearly show up as marked steps: the intro-
duction of silicon-iron alloys after 1900, causing the loss figure to drop from 8 to
2 W/kg and also the use of grain-oriented material in the late thirties. The latter
innovation enabled higher induction ratings: new branches with maximum in-
duction of 15 kG and 1 7 k G appear after 1940. The dramatic increase of the
power handled by a transformer of equal size is also shown.

co. e toss ( W a t t / k g ) transforrnator power { MVA)

6
4.
17 kG

2 15kG 1000
!
10 kG
1 50
0.8 I //
0.6 ' /t
o.4 / I
I
/
0.2 If I I
188o 1900 1920 194.0 1960 1980 2000
year
Fig. 4. Core loss of electrical grade steel (after Luborsky et al. 1978). Since 1880 the core loss
decreased from 8 W / k g to 0 . 4 W / k g at an induction of 1 0 k G and a frequency of 6 0 H z . T h e
introduction of textured sheet around 1940 led to the use of higher m a x i m u m induction (up to 17 k G
(1.7 Wb/m2)). T h e broken line shows the increase in power handled by a transformer of constant size.

3.2. Ferrites

Ferrites are mixed oxides of the general chemical composition MeOFe203, where
Me represents a divalent metal ion such as Ni, Mn or Zn. The crystallographic
structure of ferrites, and also that of the closely related ore magnetite (FeOFe203)
is the spinel structure. Ferrites can therefore be seen as direct descendants of
magnetite (see fig. 1). Many simple or mixed ferrites are magnetic at room
temperature, but due to their ferrimagnetic character the saturation magnetization
is only a fraction of that of iron. The outstanding property of ferrites, which
makes them suitable for many applications, is their high electrical resistivity as
compared to that of metals. Their specific resistivity ranges from 102 to 101° f~ cm,
which is up to 15 orders of magnitude higher than that of iron. In most high
frequency applications of ferrites eddy currents are therefore absent or negligibly
small, whereas at such frequencies eddy currents are the main drawback of
MAGNETISM A N D M A G N E T I C M A T E R I A L S 13

metals, even in laminated form. Such intrinsic properties make the ferrites
indispensable materials in telecommunications and in the electronics industry,
where frequencies in the range of 109 to 1011Hz have to be handled. The potential
usefulness of magnetic oxides for high frequency applications was realized as early
as 1909 by S. Hilpert, who investigated the magnetic properties Of various oxides
including some simple ferrites. In 1915 the crystallographic structure of ferrites,
which had remained unknown until then, was determined independently by W.H.
Bragg in England and S. Nishikawa in Japan. Contributions to the understanding
of the chemistry of ferrites were also made by Forestier (1928) in France. All of
this early work remained without a direct follow-up; at that time there was
apparently no technological need yet for such materials. The situation remained
unchanged until 1933, when Snoek of the Philips Research Laboratories started a
systematic investigation (Snoek 1936) into the magnetic properties of oxides. In
the same period of time ferrites were independently investigated by Takai (1937)
in Japan. Snoek's working hypothesis in his search for high permeability materials
was to look for cubic oxides which, for symmetry reasons, could be expected to
have a low crystalline anisotropy. Simultaneously he aimed at finding materials with
low magnetostriction values to minimize the adverse effects of the unavoidable
internal stresses present in polycrystalline materials. Snoek's approach turned out to
be fruitful: he found suitable materials in the form of mixed spinels of the type
(MeZn)Fe204, where Me is a metal of the group Cu, Mg, Ni or Mn. Permeabilities up
to 4000 were reached (Snoek 1947).
Snoek's achievement may again have remained of more academic interest, but
this time there was a clear demand for magnetic materials from the telephone
industry, which felt the need to improve the load coils of their long-distance lines
and to use bandpass filters based on low-loss magnetic materials. Ferrite inductors
proved to be well suited for these purposes, and so ferrites and telephone
technology developed in close cooperation. Six (1952) was the inventive and
leading promotor of this development, which did not, however, proceed without a
great deal of effort from chemists, physicists and electrical engineers, who
cooperated in achieving adequate material properties and practical technical
designs.
Before 1948, when most of this work was done, little was known about the
cause of the low saturation magnetization of ferrites or of the origin of their
anisotropy. This changed when N6el (1948), who had already explained the
behaviour of antiferromagnets, introduced his concept of partially compensated
antiferromagnetism, which he called ferrimagnetism. The essential point of N6el's
explanation is the antiparallel orientation of the spins of the ions in the two
sublattices, octahedral and tetrahedral, of the spinel structure. N6el's model
revealed directly the cause of the low saturation of ferrites. The work of Verwey
and Heilman (1947) on the distribution of the various ions over these lattice sites
was undoubtedly of great help. N6el's model was directly verified by neutron
diffraction work done by Shull et al. (1951) only a few years after the invention of
this powerful method of analysis. Further proof was derived from a study of the
temperature dependence of the saturation magnetization in some spinel ferrites
14 u. ENZ

(Gorter et al. 1953). A n o t h e r fundamental mechanism, ferromagnetic resonance,


was extensively studied in ferrites (Snoek 1947) shortly after its discovery by
Griffiths (1946).
The main difficulty encountered in the early use of ferrites was their high level
of magnetic losses and disaccommodation. It was found that even in the absence
of eddy currents there were still appreciable residual losses, which prevented the
design of high quality resonance circuits. It is now well known that quite a large
number of electronic and ionic relaxation processes can cause magnetic after-
effects in magnetic materials subjected to alternating fields. These processes are
the main cause of the losses in magnetic cores. Snoek (1947) tried to minimize
these after-effects by controlling the presence of relaxing ions with the aid of
special sintering procedures in suitable oxidizing or reducing atmospheres.
A n important insight concerning the use of ferrites in resonant filter circuits was
that the relevant quantity to be considered is the ratio between the loss factor
tan6 and the initial permeability/x rather than tan~ as such. This ratio can be
controlled by introducing air gaps in the magnetic circuits. Therefore a high initial
permeability is as important as a low loss factor. A reduced loss level in the ferrite
material made it possible to reduce the physical size of the inductors, a successful
development vizualized in fig. 5. Ferrite cores having about equal quality factors
are shown in a sequence ranging from 1946 to 1974. The volume of the inductors
was reduced by a factor of 32 during this time span. Compared with an air coil of

Fig. 5. Development of pot cores between 1946 and 1974. Reduction of the loss level of ferrite
materials led to the reduction of the physical size of these components, all of which fulfil the same
technical function. The quality factor Q of the ferrite components remained, as indicated, about
constant during this time span. An air coil and a Fernico coil of much lower quality, representing the
state of the art in 1936 and 1939, are shown for comparison. The total reduction in volume is nearly a
factor of 400.
MAGNETISM AND MAGNETIC MATERIALS 15

much lower quality representing the state of the art in 1936, the reduction in
volume is nearly a factor of 400. Similar, albeit less spectacular progress was also
made in another ferrite application, i.e., that in power transformers where a high
saturation induction and low hysteresis losses are of principal importance (see fig.
6). The total losses are shown to be reduced by a factor of three, while the
maximum usable inductance nearly doubled in the indicated period of time. This
type of material, a high saturation MnZn ferrite, is at present finding increasing
application in switched-mode power supplies for small to medium power levels.

loss P (Watt cm-3)


\
Induction 8 (kG)

0.2 "°"3c'3c5 ...----


tt I "~..3c8
0.1 2

J L i 0
1950 1960 1970 1980
year
Fig. 6. Reduction of loss factor and increase of usable maximum induction of ferrite cores for power
applications from 1950 to 1980.

A new and rather unexpected application of ferrites emerged with the in-
vention of the gyrator, a non-reciprocal network element. The impedance of
signals passing through such an element in the forward direction differs from that
in the backward direction. The gyrator was conceived by Tellegen (1948) on
theoretical grounds as a possible but not yet realized network element. Hogan
(1952) was the first to realize a non-reciprocal microwave device consisting of a
ferrite-loaded waveguide. The physical effect on which the device is based is
analogous to the well-known Faraday effect, i.e., the rotation of the plane of
polarization of a light wave passing through a magnetized body. One form of the
non-reciprocal device thus consists of a circular wave-guide carrying a central
ferrite rod magnetized along its axis. If the rotation of the plane of polarization of
the microwave equals 45 ° and coupling in or out also occurs with an offset angle of
45 ° the microwaves pass through the device in the forward direction while
propagation in the backward direction is suppressed. A similar device functions as
a microwave switch (see fig. 7) controlled by the bias field. Other devices based on
this principle, which have found wide application in microwave technology, are
circulators, resonance isolators and power limiters.
The discovery that some polycrystalline spinel ferrites can have a rectangular
hysteresis loop and therefore can be used as computer memory elements was of
paramount importance for computer technology. Until 1970 nearly all main-frame
16 U. E N Z

Rectangular
waveguide
Cylindrical
waveguide'-k~" ~ ] I
Rectangular ~ ~..,/', IX I I
waveguide -x~.r~.~~ ,.~1"~,, !/ L___ I
". L-
~ ~ I t ', , "':."l'-'~4"-"
Fig. 7. Microwave switch based on the rotation of the plane of polarization of microwaves propagating
along the magnetized ferrite rod in the central cylindrical part of the device.

Fig. 8. Core array of a ferrite core m e m o r y in 3D organization showing word, bit and sense lines. T h e
ferrite core matrix is shown together with the preceding storage technology, based on electron tubes,
and the succeeding technology, the semiconductor memory.
MAGNETISM AND MAGNETIC MATERIALS 17

computer memories consisted of ferrite cores, so that it is fair to say that the
whole computer development was closely connected with the development of the
ferrite core memory. A r o u n d 1968 the yearly,world production of ferrite cores
was about 2 x 10 l° cores (Jacobs 1969). J.W. Forrester and W.N. Papion, both at
that time at MIT, are generally considered to be the inventors of the principle of
coincident current selection of a core (Forrester 1951). The first square-loop cores
consisted of nickel-iron alloys, the switching speed of which suffered from an
inherent limitation due to eddy currents. Both Forrester and Rajchman (1952)
suggested the use of non-metallic cores to avoid this shortcoming. At about the
same time, Albers-Schoenberg (1954) observed the square-loop properties of
some ferrite compositions. A link with system requirements was immediately
made. This revolutionary challenge materialized in Whirlwind I, the first experi-
mental computer based on a ferrite core memory, built at the Lincoln laboratory
at MIT in 1953. Figure 8 shows a wired ferrite core matrix consisting of 1024
ferrite cores. A stack of such matrix planes forms the memory. Two individual
cores are shown in fig. 9.
Ferrites also p l a y e d an unexpected role in particle accelerators constructed to
study elementary particles (Brockman et al. 1969). The operation of these large

Fig. 9. Two individual ferrite memory cores (20 mil and 14 mil) are shown on the wings of a fly.
18 U. E N Z

machines is based on accelerating units consisting of large transformers designed


as resonance cavities. These elements accelerate the charged particles (protons) by
feeding energy into the particle beam circulating in the machine. The stations act
in such a way that the particle beam represents the secondary winding of the
transformers. During one acceleration sequence the increasing frequency has to
be followed by controlling the self-induction of the core of the transformer with
the aid of a bias field. Again, the demand for low losses, especially low eddy
current losses, favoured ferrites above other materials for this application, and so
ferrites entered this field as essential elements and played a continued role in the
subsequent stages of the development of these machines. Figure 10 shows an
acceleration unit of the alternating gradient synchrotron in Brookhaven. The
synchrotron contains 12 of such cavities, and each unit Contains about 500 kg of
ferrite.

Fig. 10. Accelerator unit of the alternating gradient synchrotron at Brookhaven, containing a hollow
cylindrical ferrite core assembled from a large number of rings. The total weight of the core is about
500 kg (after Brockman et al. 1969).

Apart from these few examples of specific implementations of ferrites we recall


that the bulk of ferrite material is used in telecommunication and consumer
applications, with roughly equal turnover in these two fields. The main consumer
products are television and radio sets, in which such parts as line transformers,
deflection coils, tuners and rod antennas contain ferrite materials. About 0.7 kg of
ferrite enters a black-and-white television set, and about 2 kg a colour set. Figure
11 gives an impression of the diversity of ferrite products for such applications.
MAGNETISM AND MAGNETIC MATERIALS 19

Fig. 11. Various ferrite componentsas used in radio and televisionsets.

Hand in hand with the implementation of ferrites and with the tailoring and
perfection of their technically relevant parameters, the investigation of their
fundamental properties continued, resulting in a considerable deepening of the
understanding of their chemical and physical properties. The main lines of
investigation concerned: (a) crystal structures, chemical miscibility regions and
preferential site occupation of the various ions in the spinel lattice; (b) the
experimental determination of data relating to spontaneous magnetization, Curie
temperature, anisotropy and magnetostriction constants; (c) micromagnetic pro-
perties such as domain walls and domain configuration in polycrysta!line and
monocrystalline materials; (d) the dynamics of the magnetization process, and
damping and resonance phenomena; and (e) the theoretical description, dis-
cussion and understanding of these properties in atomic terms, i.e., the arrange-
ment of the ionic magnetic moments in sublattices, the quantum mechanics of
magnetic interactions between localized moments, the spin-orbit interaction and
dipole-dipole interaction as a cause of anisotropy.
These methods of studying the properties of ferrites have acquired ~he status of
a scientific standard, i.e., a paradigm, which has since been applied to many other
materials. Garnets and hexagonal ferrites are examples of materials of industrial
importance discovered and investigated along such lines. Other compounds of as
yet more academic interest include sulphospinels and rare-earth chalcogenides.
20 U. ENZ

3.3. Garnets

The prototype of the family of garnets is the compound Mn3A12Si3012, a mineral


and esteemed gemstone occurring in nature. The crystal structure of garnets is
cubic with three different types of sublattices occupied by the three metals of the
above compound. The garnets existing in nature often contain other ions as well
and are in most cases non-magnetic. It is interesting to note that it was precisely
the ferrimagnetic properties, discovered by Pauthenet (1956) and Bertaut et al.
(1956) of some synthetic rare-earth iron garnets that attracted attention and
opened up a new field of research in magnetism. Since then, a wealth of scientific
and technological information on garnets has been produced. For more than two
decades, publications on garnets have been appearing at a rate of about 200
papers a year, so that garnets and especially yttrium iron garnet (YIG), are now
amongst the best known magnetic materials. One may agree with J.H. van Vleck
who compared the role of Y I G for magneticians to that of the fruit fly to
geneticists. Pioneering work on the chemical, crystallographic and magnetic
properties of garnet was done by Geller et al. (1957), while Pauthenet (1957)
investigated the magnetism of mixed rare-earth iron garnets, Re3FesO12, some of
which were found to have compensation temperatures of the magnetization.
These findings demonstrated directly their ferrimagnetic character and indicated
that the magnetic moment of the rare-earth sublattice is oriented oppositely to the
resulting moment of the two iron sublattices. The total magnetization of garnets is
therefore relatively low as compared with that of magnetic metals or ferrites. The
Curie temperatures, dominated by the iron-iron interaction, are of the order of
300°C, reflecting the same type of interaction mechanisms as those active in
ferrites. Because the magnetization of garnets is lower than that of ferrites,
garnets have not found bulk applications competing with ferrites, such as in
transformers, coils, etc.
One of the outstanding properties of YIG, which was discovered by Spencer et
al. (1956) and Dillon (1957) shortly after the publication of the first papers on
garnets, is its extremely low ferromagnetic resonance linewidth. Values of the
linewidth A H of some oersteds were then measured, but subsequent improve-
ments of crystal quality and purity yielded figures as low as A H = 0 . 1 0 e (8 A/m)
at 10 MHz for carefully polished samples of Y I G (LeCraw et al. 1958). These
exceptional resonance properties made Y I G very useful as a microwave device
material, with loss figures one or two orders of magnitude lower than those of
corresponding spinel ferrites. Moreover, Y I G also proved to have acoustic losses
lower than quartz, thus opening up prospects for magneto-acoustic devices such as
adjustable delay lines. Small amounts of rare-earth ions substituted into Y I G
produced a dramatic increase in both anisotropy and linewidths. Such substituted
materials were ideal objects on which to study the fundamentals of anisotropy
(Kittel 1959) and relaxation processes (Dillon 1962, Teale et al. 1962), studies
which contributed considerably to progress in the understanding of these
mechanisms.
MAGNETISM AND MAGNETIC MATERIALS 21

Garnets are also outstanding in their optical properties: Y I G shows a low


optical absorption in the range of visible light, so that layers up to several microns
thick are transparent. With the aid of the magneto-optical Faraday effect, domain
structures can be directly observed (Dillon 1958). Y I G m o r e o v e r exhibits a
"window" of extremely low optical absorption in the infrared region. Optical
absorption coefficients as low as c~ = 0.03 cm -1 for wavelengths between A =
1.2 ~ m and A = 4.4 p~m have been observed. The Faraday rotation of some garnets
is large enough to m a k e the material suitable for magneto-optical devices such as
light modulators and magneto-optical memories. In particular, garnets containing
bismuth show a very high specific Faraday rotation, e.g., 0 = 3°/~m at A = 0.5 ~m
for Y2.6Bi0.4Fe4Oa2 (Robertson et al. 1973). Bismuth-doped garnets have also
proved to be suitable for use as fast switching magneto-optical display components
(Hill et al. 1978). The non-magnetic yttrium aluminium garnet (YAG), was found
to be an excellent laser host (Geusic et al. 1964).
Garnets, especially Si-doped Y I G , are also interesting for their photomagnetic
properties, i.e., the light-induced change of magnetic properties as a consequence
of a light-stimulated redistribution of electrons or Fe 2+ ions (see Teale et al. 1967,
Enz et al. 1971).
Even greater prominence was achieved by the granets in their application as the
leading magnetic bubble device material. Two basic contributions, both m a d e by
Bobeck, opened up this new field. The first was the invention of the principle of
high-density information storage with the aid of bubble domains (Bobeck 1967);
the second was his observation of a growth-induced uniaxial anisotropy energy in
garnets (Bobeck et al. 1970). B o b e c k ' s basic idea was to associate a bit of
information, a digital one or zero, with the presence or absence of a bubble at
some defined location and time. Bubbles are cylindrical domains of reversed
magnetization occurring in nearly magnetized thin films having a sufficiently large
uniaxial anisotropy. The dimensions of bubbles are mainly determined by the
length l = trw/47rM2s, a material constant depending on the specific Bloch wall
energy O-w and the saturation magnetization Ms. By controlling the magnetization
of the garnet material the bubble diameters can be adjusted in a wide range from
submicron size to tens of microns. The first observation of stable bubbles, which
were in fact submicron bubbles, was made by Kooy et al. in 1960 in the hexagonal
material BaFe120~9, a material with a rather high magnetization and a large
uniaxial anisotropy. D u e to the microscopic dimensions of the bubbles, the
packing density of the bubbles and thus the density of information can be very
high; values of 1 0 6 bits cm -2 have been achieved, and densities higher than 1 0 7 bits
c m -2 a r e considered to be attainable in the future. An extensive review of the
physical and chemical properties of garnets has been given by Wang (1973) and a
survey of garnet materials for bubble devices has been published by Nielsen
(1976).
The advent of bubble domain devices stimulated work in the field of crystal
growth, both of bulk garnet crystals and of thin monocrystalline films grown by
liquid phase epitaxy, known as the L P E method. Flux-grown bulk crystals were
22 U. E N Z

first prepared by Remeika in 1956, mainly for the purpose of studying fundamen-
tal properties. This art developed rapidly and the growth of large crystals of high
perfection was finally mastered (Tolksdorf et al. 1978). L P E growth experienced a
similar evolution, starting with the work of Shick et al. 1971, who grew films
suitable for bubble devices. This method of preparation, which is an extension of
flux growth, has now become the standard method of growing bubble device films.
The growth procedure is as follows: a carefully polished wafer of a non-magnetic
garnet crystal is immersed in a liquid consisting of a flux and the dissolved garnet.
If the lattice misfit is controlled, the magnetic garnet grows isostructurally and
without dislocations on the substrate. The films made in this way meet high
standards of quality and reproducibility: composition and film thickness can be
controlled within narrow limits and the remaining level of defects is low, which is
reflected in the low values of the coercive fields achieved. Figure 12 shows bubbles
and stripe domains in a Y G d T m epitaxial garnet film as observed with the aid of the
Faraday effect. Figure 13 shows an overlay Y-bar structure of a shift register as
processed on a garnet film (Bobeck and Della Torre 1975).
A second scientific discipline which was greatly stimulated by the success of
bubbles is that of the dynamics of domain walls and bubbles. Domain wall motion

U
• • •

• •
• •

• •

.. • •
i
Fig. 12. Bubbles and stripe domains in a Y G d T m epitaxial garnet film 7 Ixm thick. The bias field is
near to the run-out field. The bubble diameter is about 7 ixm (after Bobeck and Della Torre 1975).
MAGNETISM AND MAGNETIC MATERIALS 23

Fig. 13. Bubbles moving in a Y-bar pattern of a bubble shift register. Some propagation loops and the
bubble generator (heavy black square) are shown (after Bobeck and Della Torre 1975).

has been studied in the past in the context of magnetization processes of magnetic
bodies caused by domain wall displacements. A detailed understanding of the
dynamic wall properties had not been reached along such lines owing to the
extreme complexity of the process. This situation changed when perfect mono-
crystalline layers of garnets for bubble devices became available, which made it
possible to study the motion of domain walls and bubbles by direct optical
observation. As a result of this development, the body of experimental data on
wall dynamics has greatly increased and the theoretical understanding of the
mechanisms involved has deepened considerably. Amongst the new results is the
insight that there is an upper limit to the velocity of a domain wall, a velocity limit
which also marks a limit to the speed, i.e., the data rate of bubble devices.
Extended reviews on the subject of wall and bubble dynamics have been pub-
lished by Malozemoff and Slonczewski (1980) and also by de Leeuw et al. (1980).
24 U. ENZ

3.4. Permanent magnets

Permanent magnets, i.e., the magnesian stones, marked the beginning of mag-
netism. It is interesting to observe that, at present, permanent magnets are still
being investigated, improved and increasingly applied. They are essential to
modern life as components of a wide variety of electromechanical and electronic
devices. It has been estimated that the average home contains more than fifty
permanent magnets and every car uses an average about eight of them. The
applications of permanent magnets range from loudspeakers, small electric motors
and generators, door latches and toys, to ore separators, water filters, electric
watches and microwave tubes. The function of a permanent magnet in these and
other applications is to generate a magnetic field in an air gap of a magnet system.
The air gap may either be fixed to accommodate moving electric conductors which
exert external forces, a function performed by loudspeakers and electric motors,
or it may be variable as it is in movable armatures on which the magnet exerts the
force. The latter application is found in door latches, relays, telephone sets,
magnetic levitation and contactless couplings between rotating shafts. Another
typical application of permanent magnets is the alignment of an object by exerting
a magnetic torque on it, as in a compass. A special application is found in electron
tubes where permanent magnets are used for controlling the orbits of electron
beams or for focussing them. Table 2 shows a number of functions performed by
permanent magnets together with the corresponding applications. The four func-
tions described cover the main applications of permanent magnets; they include
the conversion of electrical into mechanical energy and vice versa, and the
exertion of mechanical forces on material bodies and on moving charge carriers.
Figures 14 and 15 give an impression of two of the most common applications of
permanent magnets, the loudspeaker and the small motor (after Zijlstra 1976).
The early development of permanent magnet materials proceeded entirely by
trial and error. Nevertheless 100 years ago bar and horseshoe magnets made from

TABLE 2
Typical functions and applications of permanent magnets with some examples of machines, devices
and components (after Zijlstra 1974)

Function Application

Conversion of electrical into mechanical Small electric motors, dynamos, loudspeakers,


energy and vice versa microphones, eddy-current brakes,
speedometers, magnetos
Exerting a force on a Relays, couplings, bearings, clutches, magnetic
ferromagnetically soft body chucks and clamps, separators (extraction
of iron impurities, concentration of ores)
Alignment with respect to a field Positioning mechanisms (e.g. stepping motors),
compasses, some ammeters
Exerting a force on moving charge carriers Magnetrons, travelling-wave tubes, some
cathode-ray tubes, Hall plate
MAGNETISM AND MAGNETIC MATERIALS 25

Fig. 14. Cut-away view of a loudspeaker containing a Ferroxdure ring.

Fig. 15. Cut-away view of a windscreen wiper. The stator field is provided by two Ferroxdure
segments.
26 U. ENZ

carbon steel were well known and widely used. Since then an unparallelled
development has taken place; new permanent magnet materials have been
discovered and the existing ones improved. The foundations for the scientific
understanding of permanent magnets have also been laid in this period.
The relevant figure of merit expressing the quality of a permanent magnet is its
maximum energy product (BH).... This figure describes the ability of a per-
manent magnet to withstand the influence of a counteracting magnetic field.
Moreover the energy product is a measure of the useful magnetic flux that can be
produced by the magnet in a given volume. Figure 16 shows the magnetic flux
density B plotted as a function of the magnetic field H, i.e., the well-known
hysteresis loop. The shaded area represents the (BH)max product; the optimum
working point of the loop is that which defines the largest area. As an illustration
of the achievements of the past, the energy product is plotted in fig. 17 as a
function of time, starting in 1880. Since then, the (BH)max values have increased
by a factor of more than 100. This great achievement was not realized by
improving a single material, but rather by the discovery of new classes of material.
Each new finding was followed by a period of technological improvement, which
in turn was followed by a new discovery. The figure gives the top values reached
in any year. The sequence starts with carbon steels and tungsten steels at the end
of the last century and is continued with cobalt-containing steels around 1920.
A major advance in magnetic materials was made in 1932, when the develop-
ment of the Alnico magnets started with Mishima's AINiFe alloy (Mishima 1932).
In the illustration the various members of this family are indicated as Tic II, Tic G
(Jonas et al. 1941) etc. The coercive force of Alnico magnets was essentially
doubled as compared with earlier materials; these magnets were the first to be
truly p e r m a n e n t under adverse conditions such as stray fields, shock and elevated
temperature. The magnetic and mechanical hardness of the Alnico alloy is due to

Fig. 16. Hysteresis loop showing the optimum working point of a permanent magnet. The hatched
rectangle represents the maximum energy product of the material.
MAGNETISM AND MAGNETIC MATERIALS 27

(BHlrnax (MG Oe)

100 //
/
/
50 /

(Sin, Pr) Cos (.1/


20 Sm C o 5 / ~

10
5
// / oFxd 330
2 // tlTic lI
1 //M is~j.I, el* FxdlO0
0.5
,/ I W-steel
0.2 C-steel

0.1
i i i i / I
1880 1900 1920 19~0 1960 1980
= year
Fig. 17. Development of the maximum energy products of permanent magnets between 1880 and 1980.
Most data are due to Van den Broek and Stuijts (1977).
7

a thermal treatment leading to the precipitation of a second phase in a finely


dispersed form. This development culminated with Ticonal XX, an Alnico alloy
hardened in the presence of a magnetic field which leads to the precipitation of
oriented second-phase particles of elongated shape (fig. 18). The maximum energy
product reached was about 11 M G O e (90 kJm -3) (Fast a n d r e Jong 19591)/.
The next breakthrough in the development of high energy product materials
was made with rare-earth transition metal compounds. The systematic in-
vestigation of the physical and magnetic properties of these alloys and compounds
started around 1960 at Bell Laboratories (Nesbitt et al. 1961). However it was
Strnat who realized the potential of these compounds for high energy-product
magnets and vigorously promoted their development (Strnat et al. 1966). In 1969,
Das of the Raytheon Company announced that he had made a magnet having an
energy product of 20 M G O e (160 kJm -3) by sintering SmCos. Even higher values,
up to 35 M G O e (280 kJm-3), were recently reported to have been obtained in a
related material of the formula (RE)2(CoFe)17 (Wheeler report, 1979)*. This
quasi-binary intermetallic phase was again prepared by sintering. It is interesting

* Wheeler Associates, Inc., 1979, Rare Earth-Cobalt permanent magnets, Elizabethtown, Kentucky,
USA.
28 U. ENZ

Fig. 18. Micrograph of Ticonal XX (after De Vos, thesis, Delft, 1966).

to note in fig. 17, that the energy product, plotted as a function of time, follows an
approximately linear dependence throughout the period reported. The material
quoted last yields the top value of the energy product reached up till now. It is
possible to reach still higher values of (BH)max? Rathenau (1974) discussed this
1 2
question in detail and showed that the limit of (BH)max is given by ~Bs for any
material, provided the coercive field and the anisotropy can be made strong
enough. The saturation magnetization of iron or iron cobalt alloys is high enough
to allow for a further increase of (BH)max by a factor of four.
In addition to the materials discussed, which follow a straight line, fig. 17 also
contains materials indicated as Fxd 100 and Fxd 300. The energy products of
these materials fall clearly below the general trend. They are typical represen-
tatives of the family of low cost permanent magnet materials having a medium
energy product, the hexagonal oxides. If we plot the price per unit energy product
of various materials (fig. 19) we observe another systematic correlation, the
continuous progress made in improving the economy of permanent magnets. In
this plot the hexagonal oxides occupy a leading position and are the end point of a
long development. The importance of ferrimagnetic hexagonal oxides for per-
manent magnets was first pointed out in 1952 by Went et al. A detailed account of
these magnetically hard materials, called Ferroxdure, and their history has been
given recently by Van den Broek and Stuijts (1977). The opinion expressed at the
time of discovery was that these new materials were of great economic im-
portance. That opinion has been fully confirmed: the total world production of
magnetically hard f e r r i t e s - which, in composition and crystal structure, all belong
MAGNETISM A N D M A G N E T I C M A T E R I A L S 29

l 100
T S'mCo5
50
price per ' P t Co
unit of
energy 20
C-steel
•___. W.-steel Tic xx /
I
•~ ,~,~ Co-steel
10
Q
%
5 Tic ff X Tic GG

2 \NPZFxd 100
Tic \
\
1 \
\
0.5 ~ F x d 330
\
0.2

0.1
1800 '90 1900 '10 '20 '30 %0 '50 '60 '70 '80 '9 2 0
= year
Fig. 19. Price per unit of energy product (after Rathenau 1974).

to the same g r o u p - is now estimated at about 100,000 tonnes a year (1980), with
a value of some 400 million dollars. Figure 20 shows how these ferrites have
acquired an ever increasing share of the world production of p e r m a n e n t magnets,
measured in tonnes per annum.
C o m p a r e d with other materials for p e r m a n e n t magnets, Ferroxdure is charac-
terized by an exceptionally high coercivity, combined with a r e m a n e n c e which,
though not very high, is valuable for practical purposes. With such a material it
became possible to produce magnets of shapes such that would have almost
completely demagnetized themselves if made of a different material. Typical shapes
were flat ring magnets, magnetized perpendicular to the plane of the ring, or
transversely magnetized rods with m a n y north and south poles closely adjacent to
each other. Ferroxdure is also highly resistant to external demagnetizing fields, as
encountered in D C motors, for example. These novel properties were exploited
on a large scale, e.g., for making flat loudspeakers and compact D C motors (see
figs. 14 and 15).
The great economic success of Ferroxdure is due in the first place, however, to
the low price per unit of available magnetic energy (fig. 19). The material is
therefore mainly used not so much as a technical i m p r o v e m e n t but m o r e as a
substitute for m o r e expensive components, such as Ticonal magnets in loud-
speakers or stator coils of windscreen-wiper motors. Ferroxdure is inexpensive
because it does not contain any rare material such as nickel or cobalt, and it is
relatively easy to manufacture: it is only necessary to "mix a few cheap oxides"
and to " b a k e them to the right shape".
Finally, Ferroxdure - an oxide - has a high electrical resistivity, so that there are
30 U. ENZ

)roduction (kt)

100

tot
20
10 //~lloys

I L I I I I L I

1900 '10 '20 '30 '/.0 '50 '60 '70 '80


year
Fig. 20. Estimate of the world production of permanent magnets (after Rathenau 1974).

hardly any eddy-current losses. This is an important advantage in radio-frequency


applications and also in certain types of electric motors. A disadvantage is the
relatively high temperature coefficient of the remanence and the coercivity. This
makes the material less suitable for certain professional applications.
In 1954 the material was substantially improved by orienting the crystallites
(Stuijts et al. 1954, 1955). In isotropic material the magnetic moments of the
crystallites, in zero field after saturation, are randomly distributed over a hemis-
phere. In anisotropic Ferroxdure, on the other hand, which is the material now
most widely used, the c-axes and hence the moments after saturation are
approximately parallel. Consequently the remanence is about twice as great and
(BH)max is about four times higher. At the time it was a surprise that the attempts
to produce crystal-oriented Ferroxdure were so successful. It was feared, quite
reasonably, that the orientation of the crystallites, achieved with much difficulty in
the compacted product, would be lost during sintering. The result exceeded all
expectations: the texture was not only preserved but was even greatly improved.

4. Trends in magnetism research and technology

4.1. Magnetism research between physics, chemistry and electronics

On studying the ways to progress in magnetism one observes that magnetism


research has an interdisciplinary character and depends in an essential way on the
MAGNETISM AND MAGNETIC MATERIALS 31

cooperation of scientists and engineers working in fields quite different from each
other. The various disciplines relevant for magnetism research range from fun-
damental theoretical physics through chemistry to electric and electronic
engineering. The cooperation of these disciplines is of such a type that the
preparation and chemical or crystallographical study of new materials may
stimulate physical work, or alternatively that the discovery of a physical effect in
one material may stimulate the search for other classes of material showing
analogous effects. A similar mutual relation is observed between engineering
efforts facing material problems and basic material studies concentrating on the
relevant critical parameters. Last but not least, and most commonly recognized,
the discovery of new materials or of new physical effects may stimulate new
engineering applications.
This special character of magnetism research has consequences for the
organization of research and development laboratories. An organizational struc-
ture concentrating the various disciplines in multidisciplinary units or groups is
probably the most adequate form. Indeed, the study of the case histories of m a j o r
innovations seems to show that those research laboratories which have such an
organization are most likely to produce outstanding results. This applies equally
for university laboratories organized as "material centers", government institutes
and industrial laboratories. In the last category the multidisciplinary approach has
an even stronger weight as engineering aspects are included. Examples of research
successes obtained by the cooperation of experts in different disciplines are easily
at hand. In fact some have already been described in this chapter. The early
industrialization of ferrites is an example in which chemistry, crystallography,
physics and telecommunication technology were equally indispensable for success.
Other examples are the discovery of garnets, the invention of the gyrator or the
development of bubble devices. In all cases the study of materials played a central
role.
Some remarks concerning the growth of magnetism research may be made
here. Before the second world war only few laboratories were active in this field,
and with some famous exceptions, magnetism was not usually a subject of
research at universities. Since then the n u m b e r of laboratories occupied with
magnetism has largely increased. In particular many new or existing university
institutes turned to studies in this field. As a consequence the n u m b e r of
investigations has much increased and more and more detailed studies of materi-
als were made. The industrial laboratories, on the other hand, did not much grow
in n u m b e r or size, so that their share, especially concerning fundamental studies,
has diminished. The few government or national research centers which were
traditionally active in magnetism research maintained their position and continued
their important role.

4.2. Trends in applied magnetism

In this outlook into the future of applied magnetism we try to indicate some
trends which are discernable at this m o m e n t and which will probably be of impor-
tance for some time to come (see, Wijn 1976). We have seen that magnetic materials
32 U. ENZ

are used for a wide variety of different functions such as transformers, various types
of inductive elements or m e m o r y cores. C o m m o n to most of these applications is the
use of materials prepared separately in bulk form. The processed material is then
assembled into the magnetic device as a separate part.
Since some years, however, there seems to be a tendency towards the use of
materials as an integrated part of the device. In m a n y cases the magnetic material
is present in the form of a thin layer, having a monocrystalline, polycrystalline or
amorphous structure. The layers are often structured or shaped by methods
known from integrated circuit technology. The purpose is to reach a miniaturiza-
tion also in the case of magnetic materials and to give it its shape in situ.
Modification or control of the local compositions and the local magnetic proper-
ties of a thin film, e.g., by means of ion implantation, is another aspect of the
tendency described here. A second development is aimed at the control of the
position and displacement of individual domain walls in monocrystalline or
oriented polycrystalline layers. The bubble devices are a good example for this
tendency. Such devices depend, apart from a successful miniaturization, in an
essential way on the mastering of the material properties, the material perfection
and the internal stress distribution.
Some examples of these tendencies in device development will now be given,
starting with the magnetoresistive reading heads used in magnetic recording.
Reading of the information recorded on a magnetic tape is usually achieved by
picking up the stray flux passing the air gap of an inductive reading head. T h e
electric signal induced in the windings of the head is proportional to the rate of
change of the flux. In the magnetoresistive reading head proposed some time ago
(Hunt 1971), the flux itself is measured with the aid of the magnetoresistive effect
in a thin film of Permalloy. The effect depends on the angle 0 between the
direction of the current flowing through the film and the direction of the
magnetization. The latter is modulated by the stray flux of the tape. To obtain a
linear characteristic the equilibrium angle 0 should be 45 °. An elegant solution is
achieved in the so-called " B a r b e r pole configuration" (Kuijk et al. 1975) in which
the direction of the current flow is forced into the desired direction with the aid of
parallel conductor strips (see fig. 21).

i //
Ni'Fe z - - current flow
Fig. 21. Magnetoresistive reading head based on the "Barberpole configuration". The magnetization
of the NiFe film is parallel to the vector M, the current is forced, by conducting bars, into a direction
parallel to the vector I. The optimum angle 0 is 45° (after Kuijk et al. 1975).
MAGNETISM AND MAGNETIC MATERIALS 33

Closely related to this example are the thin film integrated recording heads of
the inductive variety. Here the emphasis is on reading and writing many tracks on
the tape or disc simultaneously. Accordingly the heads are made in large number
in integrated form by a batch process using thin film and photolithographic
techniques (Romankiw 1970). In combination with solid state integrated circuits a
new and very attractive approach of magnetic recording becomes feasible in such
a way (fig. 22). A similar development toward miniaturization can be observed
with inductors. The L-chip, a miniaturized self inductance based on multiple coils
printed on ferrite substrates, is being developed at present.
The second main trend, the control of magnetic domain walls and domain
structures at a micromagnetic level is manifest in bubble domain devices and
domain control in Fe-Si sheet. Both fields have already been discussed in the
present survey. Also thermomagnetic recording can be viewed as an example of
this line. The basic idea of this storage principle has been proposed long ago
(Mayer 1958), but new interest has arisen recently (Berkowitz and Meikle-
john 1975). The information is stored in small regions of reversed magnetization
in a thin magnetic film. Unlike the situation with bubbles, these domains remain
fixed. Reversal of the magnetization is achieved by reducing locally the coercive
force of the film by heating with the aid of a focussed light beam. The information
is read by using the Faraday effect. The bits of information are accessed by
mechanical motion. Film materials include GdFe, MnBi and garnet films. The

Fig. 22. Array of integrated recording heads shown in a fabrication stage prior to cutting off the front
part. The individual heads carry 6 windings (courtesy of W.F. Druyvesteyn, Philips Research
Laboratories).
34 U. ENZ

large Faraday effect of some substituted garnets is also used in a different type of
device, proposed recently (Hill 1980), the integrated light modulation matrix. The
matrix consists of isolated islands etched from a garnet film (fig. 23), which can be
switched individually by a cross bar system. The switching of an island occurs by a
single Bloch wall, and is initiated by local heating of the island.
These examples sufficiently demonstrate the trend mentioned and show that
magnetic materials provide an environment in which rather naturally we can build
and control certain kinds of objects having dimensions in the micron or submicron
range, a region which is generally not readily accessible. The examples also show
that applied magnetism is still very much alive.

~ranspareniresis]Gnce,J

q ' I

d subsirate y-bus
Fig. 23. Light modulating matrix based on switching cells of iron garnet single crystal films, with x-y
addressed resistance network (after Hill 1980).

4.3. Outlook and acknowledgement

In this contribution we have sketched the early historical lines leading to the
present edifice of magnetism, and indicated its importance for modern industry
and society as a whole. Some material classes received more detailed attention in
both the way they developed and their achievements. These materials represent
only a very small fraction of those which have been studied. Moreover, many
achievements in the explanation of material properties and their theoretical
understanding have hardly been touched upon. In this context we note that the
amount of knowledge and detailed information on the properties of magnetic
MAGNETISM AND MAGNETIC MATERIALS 35

m a t e r i a l s has a c c u m u l a t e d t o s u c h a d e g r e e , t h a t it is b e c o m i n g i n c r e a s i n g l y
difficult t o k e e p sight o n t h e w h o l e of i n f o r m a t i o n . P e r h a p s it is w o r t h w h i l e t o
c o n s i d e r t h e f e a s i b i l i t y of t h e i n s t a l l a t i o n of a d a t a b a n k c o n t a i n i n g i n f o r m a t i o n o n
magnetic materials.
T h e a u t h o r w o u l d l i k e to t h a n k P r o f . G . W . R a t h e n a u , P r o f . H . P . J . W i j n , D r .
R . P . v . S t a p e l e , D r . D . J . B r e e d a n d D r . P . F . B o n g e r s of this l a b o r a t o r y , f o r a d v i c e
a n d c r i t i c a l r e a d i n g of t h e m a n u s c r i p t of this c o n t r i b u t i o n .

References

Albers-Schoenberg, E., 1954, J.A.P. 25, 152. Griffiths, J.H.E., 1946, Nature, 158, 670.
Barrett, W.F., W. Brown and R.A. Hadfield, Gumlich, E. and P. Goerens, 1912, Trans.
1900, Sci. Trans. Roy. Dublin Soc. 7, 67. Farad. Soc. 8, 98.
Berkowitz, A.E. and W.H. Meiklejohn, 1975, Hill, B., 1980, IEEE-ED 27, 1825.
IEEE-MAG 11,997. Hill, B. and K.P. Schmidt, 1978, Philips J. Res.
Bertaut, F. and F. Forrat, 1956, C.R. Acad. Sc. 33, 211.
242, 382. Hilpert, S., 1909, Ber. Deutsch. Chem. Ges. Bd
Bobeck, A.H., 1967, Bell Syst. Tech. J. 46, 2, 42, 2248.
1901. Hogan, C.L., 1952, Bell Syst. Tech. J. 31, 1.
Bobeck, A.H. and E. Della Torre, 1975, Mag- Hornsveld, L., 1980, Philips Elcoma Division,
netic Bubbles (North-Holland, Amsterdam). private communication.
Bobeck, A.H., E.G. Spencer, L.G. van Uitert, Hunt, R.P, 1971, IEEE-MAG 7, 150.
S.C. Abrahams, R.L. Barnes, W.H. Grod- Jacobs, I.S., 1969, J.A.P. 40, 917.
kiewicz, R.C. Sherwood, P.H. Schmidt, D.H. Jonas, B. and H.J. Meerkamp van Embden,
Smith and E.M. Waiters, 1970, Appl. Phys. 1941, Philips Tech. Rev. 6, 8.
Lett. 17, 131. Kittel, C., 1959, Phys. Rev. Lett. 3, 169.
Bragg, W.H., 1915, Phil. Mag. 30, 305. Kooy, C. and U. Enz, 1960, Philips Res. Rep.
Brockmann, F.G., H. van der Heide and M.W. 15, 7.
Louwerse, 1969, Philips Tech. R. 30, 323. Kuijk, K.E., W.J. van Gestel and F.W. Gorter,
Das, D.K., 1969, IEEE-MAG 5, 214. 1975, IEEE-MAG 11, 1215.
De Bruyn, R. and G.J. Verlinde, 1980, Philips LeGraw, R.C., E.G. Spencer and C.S. Porter,
Elcoma Division, Eindhoven, private com- 1958, Phys. Rev. 110, 1311.
munication. Leeuw, F.H. de, R. van den Doel and U. Enz,
Dillon, J.F., 1957, Phys. Rev. 105, 759. 1980, Rept. Progr. Phys. 43, 689.
Dillon, J.F., 1958, J.A.P. 29, 539. Luborsky, F.E., P.G. Frischmann and L.A.
Dillon, J.F., 1962, Phys. Rev. 127, 1495. Johnson, 1978, J. Magn. Mag. Mat. 8, 318.
Enz, U., R. Metselaar and P.J. Rijnierse, 1971, Malozemoff, A.P. and J.C. Slonczewski, 1979,
J. de Phys. 32, C1-703. Physics of magnetic domain walls in bubble
Fast, J.D. and J.J. de Jong, 1959, J. de Phys. materials (Academic Press, New York).
Radium 20, 371. Mattis, D.C., 1965, Theory of Magnetism
Forestier, H., 1928, Ann. de Chim. 10e srr. 9, (Harper and Row, New York).
316. Mayer, L., 1958, J.A.P. 29, 1454.
Forrester, J.W., 1951, J.A.P. 22, 44. Mishima, T., 1932, Iron Age, 130, 346.
Geller, S. and M.A. Gilleo, 1957, Acta Cryst. N6el, L., 1948, Ann. de Phys. 3, 137.
10, 239. Nesbitt, E.A., H.J. Williams, J.H. Wernick and
Geusic, J.E., H.M. Marcos and L.G. van Uitert, R.C. Sherwood, 1961, J.A.P. 32, 342 S.
1964, Appl. Phys. Lett. 4, 182. Nielsen, J.W., 1976, IEEE-MAG 12, 327.
Gorter, E.W. and J.A. Schulkes, 1953, Phys. Nishikawa, S., 1915, Proc. Tokyo Math. Phys.
Rev. 90, 487. Soc. 8, 199.
Goss, N.P., 1935, Trans. Am. Soc. Metals, 23, Pauthenet, R., 1956, C.R. Acad. Sc. 242, 1859.
511. Pauthenet, R., 1957, Thesis, Grenoble, France.
36 U. ENZ

Rajchman, J.A., 1952, RCA Rev. 13, 183. Teale, R.W. and D.W. Temple, 1967, Phys.
Rathenau, G.W., 1974, Proc. 3rd Eur. Conf. on Rev. Lett. 19, 904.
Hard Magn. Mat., Amsterdam (Bond voor Teale, R.W. and Tweedale K., 1962, Phys. Lett.
Materialenkennis, P.O. Box 9321, Den Haag, 1,298.
The Netherlands) p. 7. Tebble, R.S. and D.J. Craik, 1969, Magnetic
Remeika, J.P., 1956, J. Am. Chem. Soc. 78, Materials (Wiley, London) 520.
4259. Tellegen, B.D.H., 1948, Philips Res. Rep. 3, 81.
Robertson, J.M., S.W. Wittekoek, Th.J.A. Tolksdorf, W. and F. Welz, 1978, Crystal
Popma and P.F. Bongers, 1973, Appl. Phys. growth of magnetic garnets from high tem-
2, 219. perature solutions, in Crystals Vol. 1
Romankiw, L.T., I.M. Croll and M. Hatzakis, (Springer, Berlin).
1970, IEEE-MAG 6, 597. Van den Broek, C.A.M. and A.L. Stuijts, 1977,
Shick, L.K. and J.W. Nielsen, 1971, J.A.P. 42, Philips Techn. Rev. 37, 157.
1554. Verwey, E.J.W. and E.L. Heilmann, 1947, J.
Shull, C.G., E.O. Wollan and W.C. Koeler, Chem. Phys. 15, 174.
1951, Phys. Rev. 84, 912. Vos, K.J. de, 1966, Thesis, Delft.
Six, W., 1952, Philips Tech. Rev. 13, 301. Wang, F.Y., 1973, in: Treatise on Materials
Snoek, J.L., 1936, Physica, 3, 463. Science and Technology, ed., H. Herman
Snoek, J.L., 1947, New Devel. in Ferromagn. (Academic Press, New York) 279.
Materials (Elsevier, Amsterdam). Weiss, P., 1907, J. Phys. 6, 661.
Snyderman, N., 1977, Electronics News, 28 Went, JJ., G.W. Rathenau, E.W. Gorter and
November. G.W. van Oosterhout, 1951/1952, Philips
Spencer, E.G., R.C. LeCraw and F. Reggia, Techn. Rev. 13, 194.
1956, Proc. IRE 44, 790. Wijn, H.P.J., 1970, Proc. Int. Conf. Ferrites,
Strnat, KJ., G.J. Hoffer, W. Ostertag and I.C. Kyoto.
Olson, 1966, J.A.P. 37, 1252. Wijn, H.P.J., 1976, Physics in Industry (Per-
Stuijts, A.L., G.W. Rathenau and G.H. Weber, gamon, Oxford) 69.
1954/1955, Philips Tech. Rev. 16, 141. Williams, H.J. and W. Shockley, 1949, Phys.
Taguchi, S., T. Yamamoto and A. Sakakura, Rev. 75, 178.
1974, IEEE-MAG 10, 123. Zijlstra, H., 1974, Philips Tech. Rev. 34, 193.
Takai, T., 1937, J. Electrochem. Japan, 5, Zijlstra, H., 1976, Physics in Technology (May),
411. 98.
chapter 2

PERMANENT MAGNETS; THEORY

H. ZlJLSTRA
Philips Research Laboratories
Eindhoven
The Netherlands

Ferromagnetic Materials, Vol. 3


Edited by E.P. Wohlfarth
© North-HollandPublishing Company, 1982

37
CONTENTS

1. I n t r o d u c t i o n . . . . . . . . . . . . . . . . . . . . . . . . . 39
1.1. G e n e r a l p r o p e r t i e s a n d a p p l i c a t i o n s . . . . . . . . . . . . . . . . 39
1.2. T h e h y s t e r e s i s l o o p . . . . . . . . . . . . . . . . . . . . . 4O
2. Suitability criteria for a p p l i c a t i o n s . . . . . . . . . . . . . . . . . . 42
2.1. T h e e n e r g y p r o d u c t . . . . . . . . . . . . . . . . . . . . . 42
2.2. T h e m a g n e t i c free e n e r g y . . . . . . . . . . . . . . . . . . . 46
3. M a g n e t i c a n i s o t r o p y . . . . . . . . . . . . . . . . . . . . . . 49
3.1. A n i s o t r o p y field a n d coercivity a s s o c i a t e d w i t h m a g n e t i c a n i s o t r o p y . . . . . 49
3.2. S h a p e a n i s o t r o p y . . . . . . . . . . . . . . . . . . . . . . 52
3.3. M a g n e t o c r y s t a l l i n e a n i s o t r o p y . . . . . . . . . . . . . . . . . . 53
4. F i n e p a r t i c l e s . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.1. Critical r a d i u s for s i n g l e - d o m a i n particles . . . . . . . . . . . . . . 55
4.2. B r o w n ' s p a r a d o x . . . . . . . . . . . . . . . . . . . . . . 6O
5. C o e r c i v i t y a s s o c i a t e d w i t h s h a p e a n i s o t r o p y . . . . . . . . . . . . . . 6O
5.1. P r o l a t e s p h e r o i d . . . . . . . . . . . . . . . . . . . . . . 60
5.2. C h a i n of s p h e r e s . . . . . . . . . . . . . . . . . . . . . . 64
6. C o e r c i v i t y a s s o c i a t e d with m a g n e t o c r y s t a l l i n e a n i s o t r o p y . . . . . . . . . . 66
6.1. M a g n e t i z a t i o n r e v e r s a l by d o m a i n wall p r o c e s s e s f o r / * 0 / / A > Js . . . . . . 66
6.2. T h e 180 ° d o m a i n wall . . . . . . . . . . . . . . . . . . . . 67
6.2.1. E n e r g y a n d w i d t h of a 180 ° d o m a i n wall . . . . . . . . . . . . 67
6.2.2. T h e e x c h a n g e e n e r g y coefficient A . . . . . . . . . . . . . . 69
6.3. I n t e r a c t i o n of d o m a i n walls w i t h cavities a n d n o n - f e r r o m a g n e t i c i n c l u s i o n s 74
6.3.1. D o m a i n - w a l l p i n n i n g at l a r g e i n c l u s i o n s . . . . . . . . . . . . 76
6.3.2. N u c l e a t i o n of r e v e r s e d o m a i n s at l a r g e i n c l u s i o n s . . . . . . . . . 78
6.3.3. D o m a i n - w a l l p i n n i n g at small i n c l u s i o n s . . . . . . . . . . . . 78
6.4. D o m a i n - w a l l n u c l e a t i o n at surface defects . . . . . . . . . . . . . 80
6.5. I n t e r a c t i o n of d o m a i n walls w i t h the crystal lattice . . . . . . . . . . 81
6.5.1. W a l l p i n n i n g at r e g i o n s w i t h d e v i a t i n g K and A . . . . . . . . . 81
6.5.2. P i n n i n g of a d o m a i n wall by an a n t i p h a s e b o u n d a r y . . . . . . . . 88
6.5.3. N u c l e a t i o n of a d o m a i n wall at an a n t i p h a s e b o u n d a r y . . . . . . . 93
6.5.4. T h i n - w a l l c o e r c i v i t y in a perfect crystal . . . . . . . . . . . . 94
6.5.5. P a r t i a l wall p i n n i n g at d i s c r e t e sites . . . . . . . . . . . . . 98
7. I n f l u e n c e of t e m p e r a t u r e . . . . . . . . . . . . . . . . . . . . . 100
References . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

38
1. Introduction

1.1. General properties and applications

The appearance of permanent-magnet materials such as alnico (Jonas et al. 1941)


and hexaferrite (Went et al. 1951) with much better properties than materials
previously in use was followed by a great increase in the applications of the
permanent magnet. Compared with electromagnets (including power supplies)
permanent magnets offer the advantage of a larger ratio of the useful magnetic
field volume to the volume of the magnet system. Their usefulness is of course
particularly apparent where a constant magnetic field is required. As is widely
known, the constancy of the externally generated field is related to the magnetic
"hardness" of the material, that is to say the extent to which the material retains
its magnetization in opposing fields. In this way the p o l a r i z a t i o n - a n d therefore
the external f i e l d - of a p e r m a n e n t magnet is maintained. A particular example of
an opposing field is the internal field of the poles of the magnet itself. In this case
the demagnetizing action is again unable to destroy the polarization of the
magnet. The present increasing interest in the further development of hard
magnetic materials is explained in part by the growing demand for miniaturization
in modern technology. The problem of heat dissipation is inseparable from
miniaturization, and the substitution of permanent magnets for electromagnets
obviously goes a long way towards solving that problem.
To ensure the most effective development it is desirable to start by investigating
the likely applications of permanent magnets. The next step is to decide on the
criteria that indicate suitability for these applications. These criteria can then
provide a pattern for the production of tailor-made magnetic materials. This calls
for insight into the effects that variation of such properties as remanence and
coercivity has on the suitability of the materials for a particular application, and it
also requires knowledge of the physical background. This will be the main subject
of the present chapter.
Table 1 lists various machines, devices and components in which permanent
magnets are nowadays used. The classification is based on four principles:
- m e c h a n i c a l energy is converted into electrical energy (or vice versa) in the
magnetic field;
- t h e permanent magnet exerts a force on a ferromagnetically soft body;

39
40 H. ZIJLSTRA

TABLE 1
Examples of machines, devices and components using permanent magnets, classified by four functions
which the magnet can perform.

Function Application

Conversion of electrical in'~o mechanical energy Small electric motors, dynamos, loudspeakers,
and vice versa microphone% eddy-current brakes, speedo-
meters, magnetos

Exerting a force on a ferromagneticaUy soft body Relays, couplings, bearings, clutches, magnetic
chucks and clamps, separators (extraction of iron
impurities, concentration of ores)

Alignment with respect to a field Positioning mechanisms (e.g., stepping motors),


compasses, some ammeters

Exerting a force on moving charge carriers Magnetrons, travelling-wave tubes, some


cathode-ray tubes, Hall plates

- t h e permanent magnet is subjected to a directional force exerted by a


magnetic field;
- t h e permanent magnet exerts a force on moving charge carriers, e.g., a beam
of electrons in a vacuum.
In sections 2.1 and 2.2 the two main suitability criteria are discussed which
together cover almost the entire field of applications. They are the maximum
energy product and the maximum change in the magnetic free energy. Ap-
plications not covered by these criteria can be found among the positioning
mechanisms in table 1.
Apart from the fact that the existing applications provide an incentive to search
for better magnetic materials, the converse is of course equally true: better
magnets lead to applications that had not previously been thought of or did not
seem feasible.

1.2. The hysteresis loop

Permanent magnet materials are characterized by high coercivities and high


remanent magnetizations. Before proceeding with the discussion of the structural
parameters that determine the hard magnetic properties, we must first define the
parameters that are generally used to specify the magnetic properties of per-
manent magnets. We employ the International System of Units (SI) in which the
magnetic flux density B is expressed as either

B =/~o(H + M ) ,
or

B =/~oH + J ,
PERMANENT MAGNETS; THEORY 41

where M and J are are the local material contributions to the flux density,
respectively called magnetization and magnetic polarization, and H is the con:
tribution from all other sources and is called magnetic field strength. The
quantities H and M are measured in A m -1 (1 A m -1 = 4~r x 10 - 3 0 e ) . The quan-
tities B and J are measured in Vsm -2 or tesla (1 T = 10 4 Gauss). The vacuum
permeability ~0 is equal to 47r x 10 -7 V s A -1 m -t (or Hm-1). Both expressions for B
will be used in this chapter. Although magnetic polarization is the official n a m e
for J it will often be called magnetization.
If the magnetization M or J of a p e r m a n e n t magnet material is plotted as a
function of the applied field H a hysteresis loop is obtained in which the
magnetization is not a unique function of H, but depends on the direction and
magnitude of previously applied fields. A typical hysteresis loop is shown in fig. 1.
The initial magnetization curve starting at the origin is obtained when the
material is in a thermally demagnetized state. If the m a x i m u m applied field H m is
sufficient to saturate the material the loop is referred to as a saturation loop.
When the applied field is reduced the magnetization decreases to the r e m a n e n t
magnetization J,, which is generally less than the saturation magnetization Js. In
an efficient p e r m a n e n t magnet material Jr is usually 0.8-1.0Js. If the material is
subjected to a demagnetizing field (i.e. a negative applied field H ) the mag-
netization is gradually reduced and at a critical field - H = jHc the magnetization
is zero. This critical field jHc is known as the magnetization coercivity and is
defined as the reverse field required to reduce the net magnetization of the
material to zero in the presence of the field. The latter qualifying statement is
necessary because if the field is r e m o v e d the specimen may return to a small
positive r e m a n e n t magnetization J; < Jr. Instead of J we can plot the magnetic

/]I//]
/Z
jHc k .,11] /

Fig. 1. Saturation hysteresis loop for magnetic flux density B as a function of H (drawn) and for
magnetization J as a function of H with initial magnetization curve (dashed).
42 H. ZIJLSTRA

flux density B = / x 0 H + J as a function of H (drawn line in fig. 1). We then obtain the
flux hysteresis loop with remanence Br = Jr and with a smaller value of the coercivity
which is here called the flux coercivity ~Hc. Note that by these definitions the
coercivities are positive numbers quantifying negative field strengths.
It should be emphasized that the coercivities jHc and BHc are assumed to
correspond to demagnetization of the saturated material, though we shall see later
that permanent magnet materials are rarely if ever absolutely saturated even in
exceptionally high fields. Unless otherwise stated, it can usually be safely assumed
that the values of Hc quoted in the various scientific journals, books and papers
refer to the "saturation" values as defined above. In the following the prefix J will
be omitted when jH~ is discussed.

2. Suitability criteria for applications

2.1. The energy product

The extent to which a material will be suitable for applications in which electrical
energy plays a part (the first groups in table 1) depends on the amount of
magnetic flux linkage per metre squared and the maximum opposing field that can
be tolerated without loss of polarization.
The product of the flux density B and the associated opposing field H, referred
to as the energy product, is a useful measure of the performance of a particular
magnet, since it is proportional to the potential energy of the field in the air gap.
It is useful only, however, when the magnet is not disturbed by fields from
another source. To determine the energy product it is of course necessary to have
information about the hysteresis loop of the material (fig. 2). A permanent magnet
that is subject only to the influence of its own field will be in a state represented
by a working point in the second (or the fourth) quadrant of the hysteresis loop.
In these quadrants the field is opposed to the flux density, and is referred to as the
demagnetizing field.
It can be shown quite generally that the occurrence of a magnetic field outside
the permanent magnet does in fact relate to a field inside the magnet with B and

Fig. 2. Part of a magnetic hysteresis loop for magnetic flux density B; the shaded area is equal to the
maximum energy product (BH)....
PERMANENT MAGNETS; THEORY 43

/ - / i n opposition. To do this, we have to apply Maxwell's equations to a situation


in which there are no electric currents (apart from the circular currents on an
atomic scale, which are the carriers of the magnetization of the material). The
magnetic field strength H then satisfies

curl H = O,

and for the flux density B we always have

div B = 0.

For a permanent magnet of finite dimensions we may therefore deduce (Brown


1962a)

fn(H. B ) d V = 0, (2.1)

where the integration is performed over the complete space R. If this integral is
written as the sum of the integral over the volume (Rrnagn) of the permanent
magnet and the integral over the rest of the complete space (Rrest), then

~Rmagn (/~r B) dV = --fRrest(H- B) dV.

Assuming that the space Rrest is " e m p t y " , i.e., contains no magnetic substances,
then the flux density there is given by B =/x0H. The right-hand side of the last
equation is then negative, which is possible only if B and H inside the magnet are
of opposite sense or at least include an obtuse angle at least somewhere. This
result is not affected if Rrest contains soft magnetic material in which B and H
always have the same direction.
It can also be shown directly from what we have said above why the product
BH is a good criterion of quality for the applications considered in this section. If
we assume that any field present in soft magnetic material is negligible, we may
write:

fRmagn(lt.B)dV=-tXo f nrestH2dV. (2.2)

The right-hand side of this equation is twice the potential energy of the field
outside the magnet (i.e. in the air gap). This is proportional to H • B.
The exact location of the operating p o i n t - a n d hence the value of the energy
p r o d u c t - d e p e n d s on the relative dimensions of the magnet and the magnetic
circuit in which it is used.
In the limiting cases of an infinitely long needle of a closed circuit ( H = 0) or of
an infinitely extensive plate (B = 0) the energy product is equal to zero; then there
44 H. Z I J L S T R A

is no external field. Between these two extremes a situation exists in which the
energy product has its maximum magnitude. In the case of the needle-shaped
magnet the demagnetizing field is very weak and the working point is close to the
point Br in fig. 2. The value of the flux density at this point is the remanence. If
the magnet is made shorter and thicker, the working point then moves along the
loop in the direction of the point sHe, which it reaches if the magnet is given the
form of a thin plate magnetized perpendicular to its plane. The demagnetizing
field then has its maximum value and exactly compensates the magnetization. In a
properly dimensioned design the energy product will thus assume a maximum
value, ( B H ) .... which is determined solely by the material used. The suitability
criterion sought has thus been found.
The product can be represented by the area of the shaded rectangle in fig. 2; its
magnitude is equal to twice the total potential energy of the field produced
outside the magnet divided by the volume of the magnet. The higher the
remanence, the greater the coercive force and the more convex the hysteresis loop,
the greater is the value of the product. For an ideal magnet, i.e., a magnet that
maintains the saturation value Js of its polarization in spite of the presence of an
opposing field H, the hysteresis loop in the second quadrant is formed by a
straight line going from the point where H = 0, i.e., where B = Br = Js, to the
point where - H = BHc = Js/l~o. The maximum energy product is then given by:

1
(BH)max = 4/x~ j 2 . (2.3)

To reach this maximum it is sufficient if the magnet maintains its saturation until
the opposing field reaches the value -½JJl~o. A further improvement in the energy
product is then only possible with materials that have a higher saturation value J~.
The highest known saturation value at room temperature is shown by an FeCo
alloy (2.4 T); from this value the theoretical energy product could be as much as
1150 kJm -3 (144 MGOe). However, the coercive force of this alloy is very low, which
makes it unsuitable for permanent magnets.
Figure 3 shows the improvements achieved in maximum energy products over
the years, the record values being indicated on a logarithmic scale. It is interesting
to note how closely the curve approximates to an exponential development.
Once the material and thus the hysteresis loop and the (BH)max value are given,
the magnet system has to be designed to make optimum use of the material
parameters. Very schematically this is done as follows:
Consider a permanent magnet system as drawn in fig. 4. The magnet has a
length Im and cross-sectional area Sin. The air gap has a length Ig and cross-
sectional area Sg. The pole pieces are assumed to have infinite permeability
( H = 0 at finite B). The fields H and B are assumed to be uniform in the magnet
body and in the air gap. For simplicity the field spread outside the magnet and the
air gap is taken to be zero, although this is certainly not true in the given
arrangement. We then have from flux continuity
BmSm = - B g S g ,
PERMANENT MAGNETS; THEORY 45

I00C
k Jim 3 //
50C /
/
j11
(BH)mox
20C
, ~10
100

50 j zt ~

20

10
//
/
5 /
/
//

1
I
880 19'00 1920 19 0,960 1 80
Fig. 3. Historical trend of the maximum energy product (BH)mx achieved experimentally since the
year 1880; (1) carbon steel, (2) tungsten steel, (3) cobalt steel, (4) Fe-Ni-AI alloy, (5) 'Ticonal II', (6)
'Ticonal G', (7) 'Ticonal GG', (8) 'Ticonal XX' (laboratory value, Luteijn and de Vos 1956), (9)
SmCos, (10) (Sin, Pr)Co5 (laboratory value, Martin and Benz 1971), (11) Sm2(Co0.85Fe0.11Mn0.04)17
(laboratory value, Ojima et al. 1977). The energy in MGOe is found by dividing the value in kJm -3 by 7.96.

1
Ig

s/[
Fig. 4. Permanent magnet system with pole pieces. S m and Sg are the cross-sectional areas of the
magnet and the air gap respectively, and Im and lg their respective lengths.
46 H. ZIJLSTRA

and, since no currents are present,

Hmlm - G i g = O,

where the positive direction for H and B is taken to the right. From these
equations it is easily found that

H m = Hglg/Im,

and
B m = - tzoHgSg/ Sm

B m / H m = - tZoSglm / Smlg .

The latter expression shows that the reluctance of the system and hence the
working point of the magnet is entirely determined by the dimensional ratios of
the yoke. Allowance for finite permeability of the pole pieces and for flux leakage
can be made by factors o~ (resistance factor) and /3 (leakage factor) so that the
equations for magnetomotoric force and the flux in the air gap are written as

HgLg = otHmLm,

BgSg = - /3BmSm .

For good designs c~ may have values between 0.7 and 0.95 and/3 between 0.1 and
0.8. Detailed discussions of these factors have been published by Edwards (1962)
and by Schiller and Brinkmann (1970).

2.2. T h e m a g n e t i c free energy

In applications involving clamping ability, lifting power or pull of the magnet


(ponderomotive force, the second category in table 1) the working point is also in
the second quadrant of the hysteresis loop. Whereas in the previous group of
applications it was the location of the working point that mattered, the important
thing now is how the working point moves. If, for example, the application is of a
cyclical nature, it is usually necessary for the working point to "stay well on the
loop" during the cyclical motion, so that good reversibility is important. The
amount of mechanical work spent in going anticlockwise round part of the loop
and completely recovered on going back again is used as a criterion for measuring
the performance of a magnet system for applications of this type.
In these applications there is generally a particular configuration of permanent
magnets and magnetizable objects, which are capable of relative movement.
Leaving aside the work required to overcome friction, the mechanical work
required to produce an isothermal change in the configuration is equal to the
increase in its magnetic free energy. Conversely, a decrease in the magnetic free
energy will result in the same amount of mechanical work becoming available.
PERMANENT MAGNETS; THEORY 47

According to the first law of thermodynamics (conservation of energy) a system


in which a reversible process takes place can be described by the equation

TdS + dA = dU.

The term T dS, the product of the absolute temperature T of the system and the
change of its entropy S, is equal to the amount of heat supplied to the system
from the environment. In addition the environment performs on the system an
amount of mechanical work dA, taken as positive. This sign convention for the
mechanical work performed is employed for systems in which magnetic effects
occur. Both amounts of energy are spent on the increment d U of the internal
energy of the system. The free energy F of the system is defined by

F=U-TS.

It follows from these two relations that

dF = dA- S dT.

If the state of the system changes isothermally (i.e. d T = 0), then

dF = dA.

In a system that contains magnetic material the main problem is to find the
correct expression for the mechanical work.
The criterion used for the suitability of a magnetic material for applications of the
type we are now considering is the maximum possible reversible change of its
magnetic free energy. This value is usually calculated per unit volume of the magnet.
The mechanical work d A associated with an infinitesimal change of the
configuration is equal to

dA ½f_ (H. dS - S . d H ) d V,
magn

where the integration is performed over the part Rmagn of the space occupied by
the material. T o derive this expression for the mechanical work, let us imagine a
number of bodies of various magnetizations arranged in a particular configuration.
We assume that the bodies are situated in each other's magnetic field and that
their temperature remains constant. A slight change in the configuration causes a
change in the fields and hence in the polarizations. For each body the increase in
the magnetic free energy consists of a quantity dFp, connected with the build-up
of the polarization in the material, and a quantity of interaction energy dF~,
since in a field H a piece of material with the polarization vector J possesses the
potential energy, - ( J • H). For the body considered we can now write:

d F = dFp+ d E = H . d d - d ( J - H ) .
48 H. ZIJLSTRA

T o find the change in the free energy of the whole system we must perform a
summation over all the bodies. The contributions from the interaction energy
would then be counted twice, but putting a factor 1 in front of them corrects for
this. The total increase in the free energy is therefore

dFsystem = E dVp+½E dFi,

where both summations are made over all the bodies. Using the above expres-
sions for dFp and dF~ and applying the expression d F = d A for isothermal
changes, we obtain the required expression for the mechanical work, calculated
per unit volume of the material.
W e should note here that the energy change H - d J is positive, because the
structure of the material offers a certain resistance to the change of the polariza-
tion. The interaction energy, - ( J • H ) , has a minus sign because it is customary to
take this energy by definition equal to zero for two bodies that are an infinite
distance apart.
If the polarization vector in the expression for the mechanical work is replaced
by the equivalent quantity B -/~0H, then, after integration,

d A = ½fu ( H . dB - B . dH) d V.
magi

T o evaluate this integral it is necessary to bear in mind that during a change in


the configuration the working point moves along the hysteresis loop in the second
quadrant from P to Q (fig. 5). It is then found that the work d A is equal to the
area of the sector O P Q . One configuration (point P) cannot move farther to the
right than point Br, where H is zero, and therefore the magnetic circuit must be
closed. The other configuration (point Q) cannot m o v e farther to the left than the
point BHc, where B and the force exerted are zero. W h e r e possible, cyclical
processes will be carried out in such a way that the working region in the second
quadrant extends to the vertical axis (Br). T h e magnetic circuit there is closed,
which corresponds to a state of lowest energy. In general the material chosen for

B/
P Br

S
Fig. 5. Part of a B hysteresis curve in the second quadrant. The area of the sector OPQ represents the
change in magnetic free energy when the working point moves from P to Q.
PERMANENT MAGNETS; THEORY 49

these applications is one in which the working point can move reversibly from
remanence over the greatest possible extent of the hysteresis loop. For an ideal
magnet, where the complete (linear) hysteresis branch in the second quadrant is
transversed reversibly, the maximum mechanical work made available per unit
volume during a change of configuration is given by:

½B~BH~ = (1/2/Xo)J 2 •

It will be evident t h a t the magnet must be capable of maintaining its saturation


polarization Js until the opposing field reaches the value -JJtxo. This imposes a
stronger requirement on the coercivity than when the magnet is used for static
field generation. The hexaferrite materials with their (for that time) high coer-
civity of the order of 3 x 10SAm -a ( ~ 4 k O e ) made many of these dynamic
applications possible. Today there are many materials with much higher coer-
civities, notably the rare-earth alloys, whose coercivities are of the order of
106 A m -1 (104 Oe).

3. Magnetic anisotropy

3.1. Anisotropy field and coercivity associated with magnetic anisotropy

Consider a single-crystal sphere of a material with uniaxial magnetic anisotropy,


uniformly magnetized to saturation parallel to the easy axis of magnetization. We
assume that changes in the magnetization occur by a uniform or coherent rotation
of the magnetization Ms and that the anisotropy energy density is given by

Wk = K sin 2 q~,

where ~0 is the angle between the easy axis and the magnetization vector. In the
presence of a f i e l d / - / a l o n g the easy axis we assume that the magnetization vector
is rotated through an angle ~p as shown in fig. 6. In this state the total magnetic
energy is

1 2
W = g/x0Ms + K sin 2 q~ +/x0HMs(1 - cos q~).

Note that the first term, which is the magnetostatic energy of the magnetized
sphere, is independent of the angle q~ because the demagnetization factor of the
sphere is isotropic and equal to ½. For a minimum in the energy W corresponding
to a stable position of the magnetization vector we require

dW d2W > 0.
d~ - 0 and d~ 2

Thus q~ = 0 is a stable position of Ms when


50 H. ZIJLSTRA

easy
axis

Ms

Fig. 6. Uniaxial crystal with easy axis for the magnetization.

H > -2K/IxoMs.

However, if the field H<-2K/l~oMs the position q~ = 0 is unstable and the


magnetization reverses discontinuously to ~0 = ~-. The critical field 2K/~oMs is
defined as the anisotropy field HA. It is important to note that the coercivity/arc of
this model of uniform rotation is equal to HA. W e shall see later that in real
materials the coercivity is smaller than HA and depends on the microstructural
features of the magnetic crystals. However, in any case the anisotropy field is the
upper limit for the coercivity, and magnetic anisotropy is therefore a prerequisite
for coercivity to occur. When the easy axis does not coincide with the field
direction the magnetization reverses by a reversible rotation followed by an
irreversible jump. The hysteresis loop is no longer rectangular and the coercivity
depends on the orientation angle of the easy axis. These cases have been
calculated by Stoner and Wohlfarth (1948) and are summarized in fig. 7. The
hysteresis loop of an array of non-interacting identical particles with uniaxial
anisotropy oriented at r a n d o m is given in fig. 8. The coercivity of this array is

Hc = 0.48HA

if coherent rotation is assumed.


Magnetic anisotropy can have various causes. The most important in p e r m a n e n t
magnet materials are: (a) shape anisotropy, associated with the geometrical shape
of a magnetized body; (b) magnetocrystalline anisotropy, associated with the
crystal symmetry of the material. They are discussed in sections 3.2 and 3.3.
A p a r t from these types of anisotropies, there are a few of less importance for
p e r m a n e n t magnets, but nevertheless worth mentioning. At the surface of a
crystal the atoms are in a position of deviating symmetry as compared with the
PERMANENT MAGNETS; THEORY 51

-1 0 1
I I I r I I I I I I I I [ I

0 0,10,90
1

M/M s

T
¢' 0

-1
0,10,90 0
I I I I I I I I I I I I
-1 0 ~ h 1

Fig. 7. Hysteresis loops for uniform rotation of the magnetization in a crystal with one easy axis for
the magnetization, with the orientation angle (in degrees) between easy axis and direction of applied
field as a parameter (Stoner and Wohlfarth 1948).

1.0

f Y
Sf
0.5 M/Ms

-0.5

-1.0
:-1.5 -1.0
f ~ 0..=
J
0 0.5 1.0 1.5

Fig. 8. Hysteresis loop and initial curve for an array of non-interacting identical particles with one easy
axis, oriented at random (Stoner and Wohlfarth 1948).
52 H. ZIJLSTRA

interior. This may give rise to a magnetic anisotropy experienced only by these
superficial atoms, causing the spins to be oriented normal to the surface, or in
other cases, tangential to the surface. In a non-spherical crystal this leads to a net
anisotropy effect which, of course, becomes smaller with increasing size of
the crystal. The underlying theory has been treated by N6el (1954) and the effect
may contribute somewhat to the shape anisotropy of heterogeneous elongated
particle magnets like alnico. A similar kind of "surface anisotropy" has been
observed by Berkowitz et al. (1975) on fine ferrite particles covered with a
monomolecular layer of oleic acid.
Another interfacial type of anisotropy was discussed by Meiklejohn and Bean
(1957), namely, exchange anisotropy. It occurs when an antiferromagnetic crystal
with high anisotropy and a ferromagnetic crystal constitute one solid body such that
the spins on either side of the interface are coupled by exchange forces. The
(non-magnetic) antiferromagnetic crystal then imparts its anisotropy to the ferro-
magnetic crystal. The effect has been observed with small cobalt spheres covered by
an anisotropic antiferromagnetic oxide layer.

3.2. Shape anisotropy

Shape anisotropy refers to the preference that the polarization in a long body has
for the direction of the major axis. The effect, which does not arise from an
intrinsic property of the material, can easily be described in the case of a prolate
ellipsoid. It is assumed that the ellipsoid is homogeneously magnetized in a
direction that makes an angle ~0 with the major axis (fig. 9). The demagnetizing
/
I
/
I

I
I
!
I
I
I

Fig. 9. Magnetized prolate spheroid.


PERMANENT MAGNETS; THEORY 53

field Ha due to the magnetic poles at the surface is also homogeneous within the
ellipsoid. Along each of the three principal axes of the ellipsoid we can apply one
of the relations

I.l"ondi = -- NiJi ,

where i is the number of the principal axis; the coefficients N~ are the demag-
netization factors. In the case of a prolate spheroid we have the "parallel"
demagnetization factor NIl for the direction parallel to the axis of revolution and
two "perpendicular" demagnetization factors Na, which, for reasons of symmetry,
are identical. Using eq. (2.1)

f (H.B) dV=O,

it can easily be shown that the energy Em of the demagnetizing field, which is
given by

Em = II.~OIR n 2 d V ,

depends in the following way on the parameters that describe the situation:

Em = ~ 0 {]~lJ2 + (NL - ]~l)J 2 sin 2 ~p}Rmagn •

In this expression Rmagnis the volume of the ellipsoid. The coefficient of the
directionally dependent part of the equation describes the shape anisotropy. For
small values of the angle ~0 we find from this an effective anisotropy field

HA = (N±- Nfl/~0 •

In the case of a long bar (needle) this field HA approximates to the value J/2l~o. If
the bar consists of iron (Js ~ 2 Wbm -2) it follows from the foregoing that HA
106Am -1 (~104Oe). This is the value of the coercive force when the mag-
netization is rotated uniformly.
Magnetic materials which derive their hardness from shape anisotropy consist of
a fine dispersion of magnetic needles in a matrix of non-magnetic or weakly
magnetic material, like alnico (fig. 10).

3.3. Magnetocrystalline anisotropy

There are three situations that give rise to magnetic anisotropy as an intrinsic
crystal property. The first and most important one is that in which the atoms
possess an electron-orbital moment in addition to an electron-spin moment. In
54 H. ZIJLSTRA

Fig. 10. Microstructure of an alnico magnet. The alloy consists of a precipitate of elongated particles
(approx. composition FeCo, average thickness 30 nm) in a matrix of approx, composition NiA1. Left:
plane of observation parallel to the easy axis. Right: perpendicular to the easy axis. (Electron
micrograph of 'Ticonal XX' magnet steel (De Vos 1966).)

such a situation the spin direction may be coupled to the crystal axes. This arises
through the coupling between spin and orbital moments and the interaction
between the charge distribution over the orbit and the electrostatic field of the
surrounding atoms. There will then be one or more axes or surfaces along which
magnetization requires relatively little work. The crystal will then be pref-
erentially magnetized along such an easy axis or plane.
The second situation is encountered in non-cubic crystal lattices. In these
crystals the magnetostatic interaction between the atomic moments is also aniso-
tropic, which may give rise to easy directions or planes of magnetization.
The third possibility of crystal anisotropy is found in the directional ordering of
atoms as described by N6el (1954). This typically involves solid solutions of atoms of
two kinds, A and B, linked by the atomic bonds A - A , A - B and B-B. In the presence
of a strong external magnetic field the internal energy of these bonds may be to some
extent direction-dependent. Given a sufficient degree of atomic diffuson - as a result
of raising the temperature, for example - a certain ordering can be brought about in
the distribution of the bonds; in this way it is possible to " b a k e " the direction of this
field into the material as the easy axis of magnetization.
In addition to these sources of magnetocrystalline anisotropy mechanical stres-
ses may contribute through the magnetoelastic (magnetostrictive) properties of
PERMANENT MAGNETS; THEORY 55

the crystal. This contribution, however, is considered to be negligible in hard


magnetic materials. Examples of materials with magnetocrystalline anisotropy are
Ba- or Sr-hexaferrite with H A ~ I . 3 X 106Am -1 (17kOe), MnAI with HA ~
3.2 X 106 A m -1 (40 kOe) and SmCo5 with H a ~ 24 x 106 A m a (300 kOe).

4. Fine particles

4.1. Critical radius for single-domain particles

In section 3.1 the coercivity is calculated in a model assuming uniform rotation of


the magnetization. The result Hc = HA however is in disagreement with experi-
ment, as illustrated by a few examples in table 2. Obviously the assumption of
uniform rotation is not realistic and we have to consider other modes of reversal.
This section deals with the conditions under which uniform magnetization is stable
and with the non-uniform states that may occur otherwise.
Consider a uniformly magnetized sphere. It contains a certain amount of
magnetostatic free energy due to its magnetic dipole field. This energy can be
reduced by allowing a non-uniform magnetization. By this the magnetic field
acquires multipole components at the expense of the dipole field with a general
reduction in magnetostatic energy. The field even vanishes altogether when the
magnetization vectors form closed loops inside the sphere. However, non-uniform
magnetization requires work due to the exchange interaction and the magneti c
anisotropy, and the trade-off of these energies with the magnetostatic energy
determines which mode of magnetization the sphere will have. In general the
mode depends on the radius of the sphere, and in some cases a critical radius can
be calculated below which a sphere is uniformly magnetized in its lowest state.
A rigorous determination of the critical radius for a uniformly magnetized
sphere using the micromagnetic equations (Brown 1957, Frei et al. 1957) would be
extremely difficult because the micromagnetic theory contains non-linear
differential equations. However, a simplified approach to the problem has been
made by Kittel (1949) who calculated approximate values of the critical radius Rc
while Brown (1969) calculated exact values for the upper and lower bounds for

TABLE 2
Comparison of anisotropy field HA and coercivityshe of various
permanent magnet materials.

HA jnc

Material (kAm-1) (kOe) (kAm-l) (kOe)

Ticonal 900 370 4.6 100 1.3


Ferroxdure 1200 15 400 5
SmCo5 24000 300 5500 70
MnAI 3400 40 400 5
56 H. ZIJLSTRA

Re. T h e following calculations are based on B r o w n ' s a p p r o a c h with the following


assumptions:
(1) T h e m a g n e t i z a t i o n is continuous with the magnetic polarization IJs[ constant
(2) T h e energy density due to the e x c h a n g e coupling is (see section 6.2.1):

We = A(V~p) 2 .

(3) T h e f e r r o m a g n e t i c material is uniaxial and the magnetocrystalline aniso-


tropy energy density is (apart f r o m a constant term):

W r = K sin 2 ~ ,

w h e r e ~0 is the angle b e t w e e n the easy axis of magnetization and the mag-


netization vector.
According to B r o w n the critical radius Rc0 b e l o w which a sphere is certainly
uniformly m a g n e t i z e d is

/ I.l,o A \ 1/2
Rco = 5.099~-~s2 ) •

If we assume that for iron (Kittel 1949)

A ~ 2 x 10 -11Jm -1 ,

we find that

Rc0(iron) = 13 n m (130 ~ ) .

O b
Fig. 11. Demagnetization modes of a sphere. (a) Magnetization curling. (b) Two-domain state.
PERMANENT MAGNETS; THEORY 57

In order to calculate the critical radius above which the magnetization is non-
uniform we must distinguish between two cases:
(1) The material has a low magnetocrystalline anisotropy energy density. In this
case the energy associated with the uniform magnetization state is compared with
the non-uniform magnetization state known as "curling". The latter is considered
to have the l o w e s t energy because the curling mode (shown in fig. ll(a))
minimizes the sum of the magnetostatic, anisotropy and exchange energies.
(2) The material has a high magnetocrystalline anisotropy energy density. In
this case the energy associated with the uniform-magnetization state is compared
with the two-domain state having a single plane wall as shown in fig. ll(b).
Unfortunately the micromagnetic theory does not generate the magnetization
configuration that has the lowest energy, so that the non-uniform states as
mentioned have been chosen on the assumption that they have the lowest energy.
Simple calculations suggest that this is probably correct.

(1) Low magnetocrystalline anisotropy energy density. According to Brown (1969)


this condition is defined by

K<O.1781J~/~o
or

K < 1.0686Wm, where

Wm = (1/6/x0)J 2 .

The latter expression, Win, is of course the magnetostatic energy of a uniformly


magnetized sphere. If the above condition is satisfied the curling mode has a lower
energy than the uniform magnetization mode provided that the particle radius

R > RCl ,

where

64053 ~---~ m 1 - 5 . 6 1 5 0 ~ 2 K ) -1

If K = 0 then Rcl = 1.2562Rc0. If K = 0.1781JsZ//x0 then Rcl = ~.

(2) High magnetocrystalline anisotropy energy density. According to Brown (1969)


this condition is defined by

K > 0.1781J~/~0,

in which case the two-domain state shown in fig. 11(b) has a lower energy than the
uniformly magnetized state provided that the particle radius
58 H. ZIJLSTRA

R > Rc2,
where

/x0A ~/: /x0K 1/2


Rc2= 56.129 ( J~s ) ( --77-+
Js 1.5708)

Note that when R > Rc2 the two-domain state has a lower energy than the
uniformly magnetized state even when the material has a low magnetocrystalline
anisotropy density.

When K = 0, R c l = 1.2562Rc0 and Rc2 = 13.7965Rc0 ;


when K = 0.1781J~//z0, Rcl = ~ and Rc2 = 14.5576Rc0 ;
when K = 0.1627J~/Iz0, Rca = Rc2 = 14.5Rc0.

Graphs of the ratio R c l / R c o and R c j R c 0 as functions of the parameter


x = txoK/J~ are shown in fig. 12. The calculations of the critical radius made by
Kittel (1949) are also shown for comparison.
Kittel's calculations are based on a comparison of the approximate energy of a
two-domain sphere having a plane wall through the centre with the energy of a

102

two doma ~s

10

curling

1o.2 lo-~---~ x =.uo K/J~ 10

Fig. 12. Ratios of upper bounds beyond which magnetization curling occurs, Rcl (curve a) or the
two-domain state has the lowest energy, Rc2 (curve b), both with respect to the lower bound Rc0,
below which the uniform state is stable, as a function of the reduced magnetocrystalline anisotropy
x = t.LoK/J~. T h e critical radius separating uniform from non-uniform behaviour (Kittel's ap-
proximation) is also given as its ratio with Rc0 (curve c).
PERMANENTMAGNETS;THEORY 59
uniformly magnetized sphere. H e finds the latter to have lowest energy when

#0A 1/2 /z0K 1/2


R<9tx°2Y
Js =36(js)~ (Js)

where the domain wall energy 3' = 4X/A-K. H e has assumed that the mag-
netostatic energy of the two-domain state is approximately half of that of the
uniformly magnetized state. The critical radius as determined by Kittel (1949) can
be compared with Rc0 as determined by Brown as follows:

R (Kittel) _ 3 6 / x 0 X / ~ / 5 . 0 9 9 ~ / / x 0 A
Re0 " / Z
/ p, o K \ 1/2
= 7.06~-2 ) .

The ratio R(Kittel)/Rco is also shown in fig. 12 and plotted as a function of the
parameter x = t~oK/J~. This line fits reasonably well between Brown's upper and
lower bounds but is certainly incorrect for values of x < 0.02. Note that in all the
above calculations of ratios of Rc~, Re2 and R (Kittel) with respect to Rc0 no assump-
tion has been made about the value of A.
From the above discussion it will readily be appreciated that single domain
particles, i.e., particles with radius R < Rc0, are necessarily uniformly magnetized
but that particles with R >Rc~ or R > R c 2 are not necessarily non-uniformly
magnetized. Although the latter particles can be uniformly magnetized, the
energy of that state is higher than that for the non-uniform state. Thus the
uniformly magnetized state may persist if there is an energy barrier between this
and the non-uniform state. This is true for a perfect single crystal in which the
nucleation of a domain wall requires a finite energy for nucleation (see section
4.2). It is also possible for particles with R < R c 0 to contain domain walls
provided they contain lattice defects where the domain wall energy is lower than
that in the surrounding matrix. The coercivity is determined by the height of the
nucleation energy barrier and hence by the presence of lattice defects and the
particular magnetic spin structure of the material (see section 6.5.3). The presence
of superficial features, such as scratches and sharp edges, may also influence the
coercivity owing to the associated local demagnetizing fields, which may assist
domain wall nucleation (see section 6.4). The coercivity can also be determined by
domain wall pinning at the lattice defect (see section 6.5.2). The behaviour of the
sphere for radii between Rc0 and Rcl or Rc2 is unknown, but it cannot be
excluded that the magnetization alternates from the uniform to the non-uniform
states. The region between Rc0 and Rcl or Rc2 is associated with the upper and
lower bounds to the magnetostatic energy of the non-uniform states (Brown
1962b). If this energy is zero as is indeed the case for a cylindrical bar which
demagnetizes by the curling mode, the calculation is exact, and Rc0 and Rc~
coincide and therefore correspond to a single critical rod radius (Frei et al. 1957)
(see also section 5.1).
60 H. ZIJLSTRA

4.2. Brown's paradox

For high anisotropy a supercritical (R > Rc2) sphere has the non-uniform multi-
domain mode as the lowest energy state. However, if the particle happens to be in
a uniform state it cannot spontaneously transform to the lower energy state. For
this a wall has to be nucleated, which means that one or several spins must start
rotating.
Consider one particular spin. It is subjected to an effective field H which is
composed of

H = HA + Hw+ Ha+ He,

where H A is the anisotropy field; Hw is the Weiss field, accounting for the
exchange interaction between the spin and its neighbours; Ha is the demagnetiz-
ing field; He is the externally applied field.
For instability of the spin it is required that

- (Ha + He) > HA + H w .

Now Hw is of the order of 10 9 A m -1 which far outweighs any practical value that
Ha or He could reach. The conclusion is that the uniform magnetization is
maintained under all circumstances and that when

- H e > HA

the magnetization reverses by uniform rotation.


The coercivity of a spherical crystal is thus always

H~=Ha,

which is in obvious contradiction with experiment (see table 2). This inconsistency
which is referred to as "Brown's paradox" (Shtrikman and Treves 1960) is solved
by considering that lattice defects are able to reduce Hw considerably and even
reverse it locally. Also HA can be influenced by a defect as the symmetry of the
crystal is disturbed locally. Finally sharp edges and scratches can locally increase
Ha. These matters are discussed in more detail in sections 6.4 and 6.5.

5. Coercivity associated with shape anisotropy

5.1. Prolate spheroid

From calculations using micromagnetic theory Frei et al. (1957) and Aharoni and
Shtrikman (1958) have shown that magnetization reversal of a prolate spheroid
may occur by three basic mechanisms. These are:
PERMANENT MAGNETS; THEORY 61

(a) Uniform rotation of the magnetization for which the coercivity is equal to
the anisotropy field

Ho = HA = ! (N. - N)J~, (5.1)


/Xo

where N~ and N]I are the demagnetization factors perpendicular and parallel to
the major axis of the spheroid (see also sections 3.1 and 3.2).
(b) Magnetization curling (see figs. l l ( a ) and 13(a)) for which the coercivity is

Hc=k Js 1 (5.2)
2/Zo p2,

where p = R/Ro, R is the minor half axis of the spheroid, R0 is a fundamental


length defined by R0 = (47rtxoA/J2)1/2 and A is the exchange energy coefficient as
discussed in sections 4.1 and 6.2.1. The factor k depends on the axial ratio of the
spheroid and is equal to 1.08 for the infinitely long spheroid or the infinite
cylinder. For the sphere k = 1.39. However, a sphere will rotate its magnetization
uniformly under any applied field. Therefore its coercivity is zero. The sphere can
perform a transition in zero applied field from the uniformly magnetized state to a
non-uniform one by the curling process under its own demagnetizing field. The
condition for this is

Js > 1 . 3 9 Js 1
3tZo 2/Xo p 2, (5.3)

where the left-hand member is the self-demagnetizing field of the sphere and the

£1 b c
Fig. 13. Demagnetization modes of the infinite cylinder: (a) curling; (b) twisting; (c) buckling.
62 H.' ZIJLSTRA

right-hand member follows from eq. (5.2) with k = 1.39. This is rewritten as

p2 > 2.09,
or
/ lzoA \ 1/2
R > 5.121--=~/

which is about the result obtained by Brown (1969) for the critical radius of a
sphere (see section 4.1).
It is interesting to note the similarity between the quantity R0 and the thickness of a
domain wall. As discussed in section 6.2.1, the wall thickness 6 is determined by the
exchange energy competing with the anisotropy energy, so that 3 c~ ~/--A/K. In the
present discussion we deal with a balance between exchange energy and mag-
netostatic self-demagnetization energy, the latter being proportional to J2/iXo. If we
substitute this for K in 6 we obtain

{ l~oA "ll/2
6 oc\ j2 ] o:Ro.

(c) Magnetization buckling. This mechanism of magnetization change is shown


in fig. 13(c) and is degenerate with magnetization twisting (fig. 13(b)) as first
described by Kondorsky (1952). Both of these mechanisms are nearly degenerate
with uniform rotation of the magnetization of an infinite cylinder with R < R0 and
represent a higher energy barrier than curling does for R > R0. This is illustrated
in fig. 14 where the coercivities of an infinite cylinder due to these mechanisms is
shown as a function of R/Ro. The buckling and twisting mechanisms will be
ignored in the present discussion.
When the magnetization changes by uniform rotation the associated anisotropy
energy is entirely of magnetostatic origin, whereas for magnetization curling the
associated energy is entirely due to changes in the ferromagnetic exchange energy.
In the latter case there is no magnetic flux leakage from the surface of the
spheroid so that the magnetostatic energy is zero. This result implies that for the
curling mode the coercivity is independent of the particle packing density.
For the uniform rotation of the magnetization the coercivity depends on the
particle packing density p (p is the ratio of the volume occupied by the particles
compared with the total volume of the specimen) i.e., for a system of parallel
infinite cylinders

1
Hc = ~ Js(1 - p ) . (5.4)
/-/xo

For a derivation of this result see Compaan and Zijlstra (1962). Thus in any
assembly of particles the hysteretic behaviour will be determined by the mag-
netization reversal mechanism which has the lowest coercivity (see fig. 14). In all
the above cases the particles remain uniformly magnetized until the reverse field
PERMANENT MAGNETS; THEORY 63

Uniform r o t a t i o n

Buckling or
twisting

0.2 HC/HA

o, l Curling

0.05

0.02

~- R / R o
0.01 I i i t
0.2 0.5 1 2 5 10 20 50

Fig. 14. R e d u c e d coercivity He/HA due to various demagnetization m o d e s of the infinite cylinder as a
function of reduced radius R/Ro.

nucleates an instability in the magnetization, which is then reversed either by a


sudden uniform rotation or a curling of the magnetization. In this case the
nucleation field is the same as the coercivity and the hysteresis loops are all
symmetrical and rectangular. Which mechanism of magnetization reversal occurs
depends on both R and p. The uniform rotation mode changes to the curling
mode for a system of parallel infinite cylinders when

R e > I ~~13.6 ( ~ 2A ) , (5.5)

which is obtained by putting Hc of eq. (5.4) greater than Hc of eq. (5.2). The
critical radius for an isolated cylinder is

/ i,~oA \ l/2
Rc = 3.68~---f{-2) .

If we assume that for iron A = 2 x 10-'1 Jm -1 (Kittel 1949) and Js = 2 T, the critical
radius for an isolated infinite cylinder is 9 x 10-9m. For an assembly of iron
cylinders with a packing density p = 2 (as it occurs for example in alnico 5)
Rc ~ 16 x 10-9 m. The measured coercivity Hc for alnico 5 is about 5.5 X 104 A m -1.
64 H. ZIJLSTRA

F r o m measurements with a torque m a g n e t o m e t e r the anisotropy field H A -~


2 x 105Am -1. The latter value is the coercivity which would be expected for
uniform rotation of the magnetization. According to measurements m a d e by D e
Jong et al. (1958) the rod diameters in alnico 5 are about 3 x 10 -8 m, which is too
close to the calculated critical diameter to determine whether the difference
between HA and Hc is due to curling or to the fact that the elongated particles are
not regular in shape.
More convincing evidence in support of the above theory has been provided by
Luborsky and Morelock (1964) who measured the coercivities of Fe and FeCo
whiskers of various diameters. The coercivities varied from 4 x 104Am -1 to
25 × 10 4 A m -1 for whisker diameters in the range 65 nm to 5 nm and are in very
good agreement with the theoretical curve for the curling mechanism. For
whiskers with larger diameters the experimental results deviate from the
theoretical curve, due presumably to the presence of a finite magnetocrystalline
anisotropy energy and to the non-circular cross section of the whiskers.

5.2. Chain of spheres

The appearance of electrodeposited particles in a mercury cathode, as observed by


Paine et al. (1955), inspired Jacobs and Bean (1955) to investigate theoretically the
hysteretic properties of a chain of ferromagnetic spheres, consisting of an in-
trinsically isotropic material. The spheres touch each other but have only mag-
netostatic interaction. Two mechanisms of reversal are considered: (a) symmetric
fanning, and (b) parallel rotation.

a b c

Fig. 15. Demagnetization modes of the chain of spheres: (a) symmetric fanning; (b) parallel rotation.
For comparison the uniform rotation mode of the prolate spheroid of the same dimensional ratio is
also indicated (c).
PERMANENT MAGNETS; THEORY 65

The coercivities of these models are compared with those of prolate spheroids
of the same length-to-diameter ratio. The three models are shown in fig. 15. The
symmetric fanning appears not to provide the lowest energy barrier owing to end
effects that have been ignored. Taking these into account leads to a modified
fanning process, called asymmetric fanning. The results of the calculations for a
system of non-interacting elongated particles oriented at random are given for the
various mechanisms mentioned as a function of the length-to-diameter ratio (fig.
16). The experimental points refer to samples consisting of electrodeposited
elongated particles (Paine et al. 1955) with diameters lying between 14 and 18 nm.
Coercivities are in good agreement with the asymmetric fanning model. However,
this might be a fortuitous agreement, since the experimental spheres have
certainly more than a point-like contact, and possibily exchange interaction
between the spheres has to be reckoned with. The mechanism must then be
something between symmetrical fanning and magnetization buckling as described
in section 5.1.

b
0

I I I

~- Elongation

Fig. 16. Coercivity of fine-particle iron oriented at random as a function of particle elongation. Chain
of spheres model: (a) parallel rotation; (b) symmetric fanning; (c) asymmetric fanning. Prolate
spheroid model: (d) uniform rotation. The points refer to experiments (Jacobs and Bean 1955).
66 H. ZIJLSTRA

The magnetization reversal in elongated particles has been thoroughly analyzed


theoretically by Aharoni (1966).

6. Coercivity associated with magnetocrystalline anisotropy

6.1. Magnetization reversal by domain wall processes for tXoHA > Js

The fundamental requirements for magnetization reversal by domain wall pro-


cesses are:
(a) The nucleation of a domain wall (or a reverse domain) by a nucleation field
Ha which may be either positive or negative, though for high coercivities we
require Ha to be large and negative.

Hn ~H

IHnl>lHpl

b
J
Hp Hn ~H

J IHnl<lHpl
Fig. 17. Hysteresis loops in relation to nucleation field H. and pinning field Hp: (a) Coercivity
determined by H.; (b) Coercivitydetermined by Hp.
PERMANENT MAGNETS; THEORY 67

,~ oo-o.--=

B C ':"

P I
-115 -1.10 0.~ H 1'10 L (MA/rn)
1.5

Fig. 18. Hysteresis loop of a SmCo5 particle of 5 i~m size. After magnetizing in a strong positive field a
wall is nucleated at an applied field of about -0.6MAre -1 (point A) or at an internal field of
-0.8 MAm-1 (due to the self-demagnetizing field of -0.2 MAm-1). The internal curve measured after
applying a weaker positive field reveals wall nucleation at zero applied field and subsequent wall
detachment at about -0.35 MAre-1 applied field (point B) which means a pinning field of 0.55 MAm-~
for this particular pinning site. The curve C demonstrates the absence of strong pinning sites in a large
part of the crystal (Zijlstra 1974).

(b) T h e d i s p l a c e m e n t of the wall by a reverse field, if necessary by d e t a c h i n g it


from its p i n n i n g sites, in which case a reverse field Hp k n o w n as the p i n n i n g field
m u s t be applied.
T h e resulting hysteresis loop d e p e n d s o n which of these two processes is
d o m i n a n t . Schematic hysteresis loops for m a t e r i a l s in which the coercivities are
principally d e t e r m i n e d by d o m a i n wall n u c l e a t i o n a n d p i n n i n g are s h o w n in fig.
17(a) a n d (b) respectively.
A hysteresis loop which shows m a g n e t i z a t i o n changes c o n t r o l l e d by either
n u c l e a t i o n or p i n n i n g is s h o w n in fig. 18.

6.2. The 180" domain wall

6.2.1. Energy and width of a 180 ° domain wall


C o n s i d e r a 180 ° d o m a i n wall in a uniaxial f e r r o m a g n e t as d e p i c t e d in fig. 19. T h e
o r i e n t a t i o n angle of the m a g n e t i z a t i o n is 0 at x = -co a n d 7r at x = +oo. T h e x-axis
is n o r m a l to the wall plane. T h e m a g n e t i z a t i o n has its easy axis along the z-axis.
T h e c e n t r e of the wall is at x = 0. Inside the wall the m a g n e t i z a t i o n vectors
deviate from the easy axis by an angle ~, giving rise to an a n i s o t r o p y e n e r g y density

WK = K sin 2 q~, (6.1)

w h e r e ~0 d e p e n d s o n x a n d K is the a n i s o t r o p y constant. T h e m a g n e t i z a t i o n inside


68 H. Z I J L S T R A

easy
axis
I
I

Fig. 19. Perspective view of 180 ° domain wall.

the wall is not uniform, giving rise to an exchange energy density

We=A/d ] 2 (6.2)

where A is the exchange energy coefficient. Here it is assumed that no higher


terms than K occur in the anisotropy energy and that the ferromagnetic coupling
is adequately described by nearest neighbour interaction. The wall energy per unit
area then is

Y = f_7=(W~; + We) dx, (6.3)

with ~0 depending on x in such a way that the energy is a minimum. By using


variational calculus (see e.g. Margenau and Murphy 1956), we find the Euler
equation as the minimizing condition for this problem to be

d2~
K sin 2q~ - 2A ~ = 0,

stating that the torque is zero everywhere. Multiplying by dq~/dx and integrating
from -oo to x we find

K s i n 2~0=
A{d~] 2
\~-x] ' (6.4)

stating that the energy density due to magnetic anisotropy equals that due to
exchange interaction everywhere. The relation between q~ and x follows by
integration to be

tan ½~o= e ~'/~ ,


PERMANENT MAGNETS; THEORY 69

------7------ y~

Y
0
6 Ii-

Fig. 20. Distribution of spin orientation angle q~across a 180° domain wall (schematic). The width 6 of
the wall is defined after Lilley (1950).

where x0 = X/--A/K is called the wall thickness parameter. According to this


relation a wall is infinitely thick. However, the m a j o r part of the orientation
change and hence the energy is concentrated near x = 0. Following Lilley (1950)
we define the wall thickness 6 as the distance between the intersections of the
tangent to ~ at x = 0, with ~p = 0 and ~ = zr (fig. 20). We then find

6 = 7r~/~4/K = 7rXo. (6.5)

The wall energy is calculated by substituting WK for We in eq. (6.3) and shifting
from x to ~0 as a variable using eq. (6.4):

y = 2V'A-K
L sin ~ dq~ = 4X/A--K. (6.6)

Equations (6.5) and (6.6) are derived under the assumption m a d e in eq. (6.2) that
the magnetization is continuous, i.e., the discreteness of the lattice of magnetic
atoms is ignored. This is generally allowed provided that x0 is large c o m p a r e d with
the lattice p a r a m e t e r or the distance between magnetic nearest neighbours.
However, this approximation gives rise to errors when x0 approaches the lattice
p a r a m e t e r a, i.e., when a2K becomes comparable with A. This case of very strong
anisotropy is treated in section 6.5.4.
When a wall is pinned an applied field causes a contribution to the wall energy
which is treated in section 6.5.2.

6.2.2. The exchange energy coefficient A

6.2.2.1. Determination of the exchange energy coefficient A fro m the Weiss internal
field model of ferromagnetism. The Weiss internal field model describes the
ferromagnetic interaction between localized spins in a regular lattice. A particular
spin i is aligned with its nearest neighbours by an effective field per nearest
neighbour j
H,,j = N M / Z ,
70 H. Z I J L S T R A

where N is the Weiss field constant, M the magnetization of the crystal and Z the
n u m b e r of nearest neighbours or coordination number. Consider the spin pair i, L
each spin having a magnetic m o m e n t / z . They differ in orientation by an angle Aq~.
The exchange energy change associated with this orientation difference can be
written as

wij = ~ j ~ (1 - c o s a ~ ) = (NM/Z)tx (1 - cos 2~q~),

which for A~ ~ 1 approaches

N M , ,t ,,2
wi~ = ~ - / x t a q ~ ) .

Suppose an angular gradient d~p/dx is to be present along the x direction. With a


distance ~ between the two spins and an angle h between their connecting line
and the x-axis we then have

NM *2 [d~o'~2
wij = - ~ - ~t c o s 2 ,~ \ ~ x x ] "

This expression will be applied to various crystal lattices:


(a) Body-centred cubic (bcc, fig. 21). For a bcc lattice the n u m b e r of nearest
neighbours is 8, all with cos2A = ½ if the x-axis is coinciding with a cube edge.

13

Fig. 21. Body-centred cubic crystal lattice.

With a being the lattice p a r a m e t e r we have ~2 = 3a2 and

NM a 2 [dq~ ,~2
w,j = ---fg- ~ T Td;x ] "

Per unit cell of volume a 3 there are 8 such pair interactions and two atoms
contributing with 2~//z0 to the magnetization M. Substituting this and dividing by
the cell volume results in the exchange energy density We due to the angular
gradient d~/dx:

N M 2 az{d~ ~2
We= ~ o ~ \~xl "
PERMANENT MAGNETS; THEORY 71

Cl

Fig. 22. Face-centred cubic crystal lattice.

(b) Face-centred cubic (fcc, fig. 22). For an fcc lattice the n u m b e r of nearest
neighbours is 12, of which 8 with cos2A = ½ and 4 with cos2A = 0. The nearest
neighbour distance is ~: = ½a~/2 in all cases.
Per unit cell there are 24 pair interactions and 4 atoms. Weighting the
interactions with their proper values of cos2A we find the exchange energy per
unit volume

NM 2 2/d~\ 2
We =/z0 ~ a I~x-x,] .

(c) Simple cubic (sc, fig. 23). For an sc lattice the n u m b e r of nearest neighbours

9i
i
i

i i

i i PF

O-- • - :g-'- - -O


Q,,
6

Fig. 23. Simple cubic crystal lattice.

is 6, of which 2 with COS2A = 1 and 4 with COS2A = 0. The nearest neighbour


distance ~: = a in all cases. Per unit cell there are six half interactions (the other
halves to be attributed to adjacent cells) and only one atom. The exchange energy
per unit volume is then

NM 2 a2{d~o'~2
We = ~ o - i ~ \dx-x/ "

(d) Hexagonal close-packed (hcp, fig. 24). For an hcp lattice the coordination
n u m b e r is 12. The nearest neighbour distance sc = a in all cases. Per unit cell there
are 23 pairs, which are tabulated below together with the proper value of cos2A
and the part p for which they have to be attributed to the unit cell under
consideration (table 3). There are two atoms in this cell contributing 2/x/~0 to the
72 H. Z I J L S T R A

-~ - -..-- - i- ; . ~
O- . . . . . -O-/- - / f ' -

0 / ~\
¢_~J \
~X

Fig. 24. Hexagonal close-packed crystal lattice.

magnetization. With these data we find for the exchange energy density

N M 2 a2[ d~o,~2
w° = ~°-J3-- \TX-x] "

If we express We in terms of the nearest neighbour distance ~: rather than the


lattice p a r a m e t e r a we obtain for all four structures the same result (N6el 1944a)

NM2.2{d~ \2
we = .

This is not the case when we expand the structures along the vertical axis in order
to obtain tetragonal symmetries from the cubic structures or to deviate from the
close packing in the hexagonal case. In all these cases, however, the expressions

TABLE 3
Interactions in an hcp cell.

Location in cell Number cos2A p Contribution

F r o m centre plane 4 ¼ 1 1
to upper and lower
planes 2 0 1 0
1
In centre plane 3 1 ~ 1
2 ¼ 1 g
2 ¼ _2 1
3 3

1
In upper and 2 1 g 1
1
lower planes g ¼ ½

Total contribution 4
PERMANENT MAGNETS; THEORY 73

for We in terms of a remain unchanged, because the expansion is perpendicular to


the x-axis. If we introduce the Weiss-field energy

Ew = ½1~oNM2

we have for the coefficient A of eq. (6.2):

bcc: A = ~a2Ew = l~2Ew,

fcc: A = ~2a2Ew= ~2.Ew ,


sc: A = la2Ew = ~2Ew,

hcp: A = {a2Ew = {sC2Ew.

6.2.2.2. Determination of the exchange energy coefficient A from the Heisenberg


model of ferromagnetism. According to the Heisenberg model the ferromagnetic
coupling between two spins is expressed as

w~j = -2J$~ • ~ ,

where St and Sj are spin vectors of two neighbouring atoms and J is a coefficient
called the exchange integral. For J > 0 parallel spins have minimum energy,
resulting in ferromagnetic coupling. We suppose an angular gradient d~/dx to be
present along the x-axis. The coupling energy between spins i and j can then be
written as
(d~'~ 2
Wi] = J S 2 ~ 2 COS2 a \ ~ X ] '

where sc is the distance between the spins, A is the angle between their connecting
line and the x-axis, and the moduli of the spin vectors are assumed equal. In the
same way as in the previous section these energy contributions are added for the
various spin pairs in a crystallographic unit cell. We then find for the exchange
energy density,
2 J S 2 {d~o'~ 2
bcc: We = - - 7 - ',~X ] '

4JS z {d~o"~2
fcc: We = - - 7 - - \d--x-x) '

JS 2 {dq~'~2
SC: we = --a- \ ~ x x j ,

JS22~/ 2 [ d~o'~2
hcp: W e - a k~xx] "

Expansion by a factor a of the vertical axis divides the coefficient of (dq~/dx) 2 by


74 H. ZIJLSTRA

the same factor a, as only the cell volume increases by this factor and the rest
remains the same.
The coefficients thus derived are usually written as A, e.g., by N6el and Kittel,
or as C = 2A by Brown. They are associated with the Curie t e m p e r a t u r e Tc and can
be determined by calorimetry, spin-wave resonance m e a s u r e m e n t s or t e m p e r a t u r e
dependence of magnetization. A difficulty is that the models discussed here are based
on nearest neighbour interaction, although there is much experimental evidence that
interactions at a longer range cannot be ignored. Therefore determination of A will
seldom be better than an indication of the order of magnitude. Using the
approximate relation

A~105~ZTc, (SI),

with A in J/m, ~ the nearest neighbour distance in meters and Tc the Curie
temperature in Kelvin, fulfills most requirements in the present context. In cgs
units we have A in erg/cm, ~ in cm and Tc in Kelvin, for which the relation
becomes

A ~ 106~2Tc, (cgs).

6.3. Interaction of domain walls with cavities and non-ferromagnetic inclusions

Consider an array of non-ferromagnetic spheres on a simple cubic lattice as shown


in fig. 25, and assume that the spheres have a radius p and occupy a fraction a of

radius p

0
0 © ©
0 )
d

0 )
t
Domain Wall

Fig. 25. Cubic array of spherical cavities interacting with a rigid domain wall.
PERMANENT MAGNETS; THEORY 75

the total volume of the material. The n u m b e r n of spheres per unit volume is

3
n = 0, 4 . / r p 3 ,

so that the n u m b e r of spheres which are intersected by a (100) plane is given by

/./2/3= { 3 ~ 2/3 0,2/3


\4~r/ p2 "

The distance d between the centres of the spheres on any (100) plane is given by

/4,B-'~ 1/3

If we assume that a (100) plane of these spheres is intersected by a domain wall


and that the wall m a x i m u m pinning force per sphere is fm, the m a x i m u m pinning
force on the wall is

F = r~
.~2/3g
Jm •

If the wall is m o v e d through an infinitesimally small distance dx in the presence


of an external field H, the change in magnetostatic energy is 2J~H dx. The total
change in energy is

d W = n2/3fm dx + 2J~H d x .

T h e wall will actually m o v e if d W / d x <~O, i.e., if H ~ < - H e where

( 3 )2/3fm0,2/3
Hc = \~-~] ~. (6.7)

The above model has been used by Kersten (1943) as the basis of a simple
theory of coercivity. Kersten assumed radii of the spheres to be much larger than
the domain wall width, i.e., p ~> 6, so that the wall can be regarded as a plane of
zero thickness. When a sphere is intersected by a planar domain wall, which is
also assumed to be rigid, the pinning force is due to the change in the wall
energy which occurs because an area ~rp2 is r e m o v e d when the wall intersects a
sphere through a diameter. However, it should be appreciated that the pinning
force is a m a x i m u m at the edge of the sphere where the rate of change of the wall
energy with position d y / d x is a maximum. For a sphere of radius p the area of
intersection with a plane domain wall changes at a m a x i m u m rate of 2~p so that
the m a x i m u m pinning force fm per sphere is

f~ = 2~-py.
76 H. Z I J L S T R A

Hence the coercivity for a simple cubic lattice of these spheres is given by

3 )2/3 ~-y
He=
_

4~r
_ _ _

pJs 'x
^ 2/3

" (6.8)

Unfortunately the above result does not agree with the experimental values for
ferromagnetic materials which contain dispersions of non-ferromagnetic particles,
principally because the effect of the magnetostatic energy due to the surface poles
has been omitted. When the spheres, each of volume V, are not intersected by a
domain wall the magnetostatic energy is

1 2
m = g/toMs V,

but when they are intersected across their major diameters the above magnetostatic
energy is reduced by a factor of about 2 (N6el 1944b). This magnetostatic energy
variation may not be negligible in comparison with the change in the domain wall
energy. Furthermore the assumption that the domain walls are rigid is unrealistic and
makes the model strongly dependent on the shape of the inclusions.

6.3.1. D o m a i n - w a l l p i n n i n g at large inclusions


N6el (1944b) has extended Kersten's theory and has developed a theory of the
coercivity of an array of identical non-ferromagnetic spheres which includes the
effects of their magnetostatic energies. The resulting expressions for the coercivities
depend on the size of the inclusions compared with the width 6 of the domain wall.
We consider large inclusions first (p >>6).
Consider a non-ferromagnetic sphere in a uniformly magnetized material. The
associated magnetostatic energy of the sphere due to the magnetic charges on its
surface is

W m = l g/x0Ms47rp/3.
2 3

When the sphere is intersected by a plane domain wall through its centre the
above magnetostatic energy is reduced by a factor of about 2, i.e.,

AW m 9171"[.ZoIVl
. # 2s P 3•

A rigid wall moving through the crystal will have minimum energy when it
intersects the spherical hole just through the centre. The force required to move it
away from the centre is

f = d(Wm+ W~)
dx

where W~ is the contribution of the wall energy to the total energy. N6el (1944b)
has numerically calculated the energy Wm as a function of wall position x and the
PERMANENT MAGNETS; T H E O R Y 77

maximum value of its derivative with respect to x,

d Wm') = 0.600 j2p2,


dx ,/max /~0

which maximum is attained when the wall is just tangent to the sphere• With eqs.
(6.7) and (6.8) we then find for the rigid wall pinned by a cubic array of spheres

He = 0 385o:2/3(0 3Ms+ "rr'y "~ (6.9)


• \ " /xopMs]"

Note that the magnetostatic term is independent of sphere radius and that the
wall energy term is inversely proportional to p. There is a critical radius pc, below
which the surface energy term is dominant, and thus Kerstens theory becomes
applicable.
Above pe the magnetostatic term is dominant and N6eI's theory must be
applied. The critical radius is

pe ~ ~o~/Y~ •

This value is exactly the same as that derived by Kittel (1949) for the radius below
which a ferromagnetic sphere is uniformly magnetized in its lowest state (see
section 4.1).
In order to test the validity of eq. (6.9) we substitute the parameter values of a
typical rare-earth magnet and of iron (see table 4). We then obtain for the
rare-earth magnet, with a = 0 . 1 and p = 1 0 - 6 m , H e ~ 1 0 4 A m -1, which is by
several orders of magnitude too low as compared with experiment. The coercivity
of these magnets is obviously not determined by wall pinning at large inclusions•
More likely models are discussed in sections 6.3.3 and 6.5. For iron with the same
dispersion of inclusions we find He ~ 105 Am -1, which is far too high. A possible
explanation for the latter discrepancy is discussed in the next section.

TABLE 4
Intrinsic properties (order of magnitude) of a typical hard magnetic material (lanthanide-cobalt alloy)
and a soft magnetic material (iron).

La--Co Fe (SI) La-Co Fe (cgs)

Anisotropy field HA 107 5 × 104 (Am -1) 10s 5 x 102 (Oe)


Anisotropy constant K 5 × 106 5 × 104 (Jm -3) 5 × 107 5 x 105 (erg cm -3)
Saturation magnetization Js 1 2 (T) 103 2 × 103 (erg Oe -1 cm -3)
Saturation magnetization Ms 106 2 × 106 (Am -1)
Exchange parameter A 10-11 10-11 (Jm -1) 10-6 10-6 (erg cm -1)
Wall energy 3' 5 x 1 0 -2 5 x 1 0 -3 (Jm -2) 50 5 (erg cm -2)
Wall thickness 6 5 x 10 -9 5 x 10-8 (m) 5 × 10-7 5 × 10 -6 (cm)
78 H. ZIJLSTRA

Fig. 26. Spherical cavity in a ferromagnetic crystal with reverse-domain spikes. The arrows indicate
the domain magnetization. Concentrations of surface charges are indicated by their respective signs.

6.3.2. Nucleation of reverse domains at large inclusions


If the spherical cavity is larger than the critical radius it becomes energetically
favourable to provide it with a pair of reverse domain spikes as shown in fig. 26.
The (dominant) magnetostatic energy is then appreciably reduced at the expense
of some wall energy. The latter energy becomes less important the larger the
sphere. Therefore in the magnetized crystal reverse domains may occur spon-
taneously at non-ferromagnetic inclusions or cavities. N6el (1944b) has shown that
these reverse domains expand indefinitely when the applied field H = -H~, where

Hc = 1.23y/tzopMs. (6.10)

For the same two examples of the previous section (table 4) we calculate the
nucleation coercivity at an inclusion of radius p = 10-6m and find Hc(La-Co)
5 × 104Am -1 and H c ( F e ) ~ 103 A m -1. The nucleation at an inclusion in a hard
magnetic material is thus relatively easyl Inclusions of cavities should therefore be
avoided or indefinite expansion of a nucleated domain should be prevented by
some pinning mechanism. The spike formation at a cavity is analogous to the
formation of reverse domain spikes at the flat end face of a long magnetized
crystal, where as mentioned in section 6.4 the local demagnetizing field is
Hd = -½Ms (see also fig. 29).

6.3.3. Domain-wall pinning at small inclusions


If the inclusions are small (p ~ ~), the pinning force is due to the change in the
energy of a wall which occurs when part of its volume is occupied by non-
ferromagnetic inclusions. Consider a spherical cavity of volume V and radius p in
the magnetized crystal. If this is located inside a wall the wall energy is reduced by
the following quantities:

(a) exchange energy We = a ( dq~'~2


\ d x ] V;
(b) magnetocrystalline anisotropy energy WK = K V s i n 2 ~ .

The inhomogeneous magnetization requires a correction Wn of the order p2/t~2,


which may be neglected here since p ~ 6.
PERMANENT MAGNETS; THEORY 79

The presence of the sphere adds a certain amount of magnetostatic energy,


calculated by N6el (1944b) to be

(c) Wa = _~/x0Ms2[1-25P
2_ 2//dq~ 2] g
\ dx ] J '

where the second term is due to the inhomogeneous magnetization with angular
gradient dq~/dx. Outside the wall the magnetization is uniform and oriented along
the easy axis. Hence the energies WK and We are zero and

1 2
Wd = gtx0Ms V.

The difference in energy between the two situations: sphere outside wall and
sphere inside wall then is

[
A W = - / K sin 2 ~0 + A -~
dq~ 2

or using eqs. (6.3) and (6.5),

A W = - [ 2K + 75 B21 V sin 2 q~.

Under the present approximations, K > / x 0 M ~ and p ~ B, we may ignore the


second term and find

A W = -2KV sin 2 q~.

The pinning force (using eq. (6.3) for the relation between ~0 and x) is

f- d(zX
dxW ) _ 4 K V 3 / K~ sin 2 p cos p .

This is maximum for cos 2 ~ = ½so that with eq. (6.5)

sV5 K V
Im-- ~ - - 'TT~--

Substituting this result into eq. (6.7) we find the coercivity due to a cubic array of
spheres with radius p ~ 8 as

2 K a2/3 V 1
Hc = 0.47-----0-~s p 28
(6.11)
= 1.95/-/1 ~ ol 2/s ,

where H a = 2K/~oMs, the anisotropy field.


80 H. ZIJLSTRA

The numerical factor of 1.95 depends on the geometry of the inclusions and
their distribution, but is expected to be of the same order of magnitude for a
variety of probable inclusion shapes and distributions.
Using the parameters of the two examples mentioned in section 6.3.1 (table 4),
we find for a dispersion of inclusions with a radius of 10 -9 m, occupying 0.1 of the
material Volume H c ( L a - C o ) ~ 106Am -1 and H c ( F e ) ~ 5 0 0 A m -1. Both orders of
magnitude agree well with experiment, which suggests that pinning at small
inclusions might be an explanation for the observed coercivities. For this pinning
mechanism it is assumed that the wall is rigid and moves through a cubic array of
spheres (see fig. 25). A non-rigid wall in a random distribution of spheres is able
to arrange itself so that it contains a maximum number of inclusions. The latter, a
more likely picture of a pinned wall is expected to obey eq. (6.11) as well.

6.4. Domain-wall nucleation at surface defects

Domain-wall nucleation may occur at surface defects such as pits, protrusions,


scratches or sharp edges, where the magnetization reversal is assisted by the
locally increased demagnetizing fields (Shtrikman and Treves 1960).
Consider, for example, a surface defect such as a pit (fig. 27) in the form of a
truncated cone with an apex semi-angle ~b, base radius r± and face radius r2. It can
be shown (Zijlstra 1967) that the axial demagnetizing field Ha at the apex point P
is
Ha = 1Ms sin 2 q5 cos ~b ln(rl/r2).

,,//I////////////l
z

Fig. 27. Surface defect.


PERMANENT MAGNETS; THEORY 81

At an infinitely sharp point, i.e., when r2 = 0, the demagnetizing field at the point
P is infinite. However, as pointed out by Aharoni (1962), this is physically
unrealistic since a point cannot be sharper than the atomic radius --~10-1° m. For a
conical pit with a base radius of 1 p~m and sin 2 ~p cos ~b = 0.4 (i.e. ~b ~ 60 °)

Ha ~ 1.8Ms.

Similar demagnetizing field concentrations arise at sharp corners and edges and at
the bottom of cracks and scratches. Although they are appreciably stronger than
the overall demagnetizing field of a magnetized body and perfectly capable of
explaining the persistence of reverse domains in soft magnetic materials (De Blois
and Bean 1959) they are not sufficient to account for nucleation of reverse
domains in hard magnetic crystals. However, there is experimental evidence that
sharp edges do play a role in nucleating reverse domains, as demonstrated by the
following experiments on SmCo5 and related compounds. Becker (1969) has
measured a coercivity of 8.3kAm -1 (105 Oe) on a YCo5 powder made by
mechanical grinding. Subsequent treatment in a chemical polishing solution
increased the coercivity to 266 kAm -1 (3340 Oe). Becker attributed the increase to
the rounding off of the initially sharp edges of the powder particles, which he
observed by microscope. Ermolenko et al. (1973) prepared a single-crystal sphere
of 8mCo5.3 of about 2 mm diameter. After chemical polishing the sphere had a
coercivity of 460 kAm -1 (5800 Oe). Scratching the sphere reduced the coercivity to
practically zero (Shur 1973) and subsequent polishing restored it again. The
influence on the nature of the hysteresis loop was shown by Zijlstra (1974) who
compared the hysteresis loops of two single particles taken from a ground SmCo5
powder before and after annealing (fig. 28). The coercivity of the particle as
ground appeared to be determined by easy nucleation and subsequent pinning of
domain walls. For the annealed particle the nucleation proved much less easy.
The annealing process may have removed internal defects which also have their
influence on the hysteresis. However, McCurrie and Willmore (1979) have ~hown
that a similar behaviour is obtained when the particles are smoothed by chemical
polishing rather than by heat treatment.
A special case is the long body with a flat end face. The local demagnetization
factor equals ½ at this end face, although the average demagnetization factor
approaches zero for longitudinal magnetization. The associated superficial demag-
netizing field may give rise to homogeneous nucleation of reverse domains as shown
in fig. 29.

6.5. Interaction of domain walls with the crystal lattice

6.5.1. Wall pinning at regions with deviating K and A


The nucleation of domain walls at regions with reduced K has been treated by
Aharoni (1960, 1962) and Brown (1963) using micromagnetic theory, but although
a nucleation field Hn of the order of one tenth of the anisotropy field HA could be
derived, their model was not able to explain the many orders of magnitude
82 H. ZIJLSTRA

r~ e..

"" O
E

,- ©

o~
~E

r.~
=2
o

r~

O
PERMANENT MAGNETS; THEORY 83

/
a

Fig. 29. End surface of a magnetized body. (a) The body in cross section. Reverse-domain spikes
penetrate from the surface into the body (schematic). The arrows indicate the domain magnetization.
(b) Micrograph of a Sm2Co17single-crystal surface with the spikes seen from above.

reduction of Hn with respect to HA f o u n d experimentally. Calculations of the


pinning of d o m a i n walls at regions with r e d u c e d K and A were m o r e successfully
carried out by A h a r o n i (1960), Mitzek and S e m y a n n i k o v (1969), Hilzinger (1977)
and Craik and Hill (1974). W e will discuss the p r o b l e m on the basis of the t h e o r y
by Friedberg and Paul (1975) of d o m a i n wall pinning at a planar defect region.
Consider the crystal shown in fig. 30 in which there are three distinct regions a, b
and c defined by

a f r o m - ~ to Xl ,
b f r o m xl to x2,
c from x2 to + ~ .

Their magnetic properties J, K and A are identified by subscripts i = a, b and c


where

Ja=Jc¢Jb,
A a = A~ ¢ A b ,

g . = K ~ gu.
84 H. ZIJLSTRA

I x

I ,/

c ,¸

, b , \
/ I I X2
a I
I Xl

Fig. 30. Domain wall distributed in three zones a, b and c of the ferromagnetic crystal.

The easy axis of the uniaxial crystal is along the z-axis; the planar defect is in the yz
plane. A field H is applied along the z-axis. A 180 ° domain wall parallel to the planar
defect has an energy per unit area of

f~[Ai\{d~2
Y = -~ L d x ] + Ki sin 2 ~ - H J / c o s ~ ] dx, (6.12)

where ~ is the angle between the magnetization vector and the z direction, and the
subscript i applies to the appropriate region where the wall is located. Minimizing Y
by variational calculus and integrating Euler's equations in the three regions yields
the following three equations:

- A i ( d ~ ' ~ 2 + Ki sin 2 ~ - HJ~ cos ~p = C~, (6.13)


\dx]
for i = a, b and c, where C~ are constants to be determined by the boundary
conditions. Imposing the conditions ~ ( - ~ ) = 0 and ~p(+~)--~- and noting that
d~/dx = 0 at x = _+0%determines Ca = -HJa and Cc = HJa. The'continuum approach
inherent in micromagnetic methods requires continuity of ~ at the interfaces at xl
and x2, where ~ has the values ~1 and q2 respectively. Stability of the wall requires
zero torque everywhere and thus minimum local energy density. This requirement
implies continuity of A d~/dx at the interfaces xl and x2, which can be seen from the
following argument.
Consider the interface of xl and a narrow zone of width Ax on either side (fig. 31).
The value of z~x is so small that dq/dx may be taken as constant in each zone. The
energy content of this slab is then (to a first approximation)

A T = A a (~Pl - ~0a)2 ~. ( K a sin 2 f~l -- HJa c o s ~/)1) A x


Ax

+ Ab (~Pb-- ~)2 4- (Kb sin 2 ~p~-- HJb cos ~1) Ax.


Ax
PERMANENT MAGNETS; THEORY 85

i
i

[ i
I I
i 1
I I
I I

Fig. 31. Orientation angle ~0as a function of distance x near the interface between zone a and zone b.

Minimizing this with respect to q~l with fixed values of ~0a,~b and Ax and subsequently
letting Ax approach to zero, yields

A.(~x) = Ab(~-~-~X)b at x= Xl.

The difference ~ b - ~Oahas a fixed value for fixed Ax, since it determines the local
exchange energy density. This must be equal to the anisotropy energy density, which
is fixed by ~1 lying between ~0, and q~b,which interval can be taken arbitrarily small.
Continuity of q~ and A d~p/dx at the interfaces xl and x2 produces four equations
from which dq~/dx and Cb are eliminated to give a relation between q~ and ~02 with
coefficients expressed in the parameters A, K, J and H :

H(AaJ~ - A J b ) ]2 [cos H ( A a J , - AbJb) ]2


[cos ~, + ~ - A--/~uJ - ~,2 ~ 2(--X~Ka- A---~3]

2HAaJa - O. (6.14)
AaKa - AbKb

This equation describes a hyperbola shown in fig. 32. Only the upper right hand
branch applies to our model. For H = 0 this curve degenerates to its asymptote
cos ~j = cos ¢2. The width of the defect determines which point of the curve
represents the actual situation. Narrower defects shift the point to the right. In the
small-deviations approximation, A a ~ . A b , K a ~ K b and Ja-~Jb, eq. (6.14) can be
written as

7)/t-x-+-k-)j -
[cos K,12I
= -4h
/(A_k- ,,÷)
+ , (6.15)
86 H. ZIJLSTRA

cos %

J
cos "P2
-p

\
\\

Fig. 32. The hyperbola (cos ¢1 + p)2 _ (cos ~2 + p)2_ Q = 0.

w h e r e A A = A b - - Aa, A K = Kb = K a , A J = J b - - J a and h = H/HA with H A = 2KJJa,


the anisotropy field of the undisturbed crystal. T h e relation b e t w e e n go1, go2 and the
width x 2 - xl of the defect is calculated by integrating eq. (6.13) for i = b:

f x: dx = f~i2 dgo [A~ sin z go - ~ Cb 1-1/2 '


cos go - AbJ (6.16)
dX 1

with

-- AaKa sin2 gol -- ~H (AbJb -- AaJa) cos ~1 - H AAabJ


Cb - AbKb Ab

H o w e v e r , this integral cannot be solved analytically and has to be approximated.


First consider the case H = 0. F r o m eq. (6.14) we see that cos: go1= cos 2 go2,which
means that for finite width of the defect go1 = ~ ' - ~ 2 and the wall is located
symmetrically with respect to the defect (see fig. 33). Since the width of the defect is
not specified this m e a n s that in zero field a wall finds an equilibrium position at a
defect of any size different f r o m zero; there is no critical size for defects of this kind.
M o r e o v e r this m e a n s that a field, h o w e v e r small, is n e e d e d to detach the wall f r o m
the defect.
N o w assume that the width D of the defect is small c o m p a r e d to the wall width 6
defined by eq. (6.5). T h e angular gradient dgo/dx m a y then be assumed constant
inside the defect. F r o m eq. (6.4) in section 6.2.1 we see that

d_~_~= 4 K b
dx ~ sin ~ ,
PERMANENT MAGNETS; THEORY 87

"it-- -- -- -- [ --

0 . . . . . I
XI X2 ~ X

Fig. 33. Orientation angle ~0 as a function of x of a pinned d o m a i n wall in zero field (drawn line) and
in a small positive applied field (dotted line).

w h e r e we take for ~0 the average value 1(~1 + ~pz). Integrating this gives

(Abl Kb) 112


D = x2 - Xl = sin l ( ~ o I -it- ~D2) (~2}2 - - ~ 0 1 ) "
(6.17)

N o w suppose that a small field H ~ HA is applied. F r o m eq. (6.15) we see that in


the small-field, small-deviations a p p r o x i m a t i o n

COS2 ~1 -- COS2 ~2 = -4h


/?A ~ +T "

T h e small-defect-width a p p r o x i m a t i o n w h e r e ~92~'~-~1 allows

cos 2 q~l - cos 2 ~2 ~ (~02- ~pl) sin(~j + ~2),

so that by substituting this into eq. (6.17) we find

7r D / A A AK\
h- ~- 6b t--A---+--K--) sin(qh + ¢2)sinl(qh+ ~2). (6.18)

L o o k i n g at fig. 33 we see that u n d e r increasing field h the wall will shift to the
right thus steadily decreasing the average angle of orientation ~p inside the defect.
Starting from ~p = ~-/2, the position at H = 0, we see that the angular function of
eq. (6.18) starts at zero and increases to a m a x i m u m of the o r d e r o n e at a certain
critical value of h. F u r t h e r increase of h will give no solutions for ~1 and ~2 so that
no stable wall will exist. W e identify this critical field with the unpinning field or
coercive force

(6.19)
88 H. ZIJLSTRA

The minus sign in eq. (6.18) implies that pinning occurs only if the form between
brackets in eq. (6.19) is negative, i.e., the wall energy inside the defect is lower
than if the defect were not present. If the form in brackets is positive a wall will be
repelled, and the defect will form a barrier rather than a trap. It should be noted
that eq. (6.19) is valid only in the small-field, small-deviations, small-defect-width ap-
proximation. If deviations become larger the symmetry in AA and AK will be lost. In
particular a substantial lowering of A will contract most of the wall inside the defect
so that the condition 6 >> D is no longer satisfied. This particular situation is discussed
in secion 6.5.2. In the small-deviations approximation ~b may be replaed by 6a, the
wall thickness in the undisturbed crystal. Note that hc falls within the small-field
approximation as a direct consequence of the small-defect-width, small-deviations
approximation. In this approximation a deviation of J has no influence on he. In the
case where D >> ~b the wall will be almost entirely within the defect region at H = 0.
With field increasing from zero we deal with a wall penetrating from region b into
region c, i.e., a wall pinned by a phase boundary. Using eq. (6.13) with i = b, we
impose boundary conditions ~ = 0 and d~/dx = 0 for the far left-hand end of the
wall, which yields Cb = -HJb. Similarly we find with q~(+~) = 7r and d p / d x ( + ~ ) = 0
in eq. (6.13) with i = c the integration constant Cc = HJa, recalling that region c has
the s a m e properties as region a. Eliminating d~o/dx from these equations we find

(KaAa - KbAb) sin 2 q?2


H=
(AaJa - AbJb) COS q~2+ (AaJa -{-AbJb) '

which in the small-deviations approximation becomes

(AA/A + AK/K) sin 2 ~2


H = ½Ha(b) 2 + (AA/A + AJ/J)(1 + cos ¢2)"

Looking at fig. 33 we expect that the highest rate of energy change will be found at
about ¢2 = ~-/2, so that the coercivity becomes

_ He 1/AA AK\
(6.20)
hc HA(b)~ ~-A--+--K--)'

which is the pinning force exerted upon a wall that penetrates from a phase with
low A and K into a phase with high A and K. In a material as SmCo5 with
HA = 2.5 x 1 0 7 A m -1 (300 kOe) this pinning mechanism could account for a coer-
civity of the order of 1 0 6 A m i (~104 Oe) with a 10% variation in A and K.

6.5.2. Pinning of a domain wall by an antiphase boundary

This section treats a theoretical model of the interaction of a plane domain wall
with a certain type of plane lattice defect, namely the antiphase boundary (APB).
The A P B occurs in ordered crystals where the atomic order on either side of the
PERMANENT MAGNETS; THEORY 89

X X X I0 0 0
I~ I~ I~III'IX'XI'X ~ X
I
x x × f×?×?×?×

~x~x~×~?x?x?×? x
I
APB
Fig. 34. Ferromagnetic ordered crystal with magnetically active antiphase boundary (APB).

A P B is opposite in phase. This is clarified in fig. 34 for a two-dimensional binary


crystal consisting of A atoms (circles) and B atoms (crosses). The crystal lattice is
continuous, but on the right-hand side of the defect B atoms occupy what would
have been A sites without the presence of the APB, and vice versa. Suppose now
that only the A atoms carry a magnetic m o m e n t and that these are coupled
ferromagnetically in the undisturbed crystal. However, across the A P B the much
shorter A - A distance might give rise to antiferromagnetic coupling, thus dividing
the crystal into two ferromagnetic parts which, in the lowest state and zero field,
are antiparallel as indicated in fig. 34. It has been suggested by Zijlstra (1966) that
such magnetically active A P B s are responsible for the easy nucleation of reverse
domains in MnA1 crystals. On the other hand it was expected that moving domain
walls would encounter strong pinning, which was indeed found to exist in MnA1
crystals (Zijlstra and Haanstra 1966).
Consider a magnetic domain wall as described in section 6.2.1 with its plane
parallel to an A P B as described above. T h e orientation angle of the magnetization is
0 at x = - ~ and ~r at x = + ~ . The A P B is located at x = 0 and coincides with the y z
plane. The ferromagnet has uniaxial anisotropy with the z-axis as easy axis. The
energy densities due to anisotropy and to exchange interaction are described by eqs.
(6.1) and (6.2), respectively. T h e coefficients K and A are assumed to be the same
throughout the crystal, except for the APB.
The situation at the A P B is described as two layers of atoms, at a distance ~,
one belonging to the left-hand side of the crystal and the other to the right-hand
side. The coupling energy density between these two adjacent lattice planes, is
different from the coupling energy density in the undisturbed crystal owing to
shorter A - A distance and is given by

At
w~ = g
~ - [1 - c o s ( ~ , 2 - ~ , ) ] , (6.21)

where pl is the orientation of the left-hand layer and ~2 that of the right-hand
layer. The coefficient A ' is different from A and can be negative, in which case
antiparallel coupling across the A P B is favoured. The structure of the wall in an
arbitrary position with respect to the A P B is shown in fig. 35 by the orientation
90 H. Z I J L S T R A

0 -×

Fig. 35. Orientation angle p as a function of distance x normal to a domain wall pinned at APB.

angle q~ as a function of x. T h e wall energy per unit area can then be written as

3' = 2X/)-K(1 - cos q~,) + 2 X / ~ ( 1 + cos q~2)

+ (A'/sc)[1 - cos(~2 - ~01)] nt- ½K~(sin 2 ~, + sin 2 (P2). (6.22)

T h e first two terms follow by integrating eq. (6.6) from 0 to ~1 and f r o m q~2 to ~,
respectively. T h e third term is the exchange coupling energy in the slab of
thickness ~ at the A P B and the last term accounts for the anisotropy energy in the
same slab. N o w approximating to continuous magnetization with the A P B as a
mathematical plane of zero thickness (~: ~ 0) where the j u m p in p is concentrated,
we can ignore the anisotropy energy term and write (omitting constant terms)

3' =
x/)-g 2(cos q~2- cos ~1) - r/cos(r/2 -- ~1) (6.23)

where the coefficient r / = A'/~X/--A-K can take values f r o m -oo to 0 for antiparallel
coupling, and f r o m 0 to +o0 for parallel coupling across the A P B . F o r r / ~ oo the
difference between q~l and q~2 vanishes and we have the undisturbed wall with
3' = 4X/A--K following from eq. (6.22). Using standard differential calculus with
respect to the two variables ~pl and q~2 we find for r / > 1 stability at cos qh = l/r/
and cos ~0z = - l / r / , i.e., the wall is pinned with its centre coinciding with the A P B .
F o r r / < 1 the wall collapses into the A P B with q~l = 0 and ~2 = ~r. Such a
d e g e n e r a t e d wall will still be called a wall here. T h e energy of the pinned wall
follows f r o m eq. (6.22) with the a p p r o p r i a t e values of cos ~01 and cos q~2 substituted
as

y = 4 ( 1 - 1/2r/)N/)--K, for r/> 1

and

T=2r/X/A--K, for r/<l.


PERMANENT MAGNETS; THEORY 91

-5

I
1'0

"-5

--10

Fig. 36. Energy ~/of domain wall pinned at A P B in zero field as a function of the coupling p a r a m e t e r
~7 across the APB.

These relations are shown in fig. 36. N o t e that there is no discontinuity at r / = 1,


neither in value, nor in slope. Since we have chosen the uniformly magnetized
state with q~l = ~02 = 0 as the g r o u n d state with zero energy, the wall e n e r g y goes to
- ~ when ~ / ~ - ~ . This m e a n s that eventually the pinning b e c o m e s infinitely
strong. T h e d o m a i n wall is thus pinned at the A P B for any value of r/ in zero
field. To calculate the field that must be applied to detach the wall, we first have
to see h o w the pinned wall behaves in an applied field.

Energy of a pinned wall in an applied field. Consider a part of a wall stretching


f r o m x = - w where q~ = 0, to x = 0 w h e r e q~ = ~00. T h e angle ~0 is kept fixed and
the wall e n e r g y is calculated as a function of the f i e l d / - / a p p l i e d along the positive
z direction. T h e energy of this partial wall is

Y(~o, H ) =
F[ K sin e ~ + HJ(1 - cos ~ ) + A [ d~'~2]
\ d x ] J dx, (6.24)

where J is the saturation magnetization and the o t h e r symbols are as m e n t i o n e d


before. With variational calculus we find the condition for m i n i m u m energy to be

d2~p
2 K sin ~ cos ~ + H J sin ~p - 2 A ~ = 0,

which states that the t o r q u e is zero everywhere. Multiplication by dq~/dx and


integration f r o m - ~ to 0 give
92 H. ZIJLSTRA

{dq~ ,~2
K sin 2 q~0+ HJ(1 - cos q~0)= A \d-x-x] "

Substituting this into eq. (6.24) and switching f r o m x to ¢ as variable we have

Y(q~o, h) = 2 (1 - - COS 2 ~ - - 2h cos q~ + 2h) u2 d~o


~/ A K fO~°
= -2{[cos 2 Po + 2(h + 1) cos q~0+ 2h + 1] 1/2- 2(h + 1) 1/2

+ h ln[(cos 2 ~o + 2(h + 1) cos q~o+ 2h + 1) 1/2

+ cos ¢o + h + 1] - h In[2(h + 1) 1/2+ h + 2]}, (6.25)

where h = H/HA, and HA = 2 K / J is the anisotropy field.

Detachment of a pinned wall. Consider the wall in zero field symmetrically pinned
at x = 0 with q~2= ~" - qh. A field applied along the z axis will rotate ~Pl and q~2
towards 0, i.e., the centre of the wall will be shifted from x = 0 to the right in fig.
35. If the field is varied f r o m zero to increasing positive values, the force exerted
on the wall will increase, at first being in equilibrium with the rate of change of
wall energy. But as the latter quantity reaches a m a x i m u m a further increase of
the field detaches the wall from its pinning site and makes it travel to infinity. T h e
total energy of the wall when pinned is

3' : 3'(~1, h)+ 3'(¢2, h ) - n V ~ cos(~2- ~1),

where 7(qh, h) follows from eq. (6.25) with qh substituted for q~0, and 7(qh, h)
follows by substituting 7r - q~2 for ~0 and - h for h. T h e m i n i m u m of 3' with respect

0.5

\
0.2

0.1
\ \

0.05
\ \

\
\
0.02

0.01
0.1 0.2 0.5 1 2 5 10 20 50 100
_ ~ r]

F i g . 37. R e d u c e d c o e r c i v i t y h0 = Ho/HA d u e to pinning at APB a s a f u n c t i o n o f 7.


PERMANENT MAGNETS; THEORY 93

to the independent variables ~01 and ~02 is sought and the critical value h~ of h
where this extreme ceases to be a minimum is determined. This critical value h~ is
identified with the unpinning field or the coercivity and its relation with r/ is
shown in fig. 37. The curve applies only to positive values of ~7. At negative 7/ the
zero field values of q~l and ~2 are 0 and ~-, respectively, and it would take a
stronger field than h = 1 to detach the wall. However, at h = 1 uniform rotation
occurs and the whole concept of wall detachment becomes irrelevant.

6.5.3. Nucleation of a domain wall at an antiphase boundary


Consider the crystal of fig. 34 with the coupling p a r a m e t e r ~7 at its A P B smaller
than zero. If a strong positive field is applied a situation occurs as depicted in fig.
38(a). This is a m o r e or less saturated state which is stable, though not always of
the lowest energy, for all positive values of h including zero. If a counter field of
increasing strength is applied tO this state there is a critical value hc of - h at
which the symmetric configuration becomes unstable and a wall is emitted from
the defect, leaving the defect itself with an antiparallel magnetization orientation
as given in fig. 38(b). The relation between the critical field hc and the coupling
constant is calculated in the following.
Consider the configuration of fig. 38(a) in a field h. Near the A P B there are two
partial domain walls separated by an angle (~;2-~pl). The energy of this
configuration is

T = Y(~I, h ) + T(q~2, h ) - 7/N/A--K cos(~2 - ~1),

where ~/ is a negative n u m b e r and Y(q~l, h) and 7(q~2, h) are given by eq. (6.25)
with @1 and q~2 respectively substituted for ~P0. For equilibrium the partial
derivatives of y with respect to the independent variables ~/91and q~2 must be zero.
Since in equilibrium q)l --@2 for symmetry reasons, these two conditions reduce
~

to one:
2(1 - cos 2 q~l- 2h cos ~pl+ 2h) 1/2+ ~ sin 2q~1 = 0. (6.26)

q~

q01 -~ x r

a
Fig. 38. (a) Magnetization orientation near A P B after saturation in a strong positive field. (b) Wall
emitted from A P B by a negative field moves to the right, leaving the A P B in antiparallel configura-
tion.
94 H. Z I J L S T R A

The value of q~a in the remanent state (h = 0) is given by

cos ¢~ = 1 for 0 > 7 > -1


cosqh=-l/7 for 7<-1-

By investigating second derivatives we find that the equilibrium of eq. (6.26) loses
its stability when cos q~l + h ~<0. This gives a simple relation between the critical
field hc and the critical angle qh:

he = cos q~l •

Elimination of ~1 from this formula and eq. (6.26) then gives the relation between
hc and 7 as shown in fig. 39. The nucleation described here occurs only for
negative values of 7 and at a negative field which approaches zero for large
negative values of 7/. The Lorentz microscope observations by Lapworth and
Jakubovics (1974) and by Van Landuyt et al. (1978) of the APBs in thin slices of
MnA1 crystals being always decorated by a domain wall are explained by
assuming that a strong negative coupling at the APB is present.

6.5.4. Thin-wall coercivity in a perfect crystal


In crystals with a very high magnetocrystalline anisotropy constant K, comparable
to the energy density A / a 2 due to the exchange coupling, the wall thickness
t~ = 7r~v/A/K according to eq. (6.5) becomes of the order of the lattice parameter
a. In this case we must abandon the continuum approach of section 6.2.1 and
take into account the discreteness of the crystal lattice. It then turns out that there
is a difference in energy between a wall with its centre coinciding with a lattice

0.5

02 \

ho
l 0.1
"..
0.05

0.02

0.01
0.1 0.2 0.5 1 2 5 10 20 50
\ 100

Fig. 39. R e d u c e d c o e r c i v i t y hc d u e to n u c l e a t i o n of a wall at A P B as a function of - 7 .


PERMANENT MAGNETS; THEORY 95

9_ b

Fig. 40. Spin orientation in domain wall of fig. 19 viewed along a row of atoms normal to the wall
plane. (a) High-energy transitional position. (b) Low-energy stable position.

plane and a wall with its centre just b e t w e e n two lattice planes. T h e two situations
are shown in fig. 40(a) and (b), where the (b) configuration has the lowest energy.
T h e energy difference gives rise to pinning of the wall (Zijlstra 1970a).
T o calculate the wall e n e r g y we divide the wall along its thickness into three
zones: two continuous zones away from the centre, where the angle b e t w e e n
adjacent spins is small, and a central zone of N + 2 discrete magnetic m o m e n t s .
T h e continuous zones end in the first (or, as the case m a y be, the last) discrete
m o m e n t (fig. 41). T h e energy of this wall in a field h along the z-axis is

7 = y(q~0, h ) + ')/(~DN+I,h)
hK(2- cos q~0- cos q~N+I)
+ laK(sin2 q~0+ sin 2 q~N+l)+

+ ~, [aKsin2~i+2~A{1-cos(¢i-~oi-1)}+2 h K ( 1 - cos ~i)l


i=1 a

2A
-1-
a {1 -- COS(@N+ 1 -- qON)} , (6.27)

I I I I I

i i 1 x\
I t

Fig. 41. Domain-wall model consisting of centre part with discrete spins sandwiched between
continuous zones.
96 H. Z I J L S T R A

where y(q~0, h) is given by eq. (6.25) and 7(~Pl, h) follows from eq. (6.25) by
substituting (Tr - ~N+I) for ¢P0 and - h for h. The third term accounts for half of the
anisotropy energy of the spins 0 and N + 1, the other half being included in the
continuous zones. Similarly the fourth term accounts for the magnetostatic energy
of the same spins in the field h = HJ/2K (where J is the saturation magnetization
per unit volume). The fifth term is the sum of the anisotropy energy and the
magnetostatic energy of the spins 1 to N and the exchange-coupling energy
between spins 0 to N. The last term adds the exchange-coupling energy between
spins (N + 1) and N. The energy difference between the two spin configurations of
fig. 40 in zero field has been calculated by Van den Broek and Zijlstra (1971) by
minimizing eq. (6.27) with h = 0 for variations of ~i. The result is given in fig. 42
where the energy barrier is plotted as a function of ~7 = a2K/A. The angle
between the two central spins of the wall in its lowest state is shown in fig. 43 as a
function of ~7. For values of ~7 greater than 4 the wall collapses to a configuration
with one part of the spins oriented in the + z direction and the rest in the - z
direction.
The critical field hc required to let the wall pass the energy barrier is calculated
by minimizing eq. (6.26) and finding the value of h for which the equilibrium
becomes unstable. The result is given in fig. 44 where log hc is plotted as a
function of r/. With increasing ~7 we see that a sharp rise occurs at r / ~ 1. This
phenomenon is indeed found to occur in a family of lanthanide-transition metal
compounds investigated by Buschow and Brouha (1975) and Brouha and Buschow

10-1

aK
10-2

10.3

10-~

10-5 t
o 1 2 3 /.
_-~rt

Fig. 42. Energy difference between configurations of fig. 40(a) and (b) as a function of ~t = a2K/A.
PERMANENT MAGNETS; THEORY 97

TC
3
&,.p

0 I I I I I I I ~ ~ I
1 2
~-rt
F i g . 43. A n g l e b e t w e e n t h e t w o c e n t r a l s p i n s in t h e c o n f i g u r a t i o n o f fig. 4 0 ( b ) as a f u n c t i o n o f r/ a t z e r o
field.

f
0.5

hc

0.2

O.OE

0.02

O.Ol
o.1 0.2 0.5
j 1
Dq.
5 10 20 50 100

F i g . 44. R e d u c e d c o e r c i v i t y hc = He~HA as a f u n c t i o n o~ r/.

(1975). Their results are summarized in fig. 45, in which the coercivities of various
compounds measured at 4 K are plotted as a function of HA/Ixs, where /zs is the
magnetic m o m e n t per formula unit. This variable can be rewritten as HAI.~J~, in
which expression the n u m e r a t o r is proportional to K and the denominator is
proportional to A. There is only qualitative agreement since the proportionality
factor between K/A in the model and HA/I~ in the experiment is not known.
98 H. ZIJLSTRA

/
30 - [kOe) /A
o y (Co,Ni)s /
x Th {Co,Ni)s /
,, Lo (Co,Ni)5 /

2c I
I,,
ac "1
I
~o I
I
4
I x
I
i °° ':'~ " ~ 2
°o 10 20 30 0 50
(kOe/jaB )
HA/-IJs / F.U./
Fig. 45, Coercivity of various lanthanide-transition metal compounds at liquid helium temperature, as
a function of HA/t~s, which variable is proportional to ~7 (Buschow and Brouha 1975, Brouha and
Buschow 1975).

Thin-wall pinning is a homogeneous process where the wall is stationary below


the critical field and moves to disappear beyond this field. This gives rise to typical
rectangular hysteresis loops as observed by Barbara et al. (1971) on Dy3A12 and by
Buschow and Brouha (1975).
The similarity between thin-wall pinning at the atomic lattice and the Peierls
force in the description of moving dislocations in crystals has been discussed by
Weiner (1973) and by Hilzinger and Kronmfiller (1973). Thin-wall pinning in
dysprosium crystals was observed and discussed by Egami and G r a h a m (1971).
Hilzinger and Kronmfiller (1972) applied the model to the lattice of RCo5
compounds (R = rare-earth element) taking into account various orientations of
the wall plane with respect to the crystal axes. Quantitative comparison with
experiment is h a m p e r e d by insufficient knowledge of the magnitude of A and is
particularly difficult when m o r e than one kind of magnetic atom or m o r e than
one nearest neighbour distance is present. Also exchange coupling between other
than adjacent atoms, which is not u n c o m m o n in intermetallic compounds, com-
plicates the matter considerably.

6.5.5. Partial wall pinning at discrete sites


When a wall is pinned at discrete sites with the parts in between free to move like
a stretched m e m b r a n e the coercivity is determined by the competition between
the magnetic pressure on the entire wall and the local pinning forces at the
pinning sites. Consider a wall pinned at certain sites (fig. 46). Between the pinning
sites the wall is free to move under the influence of a field H to form cylindrical
PERMANENT MAGNETS; THEORY 99

Q7
Fig. 46. Wall pinned at discrete sites and bulging under the pressure of an applied field.

surfaces if it is pinned to parallel line defects, or look like a padded surface if it is


pinned to point defects. In both cases its area S varies by an amount

A S = o l ( x / a )2 (per unit S ) ,

where x is the m a x i m u m displacement, a is the distance between the pinning


sites, and a is a geometrical factor of the order one. At the same time a volume
A V reverses its magnetization direction:

AV=/3x (per unit S),

where /3 is another geometrical factor of the order one. The energy of the
magnetic system varies by

AE = yAS - 2HJ A V

= ay(x/a) 2 - 2flHJx, (6.28)

where J is the saturation magnetization and y is the wall energy. If the wall is not
parallel to the magnetization direction a magnetostatic term is contributed to the
wall energy. This situation occurs with the " p a d d e d surface" and also with
cylindrical surfaces whose axes deviate from the magnetization direction.
However, when the anisotropy field is large compared with the demagnetizing
field, i.e.,

K > 210 J2 ,
100 H. ZIJLSTRA

as is the case in the hard magnetic materials discussed here, the magnetostatic
contribution is small compared with the wall energy and may be ignored (Zijlstra
1970b, Hilzinger and Kronmfiller 1976).
The equilibrium position of the wall follows by minimizing h E with respect to x
for fixed H :

x = HYa2/y, (6.29)

where a and/3 are taken equal to one. When H is increased until it reaches the
pinning field strength Hp the wall is about to leave its pinning sites. The force exerted
on these sites by the wall is determined by the wall energy y and by the boundary
angle ~0 (see fig. 46). This angle is geometrically connected with the ratio x/a, and
the pinning field strength Hp can thus be related to a critical value (x/a)~rit which is
a property of the pinning sites. The coercivity H~ then follows from eq. (6.29)

(x) (6.30)
Hc = ~aa crit"

This formula is not to be taken too seriously in view of the many assumptions
made. However, the inverse proportionality of Hc with the distance a between the
defects implies

n c o( /,/1/3

where n is the density of point defects distributed in a regular array. If the defects
are distributed at random one might expect different results. Hilzinger and
Kronmiiller (1976) have made computer simulations of a moving wall pinned by
randomly distributed centres and found empirical proportionalities of Hc with n x,
with x ranging from 0.5 for weak pinning to 1.5 for strong pinning. Any
distribution of the inter-defect distance will give rise to a lowest unpinning field, at
which the wall will be detached somewhere. In the new wall position there will
again be a lowest unpinning field, which is not necessarily stronger than the
previous one. By this the observed coercivity will be lower than which follows
from the average a. Moreover, thermal fluctuations will cause wall creep at fields
weaker than the coercivity. Small applied-field fluctuations may have the same
effect. For this reason permanent magnets with their coercivity based upon wall
pinning at point or line defects must be considered as less stable than other
magnets. Substantial wall creep has been observed in copper-containing rare-earth
permanent magnet alloys which, indeed, have their coercivity determined by
wall-pinning at precipitates (Mildrum et al. 1970).

7. Influence of temperature
In the foregoing sections relations between coercivity and intrinsic properties like
anisotropy, magnetization and exchange coupling energy have been discussed.
PERMANENT MAGNETS; THEORY 101

The t e m p e r a t u r e dependence of the coercivity can be determined by measuring


the t e m p e r a t u r e dependences of the relevant intrinsic properties and using these
in the appropriate equations. However, by this procedure the discussion remains
essentially at zero t e m p e r a t u r e in the sense that thermal agitation of mag-
netization orientation or wall position is ignored. This thermal agitation may
lower the coercivity and, moreover, give rise to time effects like wall creep or mag-
netic viscosity. These time effects render a magnet unstable and should thus be
avoided. In this section we shall discuss and estimate roughly where the danger zones
are. For a critical review of theories pertaining to magnetic fluctuations the reader is
referred to Brown (1965, 1979). The n u m b e r of thermally excited events to occur per
second is given by

p = v e -aE/kr, (7.1)

where v is a frequency factor, indicating how m a n y times per second the event
could occur and AE is the energy barrier which the system under consideration
has to overcome for the event to occur. The barrier AE is c o m p a r e d with the
average energy of thermal motion kT, where k is Boltzmann's constant and T the
absolute temperature. At r o o m t e m p e r a t u r e k T ~ 4 x 1 0 -21 J.
First we consider the case of an isolated particle with shape or mag-
netocrystalline anisotropy as discussed in sections 3.2 and 3.3. The thermal motion
of the magnetization orientation is expected to have peak amplitude at the
precessional resonance of the magnetic vector in the anisotropy field. This field
will be of the order of 105-107 A m -1, thus giving rise to ferromagnetic resonance at
101°-1012 s -1. W e take these frequencies as the frequency v in eq. (7.1). W e now
compare two situations: a highly unstable one with p = 10 s-1 and a practically
stable one with p = 1 0 . 7 s -1. For these situations we calculate the required barrier
height AE using the values for u and k T mentioned (table 5). There are two
points worth noting: (a) The value of AE is not sensitive to the choice of v and (b)
p has a very strong dependence on AE, establishing a critical value AE-~ 10 19j
below which the system is thermally unstable at r o o m t e m p e r a t u r e and beyond
which it is stable. The barrier A E against rotation of the magnetization vector is
associated with the anisotropy constant K and the particle volume v

AE = Kv.

TABLE 5
I n s t a b i l i t y of m a g n e t i c particle d u e to
t h e r m a l a g i t a t i o n at r o o m t e m p e r a t u r e .

p (S 1) /,' (s-l) AE (j)

10 101° 0.8 x 10 -19


1012 1.0 X 10 19
10 -7 101° 1.6 X 10 -lv
1012 1.7 x 10 -19
102 H. Z I J L S T R A

For shape anisotropy K = 1 0 6 J m -3, which determines a lower limit for v =


10 -25 m 3 for stability. In alnico the elongated precipitates have a v o l u m e of the
order of 10 -24 m 3, which is above the critical value. H o w e v e r , Street and W o o l l e y
(1949) have o b s e r v e d viscosity effects in alnico which might be ascribed to this
relaxation effect. In any case, the barrier is lowered by an applied counter-field as
discussed in section 3.1 and b e c o m e s zero at He. So viscosity effects, h o w e v e r
small, are to be expected near the coercive field strength. F o r magnetocrystalline
anisotropy, K ~ 107jm -3, which determines a lower limit of v = 10-26m 3 for
stability. M a g n e t s based on magnetocrystalline anisotropy always consist of fine
powders, often sintered to a dense body. T h e crystallite size of these magnets is
typically 10-7-10 -6 m, which means a particle volume of 10-21-10 -18 m 3, well a b o v e
the critical volume.
A n o t h e r case to consider is the thermal activation of wall displacement. A wall
pinned over its entire area has in its potential well a resonance frequency given by

{dZY/m)l/2
w = \dx2 / , (7.2)

where 3' is the wall energy as a function of the distance x from its pinning site and
m is the wall mass associated with the precessional m o t i o n of spins inside a
moving wall. For 3' we write arbitrarily

y(x) = a70 + (1 - a ) y 0 , (7.3)

which means that 3' is lowered by a fraction a at the pinning site and is equal to
its undisturbed value 3'o when it is displaced by half the thickness 6 of the undis-
turbed wall. T h e effective mass of a wall has been discussed by D 6 r i n g (1948).
F r o m his work we have the relation

47r/z0 ( 1 )
m - 2c~26 = 2 a 2 6 , cgs , (7.4)

where a is the g y r o m a g n e t i c ratio of the spins in the wall. Substituting eqs. (7.3)
and (7.4) into eq. (7.2) and using eqs. (6.5) and (6.6) we have

o) ~- ( a o l 2 K X 106) 1/2 ,

where K is the magnetocrystalline anisotropy constant. With K = 107 Jm -3, a =


2 × 105 m A -1 s -1 and a = 0.1 we calculate

w ~6x 10 ~°s -~,


or

~ 10 lo s-1 ,
P E R M A N E N T MAGNETS; T H E O R Y 103

to be used in eq. (7.1). Now consider a part of the wall of area S displaced by a
distance ½6 from its pinning site (fig. 47). This excitation requires an energy of

A E = 7 ( a S + 28X/S), (7.5)

in which the second term accounts for the peripheral wall that has to be made. W e
may safely assume that an excitation with S < 62 will not give rise to an increasing
unpinning of the wall. We thus take the energy of an excitation with S = 62 as the
minimum 2~E for wall detachment and see whether this may occur at r o o m
temperature. The activation energy for this excitation is

A E = 23/6 2 ~ 20A6,

where the fraction a is ignored with respect to 2 and y6 is replaced by the


exchange p a r a m e t e r A times 10 (eqs. (6.5) and (6.6)). With A ~ 10 -11Jm -1 and
a ~ 10 -s m, as it is in highly anisotropic materials like SmCos, we have AE
10 -18 J. Substituting this into eq. (7.1) with the value of v = 10 l° s -1 we obtain the
n u m b e r of these excitations per second:

p= 10 -207S 1,

which means that in this example no thermal instability will occur.


In materials with extremely high anisotropy we have very thin walls of the order
of the lattice p a r a m e t e r as discussed in section 6.5.4. These walls are pinned by
the atomic lattice and we consider the excitation of a wall area with S large

\,
UFeU
S

pinning site

Fig. 47. Part of wall area S displaced by half its thickness ~ from its pinning site.
104 H. ZIJLSTRA

c o m p a r e d with the lattice p a r a m e t e r squared, S>> ~2. W e e s t i m a t e the wall


r e s o n a n c e f r e q u e n c y . T h e e n e r g y as a f u n c t i o n of wall position x is, to a first
approximation,

[2X\ 2
3'(x) = a3"o~--~.-) + (1 - a)3'0 , (7.6)

a n a l o g o u s to eq. (7.3) b u t n o w with half the lattice p a r a m e t e r ~: as the excursion


for which 3' has the value 3'0. W i t h 6 = 10 -9 m, ~ = 3 x 10 -20 m, 3'0 = 10 -2 J m -2 a n d
a = 0.1 we calculate, using eqs. (7.2), (7.4) a n d (7.6), that

w ~ 3 x 10 22s -1
or
1-'~'1012S 1.

T h e activation e n e r g y for the excitation is

A E = a3"oS,

which has to b e smaller t h a n 10-29J to p e r m i t m o r e t h a n o n e excitation per


second. W i t h the a s s u m e d values for a a n d 3'0 this gives a m a x i m u m value for the
wall area i n v o l v e d of

S = 10-27 m 2 ,

which is by two orders of m a g n i t u d e m o r e t h a n ~2 a n d thus c o n s i s t e n t with o u r


p r e s u p p o s i t i o n . T h e c o n c l u s i o n is that in materials with thin wall coercivity
t h e r m a l excitations m a y well occur that give rise to wall creep. Such creep has
b e e n observed, a.o., by B a r b a r a a n d U e h a r a (1976) a n d H u n t e r a n d T a y l o r (1977).
E g a m i (1973) has w r i t t e n an extensive theoretical t r e a t m e n t of the creep of thin
walls, taking into a c c o u n t b o t h t u n n e l l i n g a n d t h e r m a l excitation.

References

Aharoni, A. and S. Shtrikman, 1978, Phys. Rev. Levinson and D.W. Forrester, 1975, Phys.
109, 1522. Rev. Lett. 34, 594.
Aharoni, A., 1960, Phys. Rev. 119, 127. Brouha, M. and K.H.J. Buschow, 1975, IEEE
Aharoni, A., 1966, Phys. Stat. Sol. 16, 3. Trans. Magn. MAG-11, 1358.
Aharoni, A., 1962, Rev. Mod. Phys. 34, 227. Brown Jr., W.F., 1957, Phys. Rev. 105,
Barbara, B., B. B6cle, R. Lemaire and D. Pac- 1479.
card, 1971, J. Physique, C1-1971, 299. Brown Jr., W.F., 1962a, Magnetostatic Prin-
Barbara, B. and M. Uehara, 1977, Physica ciples in Ferromagnetism (North-Holland,
(Netherlands) 86-88 B + C, 1477, (Proc. Int. Amsterdam).
Conf. Magn., Amsterdam, 1976). Brown Jr., W.F., 1962b, J. Phys. Soc. Japan, 17,
Becket, J.J., 1969, IEEE Trans. Magn. MAG-5, Suppl. B-I, 540.
211. Brown Jr., W.F., 1963, Micromagnetics (Inter-
Berkowitz, A.E., J.A. Lahut, I.S. Jacobs, L.M. science/Wiley, New York).
PERMANENT MAGNETS; THEORY 105

Brown Jr., W.F., 1965, in: Fluctuation Luborsky, F.E. and C.R. Morelock, 1964, J.
Phenomena in Solids, ed., R.E. Burgess Appl. Phys. 35, 2055.
(Academic Press, New York) p. 37. Luteijn, A.I. and K.J. de Vos, 1956, Philips
Brown Jr., W.F., 1969, Ann. New York Acad. Res. Rep. 11, 489.
Sci. 147, 463. McCurrie, R.A. and L.E. Willmore, 1979, J.
Brown Jr., W.F., 1979, IEEE Trans. Magn. Appl. Phys. 50, 3560.
MAG-15, 1196. Margenau, H. and G.M. Murphy, 1956, The
Buschow, K.H.J. and M. Brouha, 1975, AIP Mathematics of Physics and Chemistry (2nd
Conf. Proc. 29, 618. ed.) (Van Nostrand, Princeton) p. 198.
Compaan, K. and H. Zijlstra, 1962, Phys. Rev. Martin, D.L. and M.G. Benz, 1971, Cobalt No.
126, 1722. 50, 11.
Craik, D.J. and E. Hill, 1974, Phys. Lett. 48A, Meiklejohn, W.H. and C.P. Bean, 1957, Phys.
157. Rev. 105, 904.
De Blois, R.W. and C.P. Bean, 1959, J. Appl. Mildrum, H., A.E. Ray and K. Strnat, 1970,
Phys. 30, 225S. Proc. 8th Rare Earth Research Conf., Reno,
De Jong, J.J., J.M.G. Smeets and H.B. Haan- 1970, p. 21.
stra, 1958, J. Appl. Phys. 29, 297. Mitzek, A.I. and S.S. Semyannikov, 1969,
De Vos, K.J., 1966, Thesis Tech. Univ. Delft. Soviet Physics-Solid State 11, 899.
D6ring, W., 1948, Z. Naturf. 3a, 373. N6el, L., 1944a, Cahiers de Physique, No. 25, 1.
Edwards, A., 1962, Magnet Design and Selec- N6el, L., 1944b, Cahiers de Physique, No. 25,
tion of Material, in: Permanent Magnets, ed., 21.
D. Hadfield (Iliffe, London) p. 191. N6el, L., 1954, J. Phys. Radium 15, 225.
Egami, T. 1973, Phys. Status Solidi (13) 57, Ojima, T., S. Tomizawa, T. Yoneyma and T.
211. Hori, 1977, Japan J. Appl. Phys. 16, 671.
Egami, T. and C.D. Graham Jr., 1971, J. Appl. Paine, T.O., L.I. Mendelsohn and F.E. Lubor-
Phys. 42, 1299. sky, 1955, Phys. Rev. 100, 1055.
Ermolenko, A.S., A.V. Korolev and Y.S. Shur, Schiller, K. and K. Brinkmann, 1970, Dauer-
1973, Proc. Int. Conf. on Magn., Moskow, magnete (Springer, Berlin) p. 74.
1973, Vol. I(2), p. 236. Shtrikman, S. and D. Treves, 1960, J. Appl.
Frei, E.H., S. Shtrikman and D. Treves, 1957, Phys. 31, 72 S.
Phys. Rev. 106, 446. Shur, Y.S., 1973, private communication.
Friedberg, R. and D.I. Paul, 1975, Phys. Rev. Stoner, E.C. and E.P. Wohlfarth, 1948, Phil.
Lett. 34, 1234. Trans. Roy. Soc. (London) 240A, 599.
Hilzinger, H.R., 1977, Appl. Phys. 12, 253. Street, R. and J.C. Woolley, 1949, Proc. Roy.
Hilzinger, H.R. and H. Kronmfiller, 1972, Phys. Soc. (London) A62, 562.
Status Solidi (B) 54, 593. Van den Broek, J.J. and H. Zijlstra, 1971,
Hilzinger, H.R. and H. Kronmfiller, 1973, Phys. IEEE Trans. Magn. MAG-7, 226.
Status Solidi (B) 59, 71. Van Landuyt, J, G. van Tendeloo, J.J. van den
Hilzinger, H.R. and H. Kronmfiller, 1976, J. Broek, H. Donkersloot and H. Zijlstra, 1978,
Magn. Magn. Mat. 2, 11. IEEE Trans. Magn. MAG-14, 679.
Hunter, J. and K.N.R. Taylor, 1977, Physica Weiner, J.H., 1973, IEEE Trans. Magn. MAG-
(Netherlands) 86-88 B + C (1), 161 0aroc. 9, 602.
Int. Conf. Magn., Amsterdam, 1976). Went, J.J., G.W. Rathenau, E.W. Gorter and
Jacobs, I.S. and C.P. Bean, 1955, Phys. Rev. G.W. van Oosterhout, 1951/52, Philips Tech.
100, 1060. Rev. 13, 194.
Jonas, B. and H.J. Meerkamp van Embden, Zijlstra, H., 1966, Z. Angew. Phys. 21, 6.
1941, Philips Tech. Rev. 6, 8. Zijlstra, H., 1967, Experimental Methods in
Kersten, M., 1943, Phys. Z. 44, 63. Magnetism, Vol. I (North-Holland, Am-
Kittel, C., 1949, Rev. Mod. Phys. 21, 541. sterdam) p. 135.
Kondorsky, E., 1952, Dokl. Akad. Nauk SSSR, Zijlstra, H., 1970a, IEEE Trans. Magn. MAG-6,
80, 197 and 82, 365. 179.
Lapworth, A.J. and J.P. Jakubovics, 1974, Proc. Zijlstra, H., 1970b, J. Appl. Phys. 41, 4881.
3rd. Eur. Conf. on Hard Magn. Mat., Am- Zijlstra, H., 1974, Philips Tech. Rev. 34, 193.
sterdam, 1974, p. 174. Zijlstra, H. and H.B. Haanstra, 1966, J. Appl.
Lilley, B.A., 1950, Phil. Mag. 41, 792. Phys. 37, 2853:
chapter 3

THE STRUCTURE AND PROPERTIES


OF ALNICO PERMANENT MAGNET
ALLOYS

R.A. McCURRIE
School of Materials Science and Technology
University of Bradford
Bradford, W Yorks BD7 1DP
UK

Ferromagnetic Materials, Vol. 3


Edited by E.P. Wohlfarth
© North-Holland Publishing Company, 1982

107
CONTENTS

1, Isotropic alnicos 1-4 . . . . . . . . . . . . . . . . . . . . . . 111


1.1. Fe2NiA1 and the isotropic alnicos 1-4 . . . . . . . . . . . . . . . 111
1.2. Microstructure and origin of the coercivity in Fe2NiA1 and the isotropic
alnicos 1-4 . . . . . . . . . . . . . . . . . . . . . . . . 113
2, Anisotropic alnicos 5 and 6 . . . . . . . . . . . . . . . . . . . . 121
2.1. Thermomagnetic treatment of anisotropic alnico 5 . . . . . . . . . . 121
2.2. Cyclic heat treatment of alnico 5 . . . . . . . . . . . . . . . . 129
2.3. Anisotropic cast alnico 5 with grain orientation (alnico 5 D G or alnico 5-7) . 129
2.4. Shape anisotropy of alnico 5 and alnico 5 D G (alnico 5-7) . . . . . . . . 131
2.5. Magnetostriction of alnico 5 and alnico 5 D G (alnico 5-7) . . . . . . . 133
2.6. Microstructures of alnico 5 alloys . . . . . . . . . . . . . . . . 134
2.7. Alnico 6 . . . . . . . . . . . . . . . . . . . . . . . . . 137
3. Anisotropic alnicos 8 and 9 . . . . . . . . . . . . . . . . . . . . 137
3.1. Thermomagnetic treatment of anisotropic alnico 8 . . . . . . . . . . 137
3.2. Extra high coercivity alnico 8 . . . . . . . . . . . . . . . . . 141
3.3, Anisotropic alnico 9 with fully columnar grains . . . . . . . . . . . 142
3.4. Shape anisotropy of alnicos 8 and 9 . . . . . . . . . . . . . . . 145
3.5. Microstructures of alnicos 8 and 9 . . . . . . . . . . . . . . . . 146
4. M6ssbauer spectroscopy of alnicos 5 and 8 . . . . . . . . . . . . . . 148
5. Sintered alnicos . . . . . . . . . . . . . . . . . . . . . . . . 148
6. Moulded, pressed or bonded alnico magnets . . . . . . . . . . . . . . 149
7. Effects of thermomagnetic treatment on the magnetic properties of alnicos 5-9 149
7.1. Factors controlling development of am particle shape anisotropy . . . . . 149
7.2. Relationship between the preferred or easy direction of magnetization
and the direction of the applied field during thermomagnetic treatment 151
7.3. Dependence of the magnetic properties on the direction of the applied
field during thermomagnetic treatment . . . . . . . . . . . . . . 151
8. Effects of cobalt on' the magnetic properties of the alnicos . . . . . . . . . 154
9. Effects of titanium on the magnetic properties of the alnicos
(mainly 6, 8 and 9) . . . . . . . . . . . . . . . . . . . . . . . 155
10. Dependence of the magnetic properties on the angle between the direction
of measurement and the preferred or easy axis of magnetization . . . . . . . 158
11. Relationship between magnetic properties and crystallographic texture . . . . . 161
12. Effects of particle misalignment on the rcmanence and coercivity of the anisotropic
field-treated alnicos . . . . . . . . . . . . . . . . . . . . . . 161
12.1. Remanence . . . . . . . . . . . . . . . . . . . . . . . 161
12.2. Coercivity . . . . . . . . . . . . . . . . . . . . . . . . 163

108
13. Determination of the optimum volume fraction of the F e - C o rich particles 164
14. Interpretation of the magnetic properties in terms of the Stoner-Wohlfarth theory
of hysteresis in single domain particles . . . . . . . . . . . . . . . . 166
15. Interpretation of the magnetic properties in terms of magnetization reversal
by the curling mechanism . . . . . . . . . . . . . . . . . . . . 169
16. Magnetostatic interaction domains in alnicos . . . . . . . . . . . . . . 170
17. Comparison of the N6el-Zijlstra and Cahn theories of magnetic annealing
in alnico alloys . . . . . . . . . . . . . . . . . . . . . . . . 171
17.1. N6el-Zijlstra theory . . . . . . . . . . . . . . . . . . . . 172
17.2. Cahn's theory . . . . . . . . . . . . . . . . . . . . . . 173
17.3. Discussion of the N6el-Zijlstra and Cahn theories . . . . . . . . . . 174
18. Rotational hysteresis . . . . . . . . . . . . . . . . . . . . . . 177
19. Anhysteretic magnetization . . . . . . . . . . . . . . . . . . . . 179
20. Magnetic viscosity . . . . . . . . . . . . . . . . . . . . . . . 179
21. Temperature dependence of magnetic properties . . . . . . . . . . . . 181
22. Dynamic excitation (AC magnetization) . . . . . . . . . . . . . . . 181
23. Prospects for impi-ovement in the magnetic properties . . . . . . . . . . 181
24. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . 182
References . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

109
1. Isotropic alnicos 1-4

1.i. Fe2NiAl and the isotropic alnicos 1-4

The alnicos are an important group of p e r m a n e n t magnet alloys. They contain Fe,
Co, Ni and A1 with minor additions of Cu and Ti. The first m e m b e r s of the series,
which do not contain cobalt, were discovered by Mishima (1932) and are known as
the Mishima alloys. They have a composition in the range 55-63% Fe, 25-30% Ni
and 12-15% A1, an energy product of ~ 8 kJm -3 and a coercivity of 4.8 x 1 0 4 A m
which is m o r e than twice the coercivity of the magnet steels which were available
in 1931. Because of their commercial importance the alnico alloys have been
studied in great detail by m a n y researchers.
Burgers and Snoek (1935) found that when an alloy containing 59% F e - 28%
N i - 1 3 % A1 was slowly cooled at a controlled rate from 1200°C to 700°C the
coercivity rose to a m a x i m u m of about 48 k A m 1 (600 Oe) as the cooling time was
extended and then decreased to about 16 k A m -~ (200 Oe) as the cooling time was
prolonged. X-ray investigation showed that in the o p t i m u m high coercivity state a
precipitation reaction had occurred. F r o m m e a s u r e m e n t s of the internal demag-
netization coefficient Snoek (1938, 1939) suggested that the alloys were hetero-
geneous and that in the optimum high coercivity state there were two ferro-
magnetic phases a~ and c~2.
The phase segregation process in Fe2NiAI has been investigated by Sucksmith
(1939) who measured the magnetization versus t e m p e r a t u r e curves of the single
and two-phase alloy. The latter was formed by quenching from 800°C and the
magnetization versus t e m p e r a t u r e curve showed a dip at 450°C which indicated
that the alloy was indeed two-phase. Sucksmith (1939) found that the phase
segregation occurred according to the reaction:

3.25 FesoNi25A125~ Fe9sNizsAlz5 + 2.25 Fe30Ni35A13s,

and that the saturation magnetizations of the two phases are, respectively
2 1 2 J T -1 kg -1 (212emu/g or 212erg Oe -1 g-a) and 6 1 J T 1 kg 1(61emu/g or
6 1 e r g O e -1 g-l). Since the densities of the two phases were not known the
saturation magnetic polarizations in teslas (T) could not be determined.
Details of the o p t i m u m composition and heat treatment of these isotropic

iii
112 R.A. M c C U R R I E

Fe-Ni-A1 alloys have been given by Betteridge (1939). The best properties were
obtained for an alloy containing 59.5% F e - 2 7 . 6 % Ni and 12.9% A1 which had
been quenched at 28°Cs I from the single phase state at l l00°C and then
tempered for 4 hours at 650°C. This treatment gave a coercivity BHc =
4 1 k A m - 1 ( 5 1 5 O e ) and a maximum energy product ( B H ) m ~ = 1 0 . 8 k J m -3
(1.35 x 106G Oe). The coercivity was shown to depend very critically on the
A1 content while the remanence depended more on the Ni content. Betteridge
(1939) also investigated the effects of adding Cu to the Fe-Ni-A1 alloys and found
a Cu addition of 3.5% increased (BH)max to 12kJm -3 (1.5 x 1 0 6 G O e ) after
quenching from above 950°C and tempering at 550°C. The Cu addition increased
the rate of precipitation so that the Fe-Ni-A1-Cu alloys required more rapid
cooling or quenching.
The effects of elastic stress on the precipitation and magnetic properties of
Fe-Ni-A1 alloys with additions of Cu and Ti have been investigated by Yer-
molenko and Korolyov (1970) who obtained improved optimum permanent
magnet properties of ( B H ) m a x = 16.8 k J m -3, BHc = 63 kAm 1 and Br = 0.67 T.
The permanent magnet alloys with compositions close to Fe2NiA1 are usually
prepared commercially by cooling from above 1250°C at an approximately con-
trolled rate. Rapidly cooled castings can be improved by annealing at 600°C for
several hours. These alloys, which usually have small additions of Cu, are known
as alni or alnico 3, in spite of the fact that they contain no cobalt; they are still
produced in small quantities.
Betteridge (1939), Zumbusch (1942a) and others also found that the magnetic
properties of the Fe-Ni-A1 alloys could be significantly improved by the addition
of cobalt as shown in fig. 1. The increase in remanence follows the increase in the
saturation magnetic polarization of the alloys while the larger relative increase in
the coercivity BHo can be attributed to an increase in the difference between the

100 - 1"Or 20
Br

~ 80- 0"8I 16
I
E
~6o
-,-°= ~.6 12 "J
,ff
BH(max)
g
8 -r
Ill
40 - 0-4 f
20 - 0.2 4

0 --0
0 5 10 15 20 25
Wt % Cobalt
Fig. 1. Dependence of coercivity, remanence and energy product on cobalt content (after Zumbusch
1942b).
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 113

saturation magnetic polarization of the F e - C o rich precipitate particles and the


Ni-A1 rich matrix. The addition of cobalt has the further beneficial effect of
raising the Curie temperature.
The cobalt-containing alloys are produced by controlled cooling from above
!250°C and subsequent annealing for several hours in the range 600-650°C. Since
cobalt decreases the rate of precipitation of the c~1 and ~2 phases it was found
essential to add small quantities of copper (Betteridge 1939) and to reduce both
the nickel and aluminium contents as the cobalt content was increased. It was also
shown (Betteridge 1939, Edwards 1957) that the magnetic properties of these
isotropic alloys (alnicos 1, 2, 3 and 4) could be significantly improved by the
addition of 4-5% Ti provided that the cobalt content was increased to 17-20%.
These extra high BHc isotropic alnicos (alnico 2) have coercivities in the range
60-72 kAm 2 compared with 36-56 kAm -1 for the isotropic alnicos 1, 2, 3 and 4.
The energy product of the high coercivity form of isotropic alnico 2 is also
slightly higher. Although the Ti addition reduces the remanence this is more than
compensated by the much higher coercivities. Details of the optimum composition
and heat treatments of the isotropic alnicos have been given by Betteridge (1939)
and Edwards (1957). The magnetic properties and compositions of the alnicos 1-4
are summarized in table 1 and typical demagnetization B - H curves are shown in
fig. 2.

0.8

0.6 lib
>2
u)
0"4
"0
X
-I
m
0.2 u.

I1 I / ! I I I I I 0
60 50 40 30 20 10 0
Applied field,H (kAm-1)
Fig. 2. Demagnetization curves for isotropic alnicos 1, 2, 3 and 4.

1.2. Microstructure and origin of the coercivity in Fe2NiAl and the


isotropic alnicos 1--4

The compositions of the alnicos are complex and the F e - N i - A I system is the only
one for which the phase diagram has been investigated in detail. Bradley and
Taylor (1938a, b) and Bradley, (1949a, b, 1951, 1952) established the positions of
the phase boundaries and the general metallurgical behaviour of the Fe-Ni-A1
system and showed that the alloys with potentially interesting permanent magnet
properties lay close to the line from Fe to NiA1 and were centered round the
114 R.A. M c C U R R I E

~ _'2

tt~
,q. t-- ~ t--

e.,

©
©
©

c~
< 0 ©

"~ 0 iq
?

0 ©
e~

0
c,.)

_= ii
< ?
<

II
E"

z
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 115

composition Fe2NiAI. The isothermal section of the Fe-Ni-A1 phase diagram at


750°C is shown in fig. 3. D e Vos (1966, 1969) has suggested that for the alloys
which are of most interest as permanent magnets the phase diagram for the
alnicos can be considerably simplified if it is assumed that they are pseudo-
binaries of Fe and NiA1 or F e - C o and Ni-AI. Marcon et al. (1978a) have shown
that in the Fe-Ni-A1 system this is not strictly correct because of the extension of
the a + T phase boundary below 9 1 0 ° C - t h e a ~ 7 allotropic transformation
temperature for Fe. The approximate pseudo-binary phase diagram shown in fig.
4 (Marcon et al. 1978a) is that for a vertical section cut through the line with full
circles in the phase diagram shown in fig. 3. Since the permanent magnet alloys
contain at least 50 wt % Fe it can be seen from fig. 4 that below about 800°C alloys
containing 25-75% Fe are two-phase (al and a2). When the sub-division into the al
and o~2phases occurs on a sufficiently fine scale the Fe or F e - C o rich particles have a
significant shape anisotropy, which gives rise to high coercivities according to the
Stoner and Wohlfarth (1948) model in which the magnetization is assumed to reverse
by coherent rotation, though we shall see later that this simple process is unlikely to
occur.
The Fe or F e - C o rich particles (o~1phase) and the non-ferromagnetic or weakly
ferromagnetic Ni-A1 rich matrix (c~2phase) have bcc structures and are formed by
spinodal decomposition (Cahn and Hilliard 1958, 1959, Cahn 1961-1963, 1968,
Hillert 1961) rather than a nucleation and growth process. We shall see that this
has important effects on the microstructure and magnetic properties of the alloys.
The spinodal decomposition of the c~ phase into the c~ and c~2 phases, although
spontaneous, is of course diffusion limited and can occur only at relatively high
temperatures ~850°C. The concentrations of the Fe or Fe and Co atoms in the
two phases vary periodically (assumed to be sinusoidal by Cahn (1962)) and the

Ni

7 5 0 °C

70

80

Fe to 20 3o ~o so so m so 9o AI
At % AI
Fig. 3. Ternary equilibrium phase diagrams for Fe-Ni-A1 at 750°C (after Bradley 1949, 1951, 1952).
116 R.A. McCURRIE

1800[
1 6 0 0 ~
1400~--~
\ o

0oov/
oor/ o,.o
oor/'
oor/'
O0 02 I. = 0'I 4 i 01.6 I 01.8 I
Fe Composition NiA
Fig. 4. Pseudo binary equilibrium phase diagram for Fe-NiA1 (after Marcon et al. 1978a).

amplitudes of the composition fluctuations increases with time until the phase
separation is complete; the whole process takes a very short time - seconds or
minutes.
Cahn (1962) has demonstrated theoretically that for alloys in which the elastic
constants obey the relation

2C44 - C l l ~- C12 > 0 ,

the spinodal decomposition waves are parallel to the three {100} planes. This
results in an initial microstructure which consists of a simple cubic array of regions
rich in one component (e.g. Fe or Fe-Co) connected along the (100) directions by
rods similarly enriched. The 'body centres' of this array are regions enriched in
Ni-A1 and are connected by similarly enriched (100) rods. This produces two
interlocking 3 dimensional systems of (100) rods, one enriched in Fe or F e - C o and
other enriched in Ni-A1. Such a micros~ructure is typical of the periodic dis-
tribution of the phases in spinodally decomposed systems and may be contrasted
with the irregular distribution of the phases in systems which decompose by
nucleation and growth. Cahn (1962) also predicted that the relative volume
fractions of the two phases also affects the final microstructure. If the volume
fraction of one phase is much smaller than the other the final microstructure will
consist of nearly equiaxed particles aligned along (100) directions. As the volume
fraction of this phase increases elongated particles are formed with their axes of
elongation parallel to the (100) directions, and these are of course the directions
of easy magnetization, though the alloys are macroscopically isotropic. Since the
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 117

volume fraction of the al phase is usually in the range 0.5-0.7, the observed
microstructures (figs. 6 and 7 and De Vos, 1966, 1969) are in very good agreement
with those predicted by Cahn's (1962) theory.
The formation and growth of the particles to their final shape and size occurs
almost entirely during the spinodal decomposition at 800-850°C. The driving force
for this reaction is of course the reduction in the interracial energy between the
particles (~1) and the matrix (a2). Although the interfacial energy is small
10-3-10 -1Jm -2 (1-102erg cm -2) this is sufficient to favour particle growth. The
principal effect of the heat treatment at 600°C is to increase the difference
between the saturation magnetic polarization of the Fe or F e - C o rich particles
and the surrounding matrix (Ni-A1 rich) by a continuous change in their com-
position due to the diffusion of Fe and Co atoms to the particles. The spinodal
decomposition into two phases does not, however, produce a very large shape
anisotropy in the ferromagnetic o~ phase particles and since the difference in the
saturation magnetizations of the a~ particles and the matrix is relatively small the
effective shape anisotropy field of the particles (proportional to Js(al)-Js(o~2)) is
also small in spite of the elongation (e.g. fig. 6 (De Vos 1966, 1969)).
Further heat treatment is necessary in order to increase the shape anisotropy
and hence to obtain the highest coercivities and the best permanent magnet
properties. This heat treatment usually consists of an anneal at about 600°C for
several hours, though this is sometimes omitted in the case of the cobalt free
alnico 3 alloys with compositions close to Fe2NiA1. The variation of the intrinsic
coercivity (He) with composition and heat treatment for Fe-NiA1 alloys (De Vos
1966, 1969) is shown in fig. 5. The coercivities were measured (a) after quenching and
tempering to give the optimum coercivity and (b) after continuous controlled cooling
of the alloys.
Since the interfacial energy depends on the crystallographic orientation of the
boundary between the a~ and o~2 phases the particle growth is anisotropic (though

Fe-NiAI

~'~6
I
E
<¢ a b

24
O

I i
o
o 20 40 60 8O lOO
Fe Composition, At.% NiAI
Fig. 5. Dependence of coercivity of Fe-NiAI alloys on composition and heat treatment (a) after
quenching and tempering for optimum properties (b) after continuous cooling (after De Vos 1969).
118 R.A. McCURRIE

only slightly so in Fe2NiA1) which results in an elongation parallel to the {100)


directions. The microstructures of the alnicos have been studied in considerable
detail by D e Vos (1966, 1969) who used replication electron microscopy. An
example of the microstructure of Fe2NiA1 after o p t i m u m cooling ( H c ~
56 k A m -1) is shown in fig, 6 though the shape anisotropy is not very pronounced.
If the alloy is given an isothermal heat treatment for 2 h a t 850°C followed by
quenching the preferential particle growth along the {100) directions is clearly
visible (fig. 7), though this microstructure does not give o p t i m u m p e r m a n e n t
magnet properties because the particles are too coarse and hence magnetization
reversal occurs by curling (see Zijlstra, chapter 2 of this handbook) or by a m o r e
complex mechanism.
F r o m fig. 7 we may assume that the particles are mainly rod-like (~500 nm long
and 100 nm in diameter) though particle coalescence and the presence of plate-like
particles gives rise to the observed more complex microstructure. Since the
Bloch-Lifshitz domain wall width in Fe or F e - C o alloys is of the order of 200 nm
the particles are probably 'single domains' though they are not expected to

Fig. 6. Electron micrograph of the oq + a2 structure in an arbitrary plane, for an alloy containing 50 at.
% Fe, 25 at. % Ni and 25 at. % AI, after optimum cooling (magnification 45000 x) (De Vos 1969).
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 119

Fig. 7. Electron micrograph of the o~l + ~e2 structure for an alloy containing 50 at. % Fe, 25 at. % Ni
and 25 at. % A1, after annealing for 2 h at 8500C followed by quenching (magnification 30000 x) (De Vos
1969).
behave as ideal isolated single domain particles with uniaxial shape anisotropy.
Even if Bloch-Lifshitz domain walls could be nucleated in the alloys it seems most
unlikely that they could move freely in such a complex interconnected micro-
structure, so, even this mechanism of magnetization change could lead to
significant coercivities. Assuming that the relatively high coercivities and
remanences in the alnicos l - 4 are due principally to shape anisotropy of elongated
Fe or F e - C o rich particles in a non-ferromagnetic matrix the Stoner-Wohlfarth
(1948) theory (for a summary see e.g. Zijlstra, chapter 2 of this handbook)
predicts that the coercivity is proportional to the saturation magnetization, Ms, of
the Fe or F e - C o rich particles and to a factor related to the difference in the
effective demagnetization factors perpendicular (Dz) and parallel (Dx) to the
preferred direction of magnetization in the particles, i.e.,

Hc = f(O)(Dz - D x ) M ~ ,

where f(O) is an averaging factor which takes account of the various orientations
of the preferred axes of the particles with respect to the direction of measurement
120 R.A. McCURRIE

of He. According to Stoner and Wohlfarth (1948) the coercivity of a random array
of non-interacting uniaxial single domain particles which reverse their mag-
netization by coherent rotation of the magnetization vector Ms is

Hc = 0.479(Dz - D,)Ms.

However, since the microstructures of Fe2NiA1 and the cobalt containing alnicos
1, 2 and 4 - p r e s u m a b l y similar to those of FezNiA1-are so complex, a detailed
quantitative interpretation of the magnetic properties has not yet been attempted.
In any case it is clear that a simple Stoner-Wohlfarth coherent rotation
mechanism of magnetization reversal is very unlikely to occur. Although the
particles have dimensions small enough to be single domains they are close to the
size where magnetization reversal is likely to occur by curling. The irregular shape
anisotropy of the interconnected particles may result in more complex
mechanisms of magnetization reversal due to the complex spatial variation of the
associated demagnetizing fields. The magnetization reversal is further complicated
by particle interactions (Wohlfarth 1955), interaction domains (Bates et al. 1962)
and the ferromagnetism of the Ni-AI rich a2 phase (see section 16).
The suggestion by Stoner and Wohlfarth (1947, 1948) that the relatively high
coercivities of the alloy FezNiA1 and the other isotropic alnicos, could be due
essentially to the shape anisotropy of elongated single domain particles has been
confirmed by Nesbitt et al. (1954). The magnetic behaviour of the material was
simulated by embedding fine wires of Ni-Fe-Mo permalloy (which have a pre-
dominant shape anisotropy) in a non-magnetic matrix, with their axes in the three
mutually perpendicular (100) directions viz. [100], [010] and [001]. By comparing
the field dependence of the torque curves for single crystals of Fe2NiA1 with those
of the simulated specimen Nesbitt et al. (1954) were able to show that the
structures of the two materials were essentially the same although the inter-
pretation of the results was complicated by the magnetocrystalline anisotropy of
the FezNiA1. Further complications arise because the particles in Fe2NiA1 often
coalesce to form very complex structures and shapes, as shown in figs. 6 and 7.
Since the microstructures of the cobalt-containing alnicos are very similar to those
observed in Fe2NiA1 it seems reasonable to assume that the coercivities of these
alloys are also due to shape anisotropy.
Apart from the dipole field interactions between the particles which depend on
the volume fraction p, the magnetic behaviour is further complicated by the fact
that in some alnico alloys the Ni-A1 rich a2 matrix phase is weakly ferromagnetic
so that there is an exchange coupling between the Fe-Co rich al phase and the
Ni-A1 rich a2 phase. Thus magnetization change may occur by some kind of 'wall
motion' through the complex interconnected al and Og2 phases. The domain walls
may of course only be 'interaction domain walls' as described in section 16 rather
than Bloch-Lifshitz walls, but the motion of these interaction domain walls still
requires magnetization reversal in the individual al phase particles and in view of
the particle interaction effects this itself will be a co-operative process. The
magnetization reversal may occur by an incoherent process such as curling but in
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 121

view of the complex shapes of the particles and the exchange coupling to the matrix
phase even more complex reversal processes may occur. In the high cobalt and high
titanium alloys where the particles are very regular in shape and are isolated single
domains embedded in a non-ferromagnetic matrix (c~2 phase), it seems that
magnetization reversal is likely to occur by the curling mechanism in individual
particles although these are still expected to act co-operatively and hence give rise to
domains or regions of reversed magnetization interaction domains. A fuller
discussion of the physical concepts briefly described here is given in sections 14, 15
and 16 as well as by Zijlstra, in chapter 2 of this handbook.

2. Anisotropic alnicos 5 and 6

2.1. Thermomagnetic treatment of anisotropic alnico 5

The most important advance in the technology of alnico magnets was the
discovery by Oliver and Shedden (1938) of the beneficial effect of thermomagnetic
treatment and its development by Jonas and Meerkamp van E m b d e n (1941).
Oliver and Shedden (1938) found that when an alloy containing 54% Fe, 18% Ni,
10% A1, 12% Co and 6% Cu was cooled from 1200°C in a magnetic field of
352 kAm -1 the resulting magnet was anisotropic and that the value of (BH)max was
increased from 12 kJm -3 (1.5 x 106 G Oe) in the isotropic alloy to 14.4 kJm -3 (1.8 x
106G Oe) parallel to the field direction. Perpendicular to the field (BH)max
decreased to 10.8 kJm -3 (1.35 x 1066 Oe). The anisotropy in (BH)max was due
entirely to changes in the remanence and fullness factor ((BH)max/BrHc) of the
demagnetizing B - H curves; the coercivity remained isotropic. Following this
discovery by Oliver and Shedden (1938), Jonas and Meerkamp van Embden
(1941) heat treated an alloy with a higher cobalt content (viz. 51.5% Fe, 14% Ni,
8.5% A1, 23% Co and 3% Cu) by cooling from 1200°C to 600°C in a magnetic
field of 240 kAm -1 and then annealing for several hours at 600°C. They obtained a
maximum energy product of 41.6 kJm -3 parallel to the field direction and only
5.6 kJm -3 perpendicular to it. For an isotropic alloy of the same composition the
maximum energy product was 17.6 kJm -3. Most of the improvement in the energy
product parallel to the field direction could be attributed to an increase in the
fullness factor though there was also a significant increase in the remanence from
0.87 T to 1.24 T.
The increase in the coercivity BHc was s m a l l - f r o m 47 kAm -1 to 52 kAm -1.
Jonas and Meerkamp van Embden (1941) also found that the application of a
magnetic field during the anneal at 600°C had no beneficial effect.
More detailed investigations of these anisotropic alnico 5 alloys were made by
Jellinghaus (1943) who studied the effects of varying the aluminium content in
alloys with the composition Fe, 15% Ni, 5.7-16.6% A1, 23% Co and 3% Cu. H e
found that good magnetic properties could be obtained only in the narrow range
of A1 content, about 8-9%, where the coercivitY , remanence and maximum
energy product were all high and that at the composition 50% F e - 23% C o - 15%
122 R.A. McCURRIE

N i - 9 % A 1 - 3 % Cu the application of a magnetic field during cooling increased


(BH)max very substantially from 12 kJm -3 (no field) to 40 kJm -3 (with field). The
dependence of (BH)m~x on the aluminium content and the effect of the ther-
momagnetic field is shown in fig. 8. This conclusion is supported by Lange (1968).
The presence of the deleterious fcc Y phase (Jellinghaus 1943) was detected only
for very low A1 contents and was absent above 6% A1 where the whole alloy had
a bcc structure.
Extensive investigations of alnico 5 alloys were made by Zumbusch (1942a, b)
who studied alloys with a wider range of compositions including an addition of up to
4% Ti. H e found that the best magnetic properties were obtained for alloys with
compositions in the range Fe, 14-15.5% Ni, 7.8-9.2% A1, 21.5-23.5% Co, 3-4%
Cu and less than 1% Ti. The heat treatment recommended by Zumbusch (1942a, b)
consists of cooling fro.m 1250°C at about l°Cs -1 to 500°C in a field well above
160 kAm -1, followed by tempering at 550°C for 10 hours.
The precipitation processes in alnico 5 have been investigated by Heimke et al.
(1966) who measured the saturation magnetic polarization, the remanence and
coercivity as a function of tim e using a modified thermomagnetic treatment. They
concluded that when the O/1 (Fe-Co) particles are first formed they are spherical
but just below the Curie temperature and under the influence of the applied field
they soon become elongated and form larger interconnected particles with their
long axes parallel to the applied field direction. Later or at lower temperatures
more cross links are formed between the particles.
The effect of the Cu content of alnico 5 has been investigated by Van der Steeg
and De Vos (1964) who showed that increasing the Cu content raised the
homogenization temperature, stabilized the 3' fcc phase and lowered the Curie
temperature and hence the temperature for effective heat treatment in a magnetic
field. They also found that the annealing process was accelerated by the addition

40
O o

"~"
I
30
E
~= (b)
¢x 20 -
E

10 - // (a)

0 ~1 i
4. 6 8 10 12 14 16 18
Wt % AluminFum
Fig. 8. Dependence of maximum energy product on the aluminium content of alloys containing
Fe-23% Co-15% Ni-3% Cu (a) without thermomagnetic treatment (b) with thermomagnetic
treatment (after Jellinghaus 1943).
STRUCTURE AND PROPERTIES OF ALNICOPERMANENTMAGNETALLOYS 123

of copper and that above 2.5% Cu the coercivity increased. Ritzow and Ebert
(1957) have shown that the presence of Cu decreased the rate of cooling through
the y fcc region necessary to prevent the precipitation of the magnetically
deleterious y fcc phase which must be avoided if the best magnetic properties are
to be obtained. All the anisotr0pic alloys (including alnicos 6-9) are single phase
above about 1200°C but if they are held in the temperature range 1000-1200°C for
more than a very short time the resulting magnetic properties are inferior due to
the precipitation of the ~ fcc phase. The formation of the latter phase is favoured
if the composition deviates significantly from the 'ideal' or if the alloys are
incompletely homogenized by soaking at 1250-1300°C. The a-~ y phase trans-
formation has been investigated by Koch et al. (1957, 1959), Koshiba and
Nishinuma (1957, 1960), Pater et al. (1963), Van der Steeg and De Vos (1964),
Planchard et al. (1964a-1966a), Heimke and Kohlhaas (1966). Koch et al. (1957,
1959) have shown by high temperature X-ray studies that the fcc y phase can
coexist with the bcc a phase in the temperature range 1250-800°C and that on
cooling below 800°C this fcc y phase transforms, by an apparently diffusionless
reaction to another bcc phase c~.
The magnetic properties of alnico 5 as a function of heat treatment have also
been measured by many other authors (see e.g. Clegg and McCaig (1957), Tenzer
and Kronenberg (1958), Yermolenko and Shur (1964), De Vos (1966, 1969),
Bronner et al. (1968-1970)).
The temperature dependence of the coercivity of alnico 5 (Clegg and McCaig
1957) after various heat treatments is shown in fig. 9. Curves (c) and (g) represent
specimens with optimum permanent magnet properties. The saturation magnetic
polarization of these specimens decreased almost linearly by 20-35% in the
temperature range 0-550°C. Above 550°C up to the Curie temperature (--~850°C) the
decrease in Js was very rapid.
For heat treatments above about 500°C the magnetic properties are not
reversible on cooling because at high temperatures fundamental structural
changes occur. The temperature stability of alnico alloys is discussed in section 21.
By considering magnetization versus temperature curves (for a review of this
technique see e.g. Berkowitz (1969)) Tenzer and Kronenberg (1958) concluded
that in the optimum permanent magnet state alnico 5 (52% Fe, 23% Co, 14% Ni,
8% A1, 3% Cu) consists of two ferromagnetic phases, one is mainly Fe2Co with
some A1, Ni and Cu (Js ~ 1.4 T) in solid solution and the other is mainly NiA1 with
a small Fe content.
By means of a similar study on alnico 5 (51% Fe, 24% Co, 14% Ni, 8% A1, 3%
Cu) Yermolenko and Shut (1964) concluded that in the optimum state the
saturation magnetic polarizations of the ~1 (Fe-Co) and a2 (Ni-A1) phases were
respectively 1.6 T and 0.1 T.
De Vos (1966, 1969) found that when the controlled cooling was terminated
by quenching from different temperatures the resulting magnetic properties were
nearly independent of the quenching temperature, provided the latter was below
about 775°C. The variation of the coercivity of alnico 5 with quenching tem-
perature before and after tempering at 585°C is shown in fig. 10.
124 R.A. M c C U R R I E

,!
0 0 0 0 O o
o
~
(L-tu~M) °H8 ~.~

~ - , r,- O

OE ~b

N
~N

0 0 0 0
(~-,,,~) 0~ ;" +

"~

~sU
s~
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 125

6 c r. ~ ~ Alnico 5

5 • After tempering
for 14 hr

r- 2
o
1
!
500 600 700 800 900
Quenching Temperature (°C )
Fig. 10. Dependence of coercivity of alnico 5 on quenching temperature before and after tempering at
585°C (De Vos 1969).

Wittig (1962) attempted to obtain high energy product, titanium-free alnico 5


magnets by isothermal heat treatment in a magnetic field at temperatures between
800 and 870°C for 2-20 min but found that the (BH)max products were lower than
those obtained with the conventional thermomagnetic treatment in agreement
with the earlier results obtained by Koch et al. (1957). However, Wittig (1962) also
found that a (BH)max = 40 kJm -3 (equal to that obtained by the conventional
thermomagnetic treatment) could be obtained if the alnico 5 was given a two-
stage isothermal heat treatment in a magnetic field as follows: (1) 4 rain at 860°C
in a magnetic field (presumably of saturating strength) (2) 2min at 830°C in a
magnetic field (3) tempering for 2 h at 650°C (4) tempering for 16 h at 585°C.
The heat treatment of alnico 5 varies according to the composition and
manufacturer; a typical heat treatment is as follows (Marcon et al. 1978b):
(1) Homogenization at 1350°C to obtain the a phase.
(2) Fast cooling at 4 ° C s -1 down to about 900°C to prevent the parasitic
precipitation of the fcc 3' phase.
(3) Controlled cooling in a magnetic field (ideally of saturating strength
~320 kAm -1) down to 600°C. This treatment results in the spinodal decom-
position of the a phase into the al (Fe-Co) and (1' 2 (Ni-A1) phases and the
formation of elongated a~ (Fe-Co) particles.
(4) Tempering for 6 h at 650°C followed by 24h at 550°C. This treatment
produces the equilibrium al and a2 phases which have a maximum in the
difference between their saturation magnetic polarizations AJ = J~l - JR.
Many variations in the above process can be made without producing significant
changes in the magnetic properties; e.g., average cooling rates in the range
0.5-4°Cs -1 have b e e n quoted by many authors (see, e.g., Nesbitt and Williams
1957, Clegg and McCaig 1957, Povolotskii et al. 1963, Marcon et al. 1978b).
Marcon et al. (1978b) have also investigated the effects of varying the applied
126 R.A. M c C U R R I E

field during cooling (from 900 to 600°C) over the range 0 - 5 7 6 k A m -1. They
found that for polycrystalline alnico 5 (50.6% Fe, 13.9% Ni, 8.2% AI, 24.3% Co,
3% Cu) the optimum permanent magnet properties could be obtained by cooling
in a field of 80kAm-1; cooling in larger fields up to a maximum 5 7 6 k A m -1
produced no significant improvement in Br, BH¢ or ( B H ) . . . .
The thermomagnetic treatment of the alnico 5 alloys is most effective when they
contain about 23% Co or more. The high cobalt content is essential because it
raises the Curie temperature to about 850°C so that when the spinodal decom-
position occurs in the temperature range 700-850°C the shape and growth of the
particles can be influenced by the applied field because they are already ferro-
magnetic.
The effects and theory of magnetic annealing and the relationship between the
preferred direction of magnetization in alnico 5 and the applied field during
cooling are discussed later in sections 7 and 17.
The improvement in the magnetic properties of alnico 5 occurs only in the
direction of the applied field and there is an accompanying deterioration of the
properties in other directions. These alloys, though polycrystalline, are anisotropic
owing to the preferential growth of the particles along the field direction or as
near to it as possible depending on the orientation of the [100] directions in the
grains.
We can see from table 2 that the improvement in the remanence and energy
product of alnico 5 compared with those alnicos 1-4 (table 1) is very significant. A
typical demagnetization curve for alnico 5 (random grain) is shown in fig. 11:
(the demagnetization curves for columnar and single crystal alnico 5 are also
shown in fig. 11).
In dynamic applications of alnico 5 where the working point of the magnet

B. (kJn -3)
80 60 40 11.4

..~" H 1.0~

>~
" 16.1"~~~ 0-40"6"o;
•'=

Alnico 5
-%
~
0-2
I I i ' ' 0
80 70 60 50t''- 40 30 20 10 0
Applied field, H (kAm-1)
Fig. 11. Demagnetization curves for various forms of alnico 5: (a) equiaxed alnico 5, (b) grain orientated
alnico 5-7, (c) single crystal alnico 5.
S T R U C T U R E A N D P R O P E R T I E S OF A L N I C O P E R M A N E N T M A G N E T A L L O Y S 127

tt3

k~

"7

"2.

m.
"7

E 6
C~

C4

~5
'~
eq 0
0
03 eq P~
eq eq ¢¢3 ¢¢3 eq

H
tt~ ~ ,.~ tt~ tt3
,....,
-4 .,

e~
o b~
0
l G,
oq. zq. r-- r-- oq.
p., p..

If
Go
T
< tt')

C~

d
0 z
q;
tl

"-6 t:'-6 ~. , " ~ ~'r-


ca ~ k~
128 R.A. McCURRIE

changes, details of the recoil hysteresis loops and the recoil energy contours are
essential if the material is to be used efficiently. Hysteresis loops and their
associated recoil curves for various alnico alloys have b e e n published by Stfiblein
(1968). T h e recoil energy c o n t o u r s for various alnicos can be f o u n d in the b o o k by
Parker and Studders (1962). Typical recoil loops for alnico 5 are shown in fig. 12
which also shows various load lines and working points. T h e recoil energy
contours for alnico 5 are shown in fig. 13. N o t e that the recoil energies given in

29~2~ BIT]

I J r = 2 ~ 11"0

0.8

0"6

0"4

0"2

0
60 40 20 0
H I kAm-l]
Fig. 12. Typical recoil loops for alnico 5, and various load lines and working points (after Stiiblein 1968).

B (T)
1.2

1.0

0-8

0.6

0.4

0-2

0
-5 -4 -3 -2 -1 0
H (10 4 Am - 1 )
Fig. 13. Recoil energy contours for alnico 5 (after Parker and Studders 1962).
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 129

fig. 13 (usually referred to as the useful recoil energies) are those obtained by
using the more practical definition used by permanent magnet designers (Edwards
1962, Parker and Studders 1962, Schiller and Brinkmann 1970) rather than the
fundamental definition given by Zijlstra in chapter 2 of this handbook.
The technological importance of alnico 5 alloys is reflected by the large number
of published papers which have appeared since their discovery. For further details
the reader is referred to the papers by Bronner et al. (1967-1969, 1970a) and Gould
(1971) who have attempted to interpret the magnetic properties in terms of the
Stoner-Wohlfarth (1948) theory. Kolbe.and Martin (1960) have shown that alnico
5 can be successfully shaped into rod, wire or strip by hot working at high
temperatures. They have also shown that alnico 5 can be extruded swaged and
rolled.

2.2. Cyclic heat treatment of alnico 5

Hansen (1955) showed that the magnetic properties of alnico 5 were impaired by
heat treating at 620-790°C but could be restored by re-tempering at 585°C provided
the spoiling heat treatment temperature was below 675°C. After spoiling heat
treatment above this temperature the coercivity could only be partially restored.
Similar observations were made by Koch et al. (1957) and Fujiwara and Kato
(1960). D e Vos (1966) investigated the microstructures of an alnico 5 alloy after it
had been given the following treatments: (1) after optimum cooling in a magnetic
field (BHc = 44 kAm -1) (2) treatment (1) followed by spoiling at 750°C for 3 min
(Uric = 8.8 kAm -1) (3) treatments (1) and (2) followed by re-tempering at 585°C
(BHc = 54 kAm-1). By observing the microstructure in a plane parallel to the
direction of the field during cooling, De Vos (1966, 1969) found that there was no
significant change in the morphology of the structure after each stage of the above
cyclical heat treatment.
According to De Vos ( 1 9 6 6 ) t h e best magnetic properties of alnico 5 are
obtained when optimally field cooled magnets are aged below 600°C. The mag-
netic properties as a function of heat treatment in the range 550-750°C have been
tabulated by De Vos (1966). H e found that the optimum permanent magnet
properties: (BH)~ax = 48 kJm -3, BHc = 52 kAm -1, Br = 1.3 T and Bs = 1.4 T can be
obtained by optimum field cooling followed by heat treatment for 8 h at 585°C.

2.3. Anisotropic cast alnico 5 with grain orientation (alnico 5 D G or alnico 5-7)

It has been established by Ebeling and Burr (1953) and Zijlstra (1956) that the
best magnetic properties of alnico 5 are obtained when single crystals are given a
thermomagnetic treatment with the applied field parallel to one of the [100]
directions. Hoselitz and McCaig (1949b) showed that a further improvement in
the magnetic properties of field cooled polycrystalline alnico 5 can be obtained if
the grains have a preferred orientation so that the al (Fe-Co) particle axes in the
whole alloy are as nearly aligned as possible.
Fortunately, if iron-base alloys such as the alnicos are allowed to solidify on a
130 R.A. McCURRIE

cold surface the [100] axes tend to grow preferentially perpendicular to the cold
surface so that the desired grain orientation or columnar grain structure can be
produced relatively easily. Thus when alnico 5 alloys are cast with a columnar
crystallization or preferred grain orientation thermomagnetic treatment with the
applied field parallel to the preferred or columnar axis results in a significant
improvement in the magnetic properties compared with the alnico 5 alloys with
randomly oriented grains. This result can be readily appreciated by comparing the
demagnetization curves shown in fig. 11.
In order to obtain well oriented columnar magnets it is necessary to use
pre-heated or exothermic moulds and to cast the magnets on steel or copper slabs
(preferably water cooled). A typical columnar grain structure of an alnico 5-7
casting is shown in fig. 14 from which it can be seen that there is a high degree of
alignment but this is by no means fully columnar. The production and magnetic
properties of alnico 5-7 have been described and investigated by Ebeling and
Burr (1953), McCaig and Wright (1960), Makino (1962), Makino et al. (1963),
Makino and Kimura (1965), Lindner et al. (1963) and Gould (1964, 1971). The
compositions and magnetic properties of alnico 5-7 columnar alloys (laboratory
specimens) are shown in table 3. Unfortunately magnets with energy products as
high as 70 kJm -3 do not appear to be available commercially. Typical magnetic
properties of commercial alnico 5-7 are shown in table 3.
Alnico 5 and alnico 5-7 or their equivalents are among the most widely used
permanent magnet alloys. Many millions are used as loudspeaker magnets, though

Fig. 14. Typical columnar grain structure of alnico5 D G (alnico 5-7) (Gould 1971). (Magnification: ~2x.)
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 131

TABLE 3
Magnetic properties of alnico 5-7 (columnar) alloys (after Gould 1964).

Composition (wt %)
Br BH¢ (BH)m.x
Fe Ni Co Al Cu Nb (T) (kAm ~) (kJm -3)

51 14 24 8 3 - 1.33 56 51

50.7 13.5 24 8 3 0.8 1.33 59 57

50.3 13.8 24.3 8.4 3.2 - 1.33 58 64

49.7 13.5 25 8 3 0.8 1.43 62 70

49.7 13.5 25 8 3 0.8 1.35 63 65

49.7 13.5 25 8 3 0.8 1.38 57 58

50.5 13.5 25 8 3 - 1.4 56 64

49.7 13.5 25 8 3 0.8 1.4 60 69

48.5 13.5 25 8 3 2.0 1.37 66 69

47.3 17.7 24 8 3 - 1.4 59 64

N.B. 1 T = 104 G; 1 A m -a = (4~-/1000) Oe; 1 Jm -3 = 407r GOe.

they are gradually being replaced in this application by oriented BaFel20~9 and
SrFe~2019 magnets, which are very much cheaper.
It can be seen from fig. 11 that the demagnetization curve for single crystal
alnico 5 gives the best magnetic properties with an energy product of 80 kJm -3
compared with ~60 kJm -3 for columnar alnico 5-7 and ~42 kJm 3 for equiaxed
polycrystalline alnico 5 (Cronk 1966). By using a technique involving secondary
recrystallization of large grained alnico 5 (~51% Fe, 24% Co, 14% Ni, 8% A1,
3% Cu, contaminated with 0.08% C or 0.35% Mn) Steinort et al. (1962)
succeeded in obtaining a single crystal of mass ~0.11 kg with a maximum energy
product of 88 kJm -3. From fig. 11, demagnetization curve (c), it is clear that the
single crystal alnico 5 has significantly better magnetic properties than alnico 5
and alnico 5-7, but unfortunately single crystals are difficult to produce on a
commercial scale.

2.4. Shape anisotropy of alnico 5 and alnico 5 D G (alnico 5-7)

From torque curve measurements on alnico 5 and alnico 5 D G (i.e. alnico 5 with a
columnar grain structure also known as alnico 7) Hoselitz and McCaig (1951,
1952) have concluded that the materials are magnetically uniaxial with an aniso-
tropy coefficient of Ku ~ 105 Jm -3 (106 erg cm-3). A similar result was obtained by
132 R.A. McCURRIE

Nesbitt and Heidenreich (1952). From electron micrographs and torque curve
measurements on single crystals of alnico 5 Nesbitt and Williams (1955) concluded
that the high coercivity was due to the shape anisotropy of small elongated
particles (approximately 0.41xm long and 0.061xm in diameter) and that the
magnetocrystalline anisotropy was negligible. Typical torque curves for a disc with
a (110) plane as surface are shown in fig. 15 from which can be concluded that the
uniaxial shape anisotropy coefficient Ku is 1.08 x 105 Jm -3 with [001] as the easy
direction of magnetization. Similar torque curve measurements of the anisotropy
constants Ku for various thermomagnetic treatments of single crystals of alnico 5
have been made by Yermolenko et al. (1964). Instead of the conventional
thermomagnetic treatment (i.e. controlled cooling in a magnetic field) they used
isothermal heat treatment for various times in a field of 720 kAm -1 (9000 Oe) on
specimens which had been quenched from 1300°C. This procedure followed by
more complex tempering schedules did not produce specimens with optimum
permanent magnet properties. However, for an alnico 5 alloy which had been heat
treated in a field parallel to the [001] direction (which became the preferred
direction) they obtained from the torque curve in the (110) plane a value of
Ku = 0.84× 105jm -3, Sergeyev and Bulygina (1970) investigated the magnetic
properties and the magnetic anisotropy at various stages in the heat treatment of
single crystals of alnico 5. Some of their results are summarized in table 4 from
which it can be seen that the increases in the coercivity and the energy product
after tempering are accompanied by an increase in the anisotropy constant Ku,
which can be attributed to an increased elongation of the particles and to changes

12

8 Alnico 5

[E
4
O
m-

I,,.

=
O
I---4
0 I
20
|
40
i
60
/ I /
801
i
100
I
120
Angle 0 from o
i
140
01]
I
160
b
180

-8 Applied field = I ' 6 M A m -1

-12 I
Fig. 15. Torque curve in the (ll(I) plane of a single crystal of alnico 5 (after Nesbitt and Williams
1955).
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 133

TABLE 4
Magnetic properties and anisotropy constant Ku at various stages in the heat treatment of alnico 5
single crystals.

Heat treatment B BHc (BH)max Ku


(T) (kAm-1) (kJm-3) (Jm 3)

A Homogenization at 1300°C,
followed by cooling to 9 0 0 ° C . . . .
at a rate of 3.5°Cs-1.

1 A and cooling in a field from


900°C to 600°C at a rate of 1.4 38.4 38.4 1.07
0.5oCs i.

2 1 and tempering at 640°C for


2 h. 1.38 .47.2 43.2 1.30

3 2 and tempering at 560°C for


10 h. 1.37 52.8 60 1.52

N.B. 1 T = 10 4 G; 1 Am-1 = (4~r/1000) Oe; 1 J m -3 = 407r GOe.

in their composition and hence the difference in the saturation magnetic polariza-
tions of the al particles and the matrix, i.e. A J = Jal - Ja2.
M o r e recently T a k e u c h i and I w a m a (1976) cooled a single crystal of alnico 5
(containing 51.31% Fe, 14.04% Ni, 7.69% A1, 23.88% C o and 3.08% Cu) in a
magnetic field of 6 4 0 k A m -1 parallel to the [100] direction (which of course
b e c a m e the preferred direction of magnetization) and obtained a value of Ku =
1.56× 105 Jm -3 and an associated magnetocrystalline anisotropy constant K1 =
0.15 × 105 J m -3, a result which suggests that the p e r m a n e n t m a g n e t properties of
alnico are due principally to the shape anisotropy of the a~ ( F e - C o ) particles and
that the contribution f r o m the magnetocrystalline anisotropy is relatively small.
A technique for determining the magnetic anisotropy energy of alnico magnets
has been developed by Allec (1971). This is based on a m e t h o d originally used on
single crystals by Guillaud (1953) and enables the anisotropy constants to be
d e t e r m i n e d f r o m the magnetization curves in various directions for single crystal,
c o l u m n a r and polycrystalline specimens.

2.5. Magnetostriction of alnico 5 and alnico 5 D G (Alnico 5-7)

T h e magnetostriction in alnico 5 has been investigated in considerable detail by


Hoselitz and McCaig (1949) and McCaig (1949). T h e y m e a s u r e d the mag-
netostriction as a function of the direction of magnetization and the direction of
m e a s u r e m e n t for specimens of alnico 5-7 which had been cooled from 1300°C in a
magnetic field (a) parallel to the c o l u m n a r crystal axis and (b) perpendicular to the
c o l u m n a r axis. T h e saturation magnetostriction, hs, was shown to d e p e n d very
strongly on the particular set of angular relationships (angles were either 0 ° or 90 °)
134 R.A. McCURRIE

between the columnar axis, the field direction during cooling, the direction of
magnetization and the direction of m e a s u r e m e n t of &. The observed values of A~
(McCaig 1949) ranged from - 6 1 x 10 -6 to 54 x 10 -6.
McCaig (1949) also made measurements of As as a function of the direction of
magnetization and the direction of m e a s u r e m e n t for alnico 5 specimens with (a)
columnar crystals (anisotropic) and (b) randomly oriented crystals (isotropic)
which had n o t been cooled in a magnetic field. The observed values of A~ for (a)
also depended very strongly on the angular relationship (angles either 0 ° or 90 °)
between the columnar axis, the direction of magnetization and the direction of
m e a s u r e m e n t of As and ranged from - 7 . 5 × 10 .6 to 34.2× 10 .6 . For the (b)
specimen with randomly oriented crystals (i.e. isotropic) the longitudinal satura-
tion magnetostriction was 20 x 10 .6 while that measured perpendicular to the
direction of magnetization, i.e., the transverse saturation magnetostriction was
--8 X 10 -6.
An extensive investigation of the magnetostriction of alnico alloys has been
made by Nesbitt (1950) who showed that for polycrystalline alnico 5 alloys the
saturation magnetostriction constants As depended very strongly on the heat
treatment; the observed values of As were in the range 2.5 to 43 x 10 -6 which
includes both longitudinal (i.e. parallel to the direction of the field applied during
cooling) and transverse measurements of &. The low longitudinal value (2.5 ×
10 -6) is due to the fact that the magnetization vectors in most of the interaction
domains (see section 16) are either parallel or antiparallel to the applied field and
hence at an angle of 180 ° to the domain magnetization and therefore do not
contribute to a change in length of the specimen as a whole. The high transverse
value (43 x 10 -6) is due to the fact that after cooling in a transverse field (i.e. in the
direction perpendicular to that in which the magnetostriction is measured) most of
the magnetization vectors of the interaction domains are perpendicular to the
direction of m e a s u r e m e n t of As and so after magnetization in the measuring field
each domain contributes to the overall change in length of the specimen. For an
alnico 5 which was cooled from 1300°C at 2°Cs -1 (i.e. at the rate necessary to give
o p t i m u m p e r m a n e n t magnet properties) the saturation magnetostriction & after
tempering was 35 x 10 -6, whereas a specimen of alnico 5 which was given the
commercial thermomagnetic and tempering treatment for o p t i m u m permanent
magnet properties was found to have a saturation magnetostriction ),s of only
7.5 × 10 -6.

2.6. Microstructures of alnico 5 alloys

The first attempts to determine the microstructures of alnico 5 alloys were made
by Heidenreich and Nesbitt (1952) who used an oxide replica technique and
electron microscopy. Unfortunately they were unable to resolve the micro-
structure in the optimum permanent magnet state. The microstructures and
particle dimensions which they reported were obtained from specimens which had
been over-aged by heat treatment at 800°C in order to m a k e the microstructure
visible. Heidenreich and Nesbitt (1952) observed a rod-like precipitate which
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 135

tended to grow along (100) directions and to group into plate-like arrays. When
there was no applied field during the heat treatment the cubic symmetry of the
rods (i.e. their long axes were parallel to the (100) directions) was very clearly
discernible. However, when the alnico 5 was heat treated in an applied field the
long axes of the particles were confined to a single (100) direction when the field
was parallel to that direction. By extrapolating their results to zero time at 800°C,
Heidenreich and Nesbitt (1952) estimated that the particle dimensions were
7.5 × 7.5 × 40 nm. They also concluded that the particles were F e - C o rich and had
a Curie temperature higher than that of the Ni-A1 rich matrix.
By using very thin thermally grown oxide replicas and electron microscopy
Kronenberg (1954) was able to observe the microstructure of alnico 5 in the
optimum permanent magnet state but the results were inconclusive. Fahlenbrach
(1954, 1955, 1956) used a refinement of the oxide replica technique and was able
to show that after cooling in a magnetic field and tempering to give optimum
permanent magnet properties the particles (Fe-Co) were rods 40 nm in diameter
and 100rim long. Similar results were also obtained by Schulze (1956) who
estimated the particle dimensions to be 36nm in diameter and 120nm long.
Further improvements were made by Haanstra et al. (1957) and de Jong et al.
(1958) who obtained very clear electron micrographs of alnico 5 and alnico 8 in
the optimum permanent magnet state by using carbon replicas. They also con-
cluded that the F e - C o rich particles were rods (approximately 30 nm x 30 n m ×
120 nm) and that their long axes were parallel to the [100] direction which was
closest to the direction of the applied field during the thermomagnetic treatment.
The most comprehensive study of the microstructures of the alnicos (including
Fe2NiA1) has been made by D e Vos (1966, 1969) who confirmed all the above
conclusions concerning the microstructure of alnico 5 - an example of the latter is
shown in fig. 16, from which it can be seen that the particles are about 150 nm
long and about 40 nm in diameter. De Vos also showed that the higher coer-
civities which are observed in alnicos 8 and 9 can be attributed to the very high
degree of particle elongation, perfection and alignment which results from the
higher Co content and hence their greater sensitivity to thermomagnetic treat-
ment (Zijlstra 1960-1962). Further confirmation of these microstructures has been
made by Granovsky et al. (1967) and Pashkov et al. (1970).
The microstructures of alnico 5 alloys have also been investigated by Pfeiffer
(1969), Bronner et al. (1967), Mason et al. (1970), Nicholson and Tufton (1966)
and Kronenberg (1960, 1961) who used transmission electron microscopy.
From the above electron microscopy studies of alnico 5 it can be concluded that
the volume fraction, p, of the o~1 (Fe-Co) phase is in the range 0.6-0.7, and that
the degree of elongation (i.e. the ratio of the length to the diameter of the particles,
l/d) is in the range ~ 4 - 6 . The observed values of p and I/d depend of course on the
composition, thermomagnetic treatment and the temperature and time for which the
alloys are annealed.
According to Arbuzov and Pavlyukov (1965) both the c~1 (Fe-Co) and O~2
(Ni-A1) phases are ordered and have the bcc structure. However, Granovsky et al.
(1967) and Pashkov et al. (1969) have shown by very careful X-ray diffraction
136 R.A. McCURRIE

l
a

b
Fig. 16. Microstructures of alnico 5; (a) in a plane parallel to applied field H and (b) in a plane
perpendicular to H (magnification 45500 x) (De Vos 1966).

studies t h a t b o t h t h e a~ a n d og2 p h a s e s a r e very slightly t e t r a g o n a l . F o r alnico 5 the


l a t t e r a u t h o r s f o u n d for t h e a l ( F e - C o ) p h a s e a = 0.2876 nm, c = 0.2872 n m a n d
c/a = 0.998 a n d for t h e a 2 (Ni-A1) phase, a = 0.2870 nm, c = 0.2872 n m a n d
c/a = 1.001.
H o w e v e r , f r o m t r a n s m i s s i o n e l e c t r o n m i c r o s c o p y Pfeiffer (1969) a n d B r o n n e r et
al. (1967) c o n c l u d e d , in d i s a g r e e m e n t with the X - r a y diffraction results, t h a t only
the a2 ( N i - A I ) p h a s e is o r d e r e d a n d that the a l ( F e - C o ) p h a s e is a d i s o r d e r e d
solid solution.
F r o m m e a s u r e m e n t s of t h e m a g n e t i c p r o p e r t i e s of alnico 5 S e r g e y e v a n d B u l y g i n a
(1970) a n d B u l y g i n a a n d S e r g e y e v (1969) c o n c l u d e d t h a t t h e v o l u m e fraction, p, of
the c~1 ( F e - C o ) p h a s e a r e in the r a n g e 0.62-0.67, d e p e n d i n g on t h e t h e r m o m a g -
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 137

netic heat treatment, and hence are in excellent agreement with those observed
directly from the electron microscopical studies.

2.Z Alnico 6

This is a high coercivity form of random grain alnico 5. The higher coercivity is
due to a Ti or Nb addition but is achieved, unfortunately, at the expense of a
lower energy product (see table 2). A typical demagnetization curve for alnico 6 is
shown in fig. 17.

1"2

Alnicos 6,7,8 I-
d
0"8

0"4

0
-12 -8 - 4 0
H (~104Am - 1 )
Fig. 17. Demagnetization curves for anisotropic alnicos 6, 7 and 8.

3. Anisotropic alnicos 8 and 9

3.1. Thermomagnetic treatment of anisotropic alnico 8

The high energy products of alnico 5-7 (alnico 5DG, alnico 7) are achieved
principally as a result of" the increased remanence parallel to the 'thermomagnetic
field' direction and the preferred or columnar crystal axis. The coercivity on the
other hand remains almost the same as that in alnico 5 with randomly oriented
grains. It can be seen from fig. 11 that the various forms of alnico 5 all have nearly
rectangular demagnetization curves so that the alloys are used at m a x i m u m
efficiency only if the load line or load permeance B / H intersects the demag-
netization curve at a point corresponding to the (BH)maxenergy product. A slight
shift of the load permeance results in a significant decrease in the available
energy. Since the composition of the alloy limits the m a x i m u m remanence to
about 1.4T, further squaring of the demagnetization curve will not result in a
significant increase in the m a x i m u m energy product. In view of this limitation
considerable research effort has been expended to increase the coercivity, and this
has been achieved by increasing the cobalt content to 32-36wt % and adding
138 R.A. McCURRIE

4-5 wt % titanium. The addition of Ti alone was found to increase the coercivity but
this was, unfortunately, accompanied by a decrease in the remanence (Honda 1934).
To compensate for the latter Koch et al. (1957) found that it is necessary to
increase the cobalt content to ~ 3 5 % .
Koch et al. (1957) found that if the alnico 8 alloy containing 35.5% Fe, 34% Co,
14.5% Ni, 7% A1, 5% Ti, 4% Cu was cooled from 1260°C in a saturating magnetic
field and subsequently tempered at 585°C the following properties were obtained:
Br = 0.85 T, BHc = 92 kAm -1 and (BH)max--28 kJm -3. When the same alloy was
given an isothermal heat treatment at ~800°C for several minutes (precise details
not specified) in a saturating magnetic field there was a significant improvement in
the magnetic properties to: Br = 0.9 T, BH~ = 112 k A m -1 and ( B H ) ~ = 40kJm -~.
Koch et al. (1957) concluded that for Ti-containing alnicos the best magnetic
properties are obtained by isothermal heat treatment in a saturating magnetic
field (~200-300 kAm -1) followed by tempering for several hours at 585°C. De Vos
(1966, 1969) suggested that because the Ti ion has a large radius it has a low rate of
diffusion and therefore requires a longer time at high temperatures for the phase
separation to occur. From electron microscopy Koch et al. (1959) showed that for
alnico 5 which had been isothermally heat treated in a magnetic field the a l (Fe-Co)
particles were much more elongated than those which had been given the
conventional thermomagnetic treatment (i.e. cooling in a magnetic field). Koch et al.
(1959) also found that even higher coercivities could be obtained with a cobalt
content >34% Co and a Ti content > 5 % Ti. For alnico 8 alloys with 34-40% Co and
5-8% Ti, Wyrwich (1963), Stfiblein (1963), Planchard et al. (1964b), Bronner et al.
(1966a, b) and Vallier et al, (1967), Livshitz et al. (1970a) have shown that coercivities
BHc --~ 176 kAm -1 and energy products up to ~48 kJm -3 can be obtained. Bronner et
al. (1966a, b) found that the energy product of 48 kJm -3 was constant up to about
45% Co.
Wright (1970) and Bronner et al. (1970a) have shown that the addition of
niobium as well as titanium also has beneficial effects on the magnetic properties
of alnicos 5, 6, 8 and 9, though alnico 5-7 should contain a small addition of
niobium only.
Typical compositions and magnetic properties of alnico 8 alloys are given in
table 5.
Koch et al. (1959) found that for the alloy 35.5% Fe, 34% Co, 14.5% Ni, 7%
A1, 4% Cu, 5% Ti, the homogeneous bcc ce phase is stable only above about
1250°C and that between 1250°C and 845°C it decomposed to form another bcc c~
phase and an fcc y phase with lattice parameter, a0 = 0.365nm. In the range
845-800°C, the c~ phase decomposes spinodally to form a bcc F e - C o rich al phase
and a bcc Ni-A1 rich a2 phase. A small amount of the y phase is also present at this
stage but below 800°C this transforms by an apparently diffusionless reaction to
another bcc phase, c~v with lattice parameter 0.359 nm.
Ritzow (1963) suggested that the difficulties encountered in controlling the heat
treatment of the high Ti alloys is probably due to the fact that the Curie
temperature is practically in the y region (see section 9) so that the isothermal
thermomagnetic treatment temperature is very critical. Julien and Jones (1965a, b)
S T R U C F U R E A N D P R O P E R T I E S OF A L N I C O P E R M A N E N T M A G N E T A L L O Y S 139

TABLE 5
High coercivity field-treated anisotropic alnico 8 alloys.

Composition
B, BHc (BH)max
Fe Ni A1 Co Cu Ti Nb (T) (kAm -l) (Jm -3)

40.8 14.6 6.9 28.5 3 4.2 2 0.95 97 40

35 14.9 7 34 4.5 5 - 0.96 103 41

34.5 14.3 6,9 34 3.8 5.5 1 0,88 117 41

36 14 7,5 34 3 5 0.5 0.93 119 46


/
35 13 7.5 34 3 6 1.5 0.83 ' 150 48

29.5 14 7.5 38 3 8 - 0.74 167 48

35.5 14 7.5 34 3 4.5 1.5 0.8 162 48

33 15 7.5 34 3 7 0.5 0.8 154 49


Hf

N.B. 1 T = 10 4 G ; 1 A m -~ = (4:7/1000) Oe; 1 Jm -3 = 40~ GOe.

have shown that if an alloy of alnico 8 with 32% Co and 6.5% Ti is held at 900°C
for about 15 rain the 3' phase is produced which lowers the (BH)max product by
50% compared with the same alloy which had not been annealed at 900°C, They
also found that the tendency to precipitate the 3' phase was increased when the
Cu content was increased (a similar effect was observed in alnico 5) while the Ti
suppressed its formation. The homogenization temperature was just below the
melting point. If alnico 8 alloys are cooled too slowly the appearance of the 3'
phase results in a deterioration in magnetic properties which manifests itself in a
concave demagnetization curve just as occurs in alnico 5 (see fig. 18). Such
concave demagnetization curves can be synthesised from three normal demag-
netization curves. The effects of the a~ phase on the hysteresis loops have also
been discussed by Julien and Jones (1965b). In view of the tendency of Ti-
containing alnico 8 alloys to form the fcc Y phase their heat treatment must be
Carefully controlled in order to obtain optimum magnetic properties. This inevit-
ably increases the cost of their p r o d u c t i o n - though this is largely due to the high
cobalt content. The kinetics of the a ~ 3, transformation in alnico 8 (and alnico 5)
alloys have also been investigated by Planchard et al. (1964a, 1965, 1966a). A study of
the influence of cobalt on the c~ ~ 3, transformation has been made by Marcon et al.
(1971) who showed that for alnicos containing more than 28% Co it is practically
impossible to avoid the precipitation of the deleterious y phase during thermomag-
netic treatment unless appropriate amounts of a-stabilizing elements such as silicon
or titanium are added.
A typical heat treatment for alnico 8 alloys is as follows. After controlled
140 R.A. McCURRIE

1"2

Alnico 5 1.0

0.8 ~ -
m
0.6

O
"O
0"4 x
_=
I.I.

0"2

t 0
60 50 40 30 20 10 0
Applied field,H (kAm-1.)

Fig. 18. Demagnetization curve for alnico 5 showing deleterious effect of the presence of the o~ fcc
phase on the shape of the curve.

(BH) max (k Jrn-3)


80 60 40
1-4
// i
¢I 1.2
/I ~ /I
, 1I
1"0 ~ "
ii m
/ i/./ // //
0"8
¢n
c
0.6 ¢~
qD
X
0.4

0.2

! I
160 140 120 100 80 60 40 20 0
Applied field, H CkAm-1)
Fig. 19. Demagnetization curve for alnicos 8, 9 and 5-7. The latter two curves are given for
comparison. The alloy 8(b), which has the highest coercivity is also known as sermalloy A1 (Bronner et
al. 1966b).

c o o l i n g f r o m a b o v e 1250°C, t h e a l l o y is h e a t t r e a t e d f o r a f e w m i n u t e s at a b o u t
820°C in a s a t u r a t i n g m a g n e t i c field ( ~ 3 0 0 k A m -1) a n d t e m p e r e d f o r a b o u t 6 h at
650°C f o l l o w e d by a b o u t 24 h at 550°C.
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 141

A typical B - H demagnetization curve for alnico 8 is shown in fig. 19, the


compositions and magnetic properties of various alnico 8 alloys are given in table
5.

3.2. Extra high coercivity alnico 8

A higher coercivity field-treated random grain form of alnico 8 can be produced


by increasing the cobalt and titanium contents still further to 37-40% Co and
7-8% Ti. Sometimes a little (0.5 to 1.5%) of the titanium is replaced by niobium.
The maximum coercivity of these extra high coercivity alnico 8 alloys is about
180 kAm -1 which is higher than that of any other alnico. The production and
properties of extra high coercivity alnico 8 alloys have been described and
investigated by Wyrwich (1963), Planchard et al. (1964b, 1966b), Bronner et al.
(1966a, b), Vallier et al. (1967) and Bronner et al. (1970a). Planchard et al. (1964b)
and Bronner et al. (1966a) studied the dependence of the magnetic properties on the
Ni, Ti, A1 and Cu content for alloys containing 40% Co. Planchard et al. (1964b)
obtained a maximum coercivity BHc = 161 kAm -1 with Br = 0.71 T and (BH)m~ =
44 kJm -3, for an alloy containing 27.5% F e - 4 0 % C o - 14% N i - 8 % A 1 - 7 . 5 %
T i - 3 % Cu. The alloy was homogenized at 1250°C for 1 h and then heat treated
isothermally for 5 minutes at 810°C in a magnetic field of 336 kAm -1. After this
treatment the alloy was annealed for 6 h at 650°C and then for 24 h at 550°C.
In a later investigation Bronner et al. (1966b) developed an alloy containing
29.5% F e - 3 8 % C o - 1 4 % N i - 7 . 5 % A 1 - 8 % T i - 3 % Cu with a coercivity
BHo = 168 kAm -I with Br = 0.74 T and (BH)max = 48 kJm -3. These properties were
obtained by homogenization at 1250°C in a neutral or reducing atmosphere
followed by cooling in compressed air to 600°C in order to avoid the precipitation
of the deleterious 3, phase. The alloy was then isothermally annealed for a few
minutes at 820°C in a magnetic field ~336 kAm 1 and then tempered for 6 h at
640°C and finally for 24 h at 550°C. This alloy, which is an extra high coercivity
alnico 8, is also known as sermalloy A1. Apart from its high coercivity and energy
product, sermalloy A1 has high useful recoil energy (a maximum Erec ~ 20 kJm -3)
and a high Curie temPerature Tc = 874°C so that it is very stable under demag-
netizing conditions and at temperatures up to about 550°C. An alloy, with almost
identical properties to sermalloy A~, also known as hycomax IV has also been
developed by Harrison and Wright (1967). A typical demagnetization curve for
this special alnico 8 alloy (A1) is shown in fig. 19.
Koch et al. (1957) investigated the effectiveness of Ti and Nb additions in
increasing the coercivity of alnico 8 alloys and concluded that Ti was more
effective. However, Bronner et al. (1970a) and Bronner (1970) found that the
addition of Nb (0.5-2.0%) to alnico 8 alloys containing 4.5-6.5% Ti could be
beneficial. For example the alloy 35.5% F e - 34% C o - 14% N i - 7.5% A 1 - 4.5%
T i - 3 % C u - 1 . 5 % Nb B H c = 1 6 2 k A m -1, B r = 0 . 8 T and (BH)max=48kJm -3.
Bronner (1970) reported that the addition of hafnium was also beneficial and
found that the alloy 33% F e - 3 4 % C o - 15% N i - 7.5% A 1 - 7% T i - 3% Cu with
0.5% Hf had a coercivity td-/c = 154 k A m -1, a Br = 0.8 T and a (BH)m~ = 49 kJm -3.
142 R.A. McCURRIE

Similar investigations of extra high coercivity alnico 8 alloys have been made by
Livshitz et al. (1970a) who used a wide variety of thermomagnetic and tempering
treatments. They obtained magnetic properties in the ranges Hc = 160-176 kAm -1,
Br = 0.65-0.75 T and (BH)m~x = 36-44 kJm -3 for alloys containing 38% Co and
8.0-8.5% Ti after optimum thermomagnetic and tempering treatments. The
compositions and magnetic properties of some typical extra high coercivity alnico
8 alloys are given in table 5.
A simplified heat treatment, using continuous cooling, rather than isothermal
heat treatment in a magnetic field (240 kAm -1) for alnico 8 alloys has been
developed by Wright (1970). He showed that by cooling alloys with compositions
in the range 34-35% Co, 14-15% Ni, 6.8-7.2% A1, 5-5.4% Ti, 0.8-1.1% Nb,
3 - 4 % Cu, balance Fe from 1250°C (at average rates of 2, 1.2, and 0.6°Cs -I from
1200-600°C) in a magnetic field of 240 kAm 1 followed by tempering for 4 h at
640°C and then for 16h at 570°C, magnetic properties better than Br = 0.85 T
BHc = l l 2 k A m -~ and (BH)max = 36kJm -3 can be obtained. Wright (1970) also
showed that if 0.25% S is added to the alloys, magnets with columnar crystals can
be produced by casting in a mould with heated sides and chilled at the end faces.
For these columnar magnets typical properties are, Br = 1.03 T, BHc = 118 kAm -~,
and (BH)max = 58 kJm -3. Columnar alnico 8 alloys i.e. alnico 9 alloys are discussed
in section 3.3.

3.3. Anisotropic alnico 9 with fully columnar grains

Alnico 9 is produced by the columnar crystallization of alnico 8. Unfortunately,


the high Ti content of the latter reduces the grain size (Luteijn and De Vos 1956)
so that columnar crystallization of alnico 8 is difficult. Luteijn and De Vos (1956)
succeeded in producing a grain oriented alnico 9 magnet by using special
techniques. For the alloy containing 35% Fe, 34% Co, 15% Ni, 7% A1, 5% Ti and
4% Cu by weight, they obtained a magnet with Br = 1.18 T, BHc =- 105 kAm -1 and
(BH)max = 88 kJm -3. These very high values were obtained by using very pure
starting materials and isothermal heat treatment parallel to the [100] axis
of the columnar grains. The demagnetization curves for the alloy are shown in
fig. 19.
Gould (1964), Makino and Kimura (1965), Wittig (1966) and Harrison and Wright
(1967) have shown that these difficulties in producing fully columnar alnico 9 magnets
can be overcome by the addition of small quantities of S, Se or Te.
Alnico 9 can be prepared as follows. After melting an alnico 8 alloy with a small
addition of S, Se or Te (sometimes two of these elements are added) the fully
columnar structure is obtained by casting the alloy in a heated or exothermic
mould in which the end faces are chilled so that grains with (100) axes grow
perpendicularly to the chilled surfaces. The alloy is then solution treated at
1250°C cooled at a controlled rate and heat treated in a saturating magnetic field
(~300 kAm -1) for a few minutes at about 820°C and finally given a two-stage
tempering treatment for several hours at 650°C and at 550°C. The grain structure
of a fully columnar alnico 9 magnet is shown in fig. 20.
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 143

Fig. 20. Grain structure of fully columnar alnico 9 (Gould 1964) (magnification: ~2x).

The production of columnar alnico 9 has also been described and investigated
by Fahlenbrach and Stfiblein (1964), Naastepad (1966), Harrison (1966), Hoffmann
and Stfiblein (1966, 1967, 1970), Palmer and Shaw (1969), Dean and Mason (1969),
Hoffmann and Pant (1970) and Pant (1974). Hoffmann and Pant (1970) produced
magnets (containing 7.7% Ti) with coercivities up to a m a x i m u m of Uric =
1 7 1 k A m -1, iHc = 1 7 9 k A m -l, B r = 0 . 8 7 T and (BH)max = 7 7 k J m -3. They also
produced a magnet (containing 7.1% Ti) with an energy product (BH)max =
98 kJm -3 in combination with ~Hc = 136 k A m -1, i S c ~ 138 k A m -~ and Br = 1.03 T.
Pant (1974) produced alnico 9 magnets having diameters ~ 1 5 - 8 0 m m with a
columnar crystallization length of ~120 m m for which (BH)max ~ 90 kJm -3 (with
BHo = 150 k A m -1) and (BH)max ~ 80 k J m -3 (with BHc ~ 160 kAm-1).
Although the coercivity of alnico 9 is nearly the same as that of alnico 8 the
higher energy product is due to the increased remanence parallel to the columnar
axis i.e. [100]. The improved properties of alnico 9 parallel to the columnar axis
are of course achieved at the expense of those perpendicular to the columnar axis
but in many applications this is not a serious disadvantage.
Demagnetization curves for alnicos, 8 and 9 are shown in fig. 19 from which it
can be seen that the magnetic properties of alnico 9 are considerably better than
those for alnico 8; the m a x i m u m energy products for alnico 9 are typically in the
range 60-75 kJm -3. The higher energy products of alnicos 8 and 9 compared with
alnico 5-7 are, unfortunately, obtained at the expense of a reduced remanence.
T h e relationship between the magnetic properties and the crystal textures of
alnico 9 alloys have been investigated by Higuchi and Miyamoto (1970) and
D u r a n d - C h a r r e et al. (1978). T h e latter used Schulz's (1949) X-ray diffraction method
and found that the observed metallurgical texture could be correlated with the
solidification rates.
Grain growth by solid state recrystallization in various alnico 9 alloys has been
144 R.A. M c C U R R I E

investigated by Wright and Ogden (1964) but although some large crystals were
grown the technique was not considered to be very successful. The dependence of
the magnetic properties of columnar alnico 9 on the angle between the columnar
axis and the direction of thermomagnetic treatment, and the angle of measure-
ment is discussed in section 7.3.
The development and chronology of alnico magnets are shown in fig. 21 (Cronk
1966). The best magnetic properties are of course obtained for single crystal
alnicos. Naastepad (1966) showed that a single crystal containing 35 wt % Fe,
34.8% Co, 14.9% Ni, 7.5% A1, 5.4% Ti and 2.4% Cu had a coercivity 8He of
122kAm -1 and an energy product (BH)max of 107kJm -3. The latter energy
product is the highest yet reported for any alnico. A summary of the magnetic
properties and compositions of columnar alnico 9 alloys is given in table 6.

120
Alnicos MC9
100 DG -- Directed grain
¢0 MC - Monocrystal
I
E MC5
-~ 80

5-7
x
60
E 5DG
.,p
m 40

20 2

o II I I ! I

1930 1940 1950 1960 1970


Calendar year
Fig. 21. Development and chronology of alnico magnets (after Cronk 1966),

TABLE 6
Compositions and magnetic properties of anisotropic field-treated alnico 9 columnar alloys.

Composition
Br BEe (BH)max
Fe Ni A1 Co Cu Ti Nb Others (T) (kAm -1) (kJm -3)

26.1 14.7 6.8 40.3 2.9 8,2 - 0.77 Te 0.895 160 67


O.22 S
35.2 14.8 7 33.1 4.5 5.5 - 0.22 S 1.095 123 83.5
35 15 7 34 4 5 - - 1,18 104.5 87.5
36.1 14.5 7 34 3 5.2 0,2S 1.11 129 91.5
29 14 7 38.9 3,5 7.1 - 0.45 S 1.03 136 97
0.05 C
35 14.9 7.5 34.8 2.4 5.4 - - 1.15 121 106

N.B. 1 T = 10 4 G; 1 A m -1 = (4~'/1000) Oe; 1 Jm -3 = 40~r GOe.


STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 145

3.4. Shape anisotropy of alnicos 8 and 9

The effects of varying the cobalt and titanium contents on the induced shape
anisotropy constant K. of alnico alloys have been studied by Takeuchi and Iwama
(1976) (see also section 2.4). They found that for an alnico 8 alloy (29.5%
F e - 39% C o - 14% N i - 7% A1- 7.5% T i - 3% Cu) Ku = 2.4 × 105 Jm -3 and a
magnetocrystalline anisotropy constant K1 = 0.26 x 105 Jm -3, thus confirming that
the anisotropy and hence the coercivity of alnico alloys is due predominantly to
the shape anisotropy of the individual Fe-Co rich particles. The induced aniso-
tropy constant K. was of course measured on single crystal specimens in which the
individual particles were almost fully aligned by the thermomagnetic treatment.

TABLE 7
Compositions of single crystal alnicos.

Specimen
No. Fe Ni Al Co Cu Ti

1 51.4 14 7.7 23.9 3 0

2 47.9 14.4 7.9 26 2.5 1.3

3 43.5 14 7.6 29 2.9 3

4 29.5 14 7 39 3 7.5

TABLE 8
Apparent anisotropy constants Ku and K,] ( x 104 J m -3) determined by torque measurements, volume
fraction, p, particle diameter d, elongation l/d intrinsic coercivity iHc and flux coercivity BHc (K" = Ku/p)
(Takeuchi and Iwama 1976).

Specimen Ti Ku d inc BH¢ K"


No. (wt % ) (104Jm -3) p (nm) lid (kAm-1) (kAm-1) ( 1 0 4 j m 3)

1 0 15.6 0.68 44 4-5 57 - 22.9

2 1.3 16.1 0.67 42 6-7 59 - 24

3 3.0 17.3 0.63 37 8-12 67 - 27.5

4 7.5 24.2 0.46 ~ 30 30-50 167 - 52.6

Alnico 5* 0.0 15.2 0.67 40 4-6 - 53 22.7

Alnico 8* 5.0 23 0.54 30 13-16 - 114 42.6

Alnico 8"* 7.5 - - 20 30-35 - 147 -

*Sergeyev and Bulygina (1970) and **Granovsky et al. (1967).


N.B. 1 T = 104 G ; 1 A m -1 = (47r/1000) O e ; 1 J m -3 = 40~" G O e .
146 R.A. McCURRIE

This is clearly shown by the electron micrographs obtained by Takeuchi and


Iwama (1976). Their value of Ku = 2.4 x 105 Jm -3 is in very good agreement with
that obtained by Sergeyev and Bulygina (1970) who found a Ku = 2.3 x 105 Jm -3
for an alnico 8 alloy containing 26.5% F e - 40% C o - 1 4 % N i - 7.5% A 1 - 7.5%
T i - 4 . 5 % Cu. The alloy was homogenized at 1240°C, cooled to 800°C at 5°Cs -1
and then annealed for about 12 rain at 820°C in a magnetic field of 160 kAm -1
after which the alloy was tempered for 5 h at 650°C and 20 h at 560°C.
The compositions and properties of the alloys used in the above investigations
are summarized in tables 7 and 8.
The high coercivities of alnico 8 (and alnico 9) can be attributed to the
increased shape anisotropy energy resulting from the high cobalt content and the
addition of titanium. The effects of cobalt and titanium contents on the magnetic
properties are discussed in sections 8 and 9 respectively.

3.5. Microstructures of.alnicos 8 and 9

Replication electron micrographs of the highly oriented microstructure of an


alnico 8 alloy were obtained by De Vos (1966, 1969). Typical micro-
structures in planes parallel and perpendicular to the thermomagnetic field
direction are shown in figs. 22(a) and (b). The regular distribution of the
precipitated particles in both orientations is characteristic of a system in which
phase separation has occurred by spinodal decomposition. From measurements on
the electron micrographs the average dimensional length to diameter ratio 1/d is
>10 and from the application of quantitative metallography (Hilliard 1962, 1967,
1968, Hilliard and Cahn 1961, Underwood 1970, 1973, Saltykov 1970) the volume
fraction of the particles is ~0.65. Although the volume fraction of the particles is
comparable to that of the other alnicos there is a marked increase in the
dimensional ratio, degree of particle alignment and particle perfection; all of these
factors contribute to the increased coercivity of the alnico 8 alloys compared with
alnico 5 alloys. Since alnico 9 alloys have similar compositions to the alnico 8
alloys (alnico 9 is simply a columnar crystallized alnico 8) it may be assumed that
the microstructures are similar to those of alnico 8. The microstructures of alnico
8 alloys have also been investigated by Bronner et al. (1967), Granovsky et al.
(1967), Pfeiffer (1969), Mason et al. (1970), Livshitz et al. (1970b), Iwama et al.
(1970) and Takeuchi and Iwama (1976). From the electron micrographs obtained
by the above authors it can be seen that the F e - C o rich particles have length to
diameter ratios l/d in the range 10-50 where l ~ 400-1000 nm and d is in the
range 20-40 nm. The application of quantitative metallography to the electron
micrographs shows that the volume fraction of the F e - C o rich particles is ~0.65.
Although the volume fraction of the particles in alnico 8 is comparable to that in
the other alnicos, there is a marked increase in the elongation (l/d), the degree of
particle alignment and particle perfection; all of these factors contribute to the
increased coercivity of alnico 8 alloys compared with alnico 5 alloys.
By using X-ray electron diffraction Pashkov et al. (1969) measured the lattice
parameters of the bcc a~ and a2 phases in three alnico 8 alloys containing
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 147

Fig. 22. Electron micrograpti of the al + ~2 structure of alnico 8 after an isothermal heat treatment at
800°C for 9 min; (a) in a plane parallel to the direction of the applied field during heat treatment, (b) in
a plane perpendicular to the field direction (magnification 45500 ×) (after De Vos 1966).

TABLE 9
Lattice parameters of the bcc oq and c~2 phases in alnico 8 alloys (A).

c~1 phase (Fe-Co) c~2 phase (Ni-A1)

Alnico 8 a c c/a a c c/a

35% Co, 5% Ti 2.909 2.873 0.99 2.855 2.873 1.006

40% Co, 7% Ti 2.909 2.872 0.987 2.850 2.872 1.008

42% Co, 8% Ti 2.915 2.872 0.985 2.852 2.872 1.007


148 R.A. McCURRIE

respectively, 35, 40, 42% Co and 5, 7, and 8% Ti; their results are shown in table 9,
from which it can be seen that both the eel and O~2 phases are very slightly
tetragonal with c/a < 1 and c/a > 1 respectively.

4. M6ssbauer spectroscopy of alnicos 5 and 8

From M6ssbauer spectroscopy measurements on alnico 5, Shtrikman and Treves


(1966) have concluded that in the optimum permanent magnet state, the alloy
consists of two phases one of which is an Fe-rich ferromagnetic phase and the
other is a Ni and Cu rich paramagnetic phase. M6ssbauer measurements on
alnico 8 made by Albanese et al. (1970) show that 7-8% of the total Fe content is
paramagnetic. From the quadrupole splitting of the M6ssbauer spectra it was also
concluded that the two phases have a small tetragonal distortion in agreement
with the X-ray work of Bulygina and Sergeyev (1969). Van Wieringen and Rensen
(1966) also detected the presence of Fe in both the al and Og2 phases. They have
also shown that after continuous cooling of alnico 5 all the Fe atoms have a
ferromagnetic environment suggesting that both the c~1 and O~2 phases are ferro-
magnetic, whereas after tempering (Ho = 50 kAm -1) 5% of the total iron content is
in a non-ferromagnetic phase. The M6ssbauer spectra of alnico 5 and alnico 8,
particularly the latter, have also been investigated in detail by Makarov et al.
(1967), Belova et al. (1969), Povitsky et al. (1970), Belozersky et al. (1971) and
Makarov et al. (1972). A short review of M6ssbauer measurements on the alnicos
has been published by Schwartz (1976).

5. Sintered alnicos

The alnico alloys whether isotropic or anisotropic can also be made by mixing
suitably fine metal powders, pressing to a green compact followed by a sintering
heat treatment at about 1250-1350°C to produce a homogeneous solid with low
porosity. The solid compact is then given a heat treatment appropriate to the
particular composition. The resulting magnetic properties of the sintered alloys
are comparable, though slightly inferior, to those of the cast alloys. The energy
product for anisotropic sintered magnets may sometimes be as much as 20%
lower than that for the cast magnets but for isotropic cast magnets the difference
is usually smaller than this. A comparison of the magnetic properties of various
sintered and cast alnicos has been given by Bronner et al. (1970b). Typical
demagnetization curves for four sintered alnicos are shown in fig. 23. The
advantage of the sintering process is that very small magnets of intricate shape can
be made which is not possible or very expensive by the usual casting and grinding
process. However, sintered alnicos have a very small share of the market.
The production and magnetic properties of sintered alnicos have been discussed
by Schiller (1968), Schiller and Brinkmann (1970) and Heck (1974).
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 149

- 1-2

Sinteredalnicos/~~ 1.0

6 0.8~
0.6
(D
"0

D'4
I,I.

D.2
n 0
70 60 50 40 30 20 10 0
Applied field, H (kAm-1)
Fig. 23. Typical B-H demagnetization curves for sintered alnicos 2, 5, 6 and 8.

6. Moulded, pressed or bonded alnico m a g n e t s

Alnico alloys can be milled to fine powders without a marked deterioration in


their magnetic properties. If the powders are then bonded into a thermosetting
resin they can be accurately pressed or moulded to the required magnet shape.
The properties of the bonded magnets are inferior (Dehler 1942, Heck 1974) to
those of the cast or sintered magnets. This is due to the fact that the volume
fraction of the powder cannot exceed about 60% so that there is a corresponding
reduction in the remanence.

7. Effects of t h e r m o m a g n e t i c t r e a t m e n t on the magnetic properties of alnicos 5 - 9

7.1. Factors controlling development of Ol 1 particle shape anisotropy

The factors which control the development of the magnetic anisotropy of alnico 5
during thermomagnetic treatment have been studied in detail by Zijlstra (1960-
1962) following a suggestion made by N6el (1947b). H e found that the rate of
elongation of the particles is related to the difference between the decrease in the
magnetic free energy of the particles in the magnetic field and the simultaneous
increase in the interfacial or surface free energy. The equilibrium value of the
elongation is high when the magnetic free energy is large compared with the
interfacial free energy between the particles and the matrix. Since the former is
measured as energy per unit volume and thus independent of particle size and the
150 R.A. McCURRIE

latter is a surface energy and hence inversely proportional to the particle size, the
particles will always become considerably elongated when they are sufficiently
large. The efficacy of the thermomagnetic treatment on alnico 5 can be attributed
to the particularly low value of the interfacial energy or surface free e n e r g y -
typically this is ~10-3Jm -2 (1 erg cm-2). This enables the particles to become
elongated when they are still small enough to display single domain behaviour and
thus to form a magnet with high shape anisotropy together with a high coercivity.
Zijlstra (1961) suggested that the same reasoning leads to the conclusion that the
Mishima alloy Fe2NiAI, which is usually not considered to respond to magnetic
annealing owing to its much larger interfacial free energy, may be expected to do
so only when the particles are large enough. He verified this conclusion by heat
treating a specimen of Fe2NiA1 in a magnetic field for two weeks at 725°C during
which it developed a uniaxial magnetic anisotropy of 4.5 x 10 4 Jm 3 which is of the
same order of magnitude as that observed in alnico 5. This Fe2NiA1 alloy did not,
however, have a high coercivity because after the long thermomagnetic treatment
the particles were too large to be single domains. Thus it appears that the function
of the cobalt in the alnico alloys, apart from increasing the saturation magnetic
polarization and the Curie temperature, is to decrease the interfacial free energy
between the Fe-Co rich particles and the Ni-A1 rich matrix and hence to enable
the alloys to respond to thermomagnetic treatment with a resulting improvement
in their magnetic properties.
If the magnetic field is applied only during the cooling over the narrow
temperature range 840-790°C with subsequent tempering at 600°C without an
applied field, the resulting magnetic properties are only slightly inferior to alloys
which have been given the full thermomagnetic treatment. If the field is applied
only below 790°C the anisotropy is less than half that of the fully thermomag-
netically treated specimens. If the specimen is cooled to 790°C in the presence of
the field and then allowed to cool with the field off, the maximum energy product
is lower than the value obtained by quenching the alloy from 790°C. The
decrease is due presumably to a partial destruction of the shape anisotropy by
inhomogeneous demagnetizing fields and thermal fluctuations. The most effective
temperature range for the particle elongation and alignment in a magnetic field is
therefore 840°C to 790°C; the lower limit is particularly critical.
Nesbitt and Williams (1957) suggested that in the temperature range 850-790°C
only nucleation takes place, while particle growth occurs at 600°C. However,
Zijlstra's (1960-1962) and De Vos' (1966, 1969) experiments showed that in the
field cooling treatment the final particle shape is attained in the very first stage
(i.e. in the range 850-790°C) and that subsequent heat treatments did not alter the
shape or size of the particles. As was mentioned earlier tempering at 600°C
increases the difference in the saturation magnetic polarization between the c~1
(Fe-Co rich) and c~2 (Ni-A1 rich) phases and leads to an increase in the coercivity
and remanence.
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 151

7.2. Relationship between the preferred or easy direction of magnetization and the
direction of the applied field during thermomagnetic treatment

The relationship between the orientation of the particles and the direction of the
magnetic field during heat treatment has been investigated by Heidenreich and
Nesbitt (1952). They found that the orientation of the particles is not strictly
bound to the (100) directions, but could be forced into other directions by the
thermomagnetic treatment. In a later paper Nesbitt and Heidenreich (1952a)
deduced from magnetic torque measurements that the magnetic anisotropy of
alnico 5 was a maximum when the magnetic field was directed along a (100)
direction of a single crystal sample during the thermomagnetic treatment. If the
field was applied in another direction the anisotropy was lower and the preferred
direction of magnetization appeared to lie between the field direction and the
nearest (100) direction. Similar results were obtained by Hoselitz and McCaig
(1951) but they found that the preferred direction of magnetization was closer to
the [100] direction than was indicated by the experiments of Heidenreich and
Nesbitt (1952). These differences may be attributable to difference in the mag-
nitudes of the applied fields which were chosen for the thermomagnetic treat-
ments. In another series of experiments McCaig (1953) found that the uniaxial
anisotropy of alnico 5 was approximately 105 Jm 3 and that for normal cooling
rates (~l.4°Cs -1) and angles less than about 45 ° between the crystal and field
directions, the easy direction of magnetization was close to the columnar axis.
However, for faster cooling rates, larger angles and less than optimum heat
treatments the easy direction of magnetization may be closer to the field direction
with a correspondingly lower value of the anisotropy coefficient.
An extensive investigation has also been made by Yermolenko et al.
(1964) who found that when the thermomagnetic field was parallel to the [100]
direction the O/1 (Fe-Co) particles were elongated parallel to this direction but for
other angles between the applied field and the [100] direction, the axes of
elongation of the particles were not parallel to the field direction. They concluded
in contrast to the theory proposed by N6el (1947b) and later developed by Zijlstra
(1960-1962) that the axes of elongation are not entirely determined by the
minimization of the magnetic and interface energies of the particles in the
thermomagnetic field but that their shapes and orientations are partly determined
by their elastic energies. It should be mentioned that Yermolenko et al. (1964)
used isothermal heat treatment in a magnetic field (720 kAm 1) but this enabled
them to study the microstructures of the alloy at different stages in the formation
and growth of the particles.

7.3. Dependence of the magnetic properties of the alnicos on the direction of the
applied field during thermomagnetic treatment

The dependence of the magnetic properties of columnar and single crystal alnico 5
(and alnico 6) on grain orientation and the direction of the thermomagnetic field
have been investigated in detail by Ebeling and Burr (1953). They found that the
152 R.A. McCURRIE

best magnetic properties are obtained when the direction of the thermomagnetic
field is as close as possible to the [100] direction in the single crystals or to the
long axis of the columnar single crystals- i.e., approximately parallel to the [100]
direction. When the thermomagnetic field was applied at other angles to [100] or
to the [100] columnar axis the maximum energy product decreased according to
the cosine of the angle between the thermomagnetic field direction and the [100]
or the [100] columnar axis:

(BH)max(O)= (BH)max(0)cos 0.

The variation of (BH)max with the angle 0 for various sintered single crystals of
alnico 5 is shown in fig. 24.
From measurements of the energy products of single crystal alnico 5 (51 wt %
F e - 24% C o - 14% N i - 8% A 1 - 3% Cu) after heat treatment in a magnetic field
parallel to the directions [100], [110] and [111] Zijlstra (1956) showed that the
energy product could be expressed as a power series of the direction cosines/31,/32
and/33 measured with respect to the cube axes of the crystal. This was used to
calculate the energy product for polycrystalline alnico 5 as a function of crystal
orientation. The columnar texture of the magnet was represented by assuming
that there was a constant density of [100] axes inside a cone of revolution around
the measuring direction which enclosed an angle 2,/. By averaging the above
power series for (BH)max Zijlstra obtained an expression for (BH)max in terms of
r/. The energy product was shown to decrease almost linearly from 60.8 kJm -3 for
~7 = 0° to 40 kJm -3 for r / = 45 °. The calculated value of the energy product for a

50- 0

~3o
\cosO
"--"20

0 I I I I I
0 30 60 90
Angle 0 °
Fig. 24. Dependence of (BH)max energy product on the angle between the field applied during
thermomagnetic treatment and the [100] preferred or easy axis of magnetization (after Ebeling and
Burr 1953).
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 153

polycrystalline alnico 5 magnet (i.e. one for which ~/= 180 °) was 39.2 kJm -3 which
was in good agreement with experiment and suggests that the analysis was
essentially correct. An expression for the energy product as a function of ~/ was
also developed for magnets in which the crystals are radially oriented in a plane
perpendicular to the magnet axis. Magnets with this texture are significantly better
than those with randomly oriented grains.
The magnetic properties of equiaxed and columnar alnico 5 have been investigated
in detail by Makino et al. (1963) and Makino and Kimura (1965). They measured the
demagnetization curves parallel and perpendicular to the columnar axis (a) as a
function of the angle 0 between the columnar axis and the direction of the
'thermomagnetic field', and (b) as a function of the angle 0' (0' = 90 - 0) between the
direction of measurement perpendicular to the columnar axis and the direction of the
thermomagnetic field. The results of these measurements are shown in fig. 25 from
which it can be seen that the maximum coercivity (58 kAm-1), remanence (1.35 T)
and energy product (64 kJm -3) are obtained when the 'thermomagnetic field' is
applied parallel to the columnar axis. For the measurements in the direction
perpendicular to the columnar axis the maximum coercivity (50kAm-1),
remanence (1.32 T) and energy product (42 kJm -3) are obtained when the ther-
momagnetic field is parallel to the direction of measurement, i.e., perpendicular to
the columnar axis. Although the maximum energy product for (b) of 42 kJm -3 is
lower than that of 64 kJm -3 for (a) the fact that such a high value was obtained for
(b) is an important result because it provides further evidence for the theory that

- 1-4

~ 1 . 2
0=0¢

60 50 40 30 20 10 0
Applied field, H (kAm-1}
Fig. 25. B - H demagnetization curves measured in the direction M of the columnar axis after heat
treatment in a magnetic field at an angle 0 to the columnar axis c (after Makino and Kimura 1965 and
Makino et al. 1963).
154 R.A. McCURRIE

15 45

[010] . H ~ /

~"~'I0 ~ . , . 4 ~...Ku " ~ / 30

. ool e.ol,
0~ I 1 I I I I 0
0 15 30 45
Angle [30

Fig. 26. Variations of uniaxial anisotropy constant Ku and its direction o~ with direction/9 of the heat
treatment field H.

the preferred direction of magnetization is largely determined by the direction of


the thermomagnetic field (N6el 1947(b), Zijlstra 1960, 1961) and that crystallo-
graphic anisotropy is of secondary importance. Zijlstra (1960, 1961) has shown by
experiments on Fe2NiA1 that the time for which the alloy is subjected to the
thermomagnetic field treatment is also of fundamental importance in determining
the results. (The theories of the effects of thermomagnetic treatment of the alnicos
are discussed in section 17.)
Iwama et al. (1976) measured the anisotropy constant Ku of single crystal alnico
5 after cooling in a magnetic field at angles 0 °, 15°, 30 ° and 45 ° to the [100]
direction. The results are shown in fig. 26; Ko decreased from 15.6× 104Jm -3
parallel to the [100] direction to 4.5 x 104 Jm s at 45 ° to [100]. Figure 26 also shows
the variation of the angle a between the preferred axis of Ku and the [100]
direction as a function of the angle/3 between the direction of the heat treatment
field and the [100] direction. It can be seen that only when /3 = 45 ° does the
preferred direction of magnetization (i.e. the Ku axis) follow that of the heat
treatment field, i.e., a 2 4 5 °. At lower values of /3 the Ku axis lies between the
direction of the heat treatment field and the [100] direction. These results are in
good agreement with those obtained by McCaig (1953).

8. Effects of cobalt on the magnetic properties of the alnicos

The addition of cobalt to both the isotropic and anisotropic alnicos has several
beneficial effects. In the isotropic alloys it increases the saturation magnetic
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 155

polarization Js and hence the remanence Jr. Since the cobalt is precipitated in the
c~1 phase particles (i.e. the F e - C o rich phase) this increases the difference between
the magnetic polarizations of the al and a2 phases and hence increases the
coercivity due to the resulting increase in the shape anisotropy. The same
improvements are also observed in the anisotropic alnicos 5-9 but in these alloys
the cobalt content is in the range 23-40%. The high cobalt content also increases
the Curie temperature Tc which has the beneficial effect of increasing the
sensitivity (of the alnicos 5-9) to magnetic annealing. This is due to the fact that
the spinodal decomposition temperature Ts is then further below Tc so that the
magnetic polarization of the particles at Tc is higher, and therefore increases the
elongation of the particles in the magnetic field direction because their magnetic
free energy is thereby lowered. Cahn (1963) has shown the magnetic energy is
proportional to the square of the rate of change of the saturation magnetic
polarization with composition i.e. (OJs/Oc)2 and that the latter is very large when
the spinodal decomposition temperature is close to the Curie temperature. Thus
heat treatment in a magnetic field is expected to be most effective when T~ is close
to but lower than To. According to Zijlstra (1960, 1961) the high cobalt content
also decreases the interfacial energy ,,/between the F e - C o rich c~i phase and the
Ni-A1 rich a2 phase so that the increase in the total surface energy which results
from the particle elongation is reduced. N6el (1947b) and Zijlstra (1960, 1961)
have suggested that the sensitivity of alnico alloys is proportional to (A J)2~3,
where AJ is the difference between the saturation magnetic polarizations of the
F e - C o rich c~ phase and the Ni-A1 rich a2 phase. Thus when AJ is high and 3'
low the rate of elongation of the particles in the thermomagnetic field is high. This
results in an increase in the coercivity, the remanence and the energy product
parallel to the direction of the applied field. From tables 2, 3, 5, and 6 it can be
seen that there is a substantial improvement in BHc, Br and (BH),nax in the cobalt
containing alloys compared with the original F e - N i - A I alloy (table t).
Unfortunately the increased cobalt content, particularly in the anisotropic
alnicos 5-9, favours the precipitation of the magnetically deleterious fcc phase so
that more carefully controlled cooling through the range 1200°C to 850°C is
required.
Detailed investigations of the effects of cobalt on the magnetic properties of the
alnicos have been made by many authors, see e.g., Betteridge (1939), Jellinghaus
(1943), Zumbusch (1942a, b), Bronner et al. (1966a-1970). According to Wyrwich
(1963) the best magnetic properties are obtained with a cobalt content -~38%.

9. Effects of titanium on the magnetic properties of the alnicos (mainly 6, 8 and 9)

The effects of titanium additions on the magnetic properties of the alnico 8 are
summarized in fig. 27 from which it can be seen that titanium increases the
coercivity and the energy product and the time required to form the undesirable
fcc 3, phase at 1050°C, though there is unfortunately a rapid decrease in the
remanence with increasing Ti content (Vallier et al. 1967). Similar results were
obtained by Iwama et al. (1970) who also showed that the Curie temperature of
156 R.A. M c C U R R I E

1 " 1 -- m oO. 120


O

1"0-- 4~,16 ~"


3C ® E

0-9 - ~ 3 E 12 0
Br
0

toO'8 o =
2 8
C

E
[~'mlc c
0"7- 1 ®
E
I--
I

0 . 6 [- 0 [0
2 4 6 8
Wt % Ti

Fig. 27. Effects of titanium additions on the magnetic properties of alnico 8 (41% F e - 3 4 % C o - 15%
N i - 7% A I - 3% Cu) (after Vallier et al. 1967).

the Ni-A1 rich OL2 phase decreased linearly with increasing Ti content as shown in
fig. 28 from which it can be seen that if the alnico 8 alloy containing more than
5% Ti is annealed at 600°C the Curie temperature falls below room temperature.
This increases the coercivity because in normal use (~20°C) the F e - C o rich

6 0 0 I- . o
~,AnneaEed for 48hr at T C

I "~

0
"~_

0.200 I I I I
0 2 4 6 8
Wt % titanium
Fig. 28. Variations of Curie temperature of the Ni-AI rich o~2 phase with Ti content for alnico
specimens annealed at various temperatures (after Iwama et al. 1970).
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 157

particles are then surrounded by a non-ferromagnetic matrix and are not there-
fore in exchange contact so that the difference between the saturation magnetic
polarizations of the F e - C o rich al phase and the Ni-AI rich ~2 phase is as large as
possible; the latter therefore increases the shape anisotropy energy of the particles
and hence the coercivity.
The effects of Ti additions on the magnetic properties of various thermomag-
netically treated alnico single crystals have been investigated by Takeuchi and
Iwama (1976). They measured the anisotropy coefficient Ku, the volume fraction of
the F e - C o rich particles, p, the particle diameter d, the elongation l/d and the
intrinsic coercivity iHc. Their results are summarized in tables 7 and 8. The latter
includes results obtained by Sergeyev and Bulygina (1970) and Granovsky et al.
(1967) for comparison. These results show that the anisotropy Ku (and K ' , the
anisotropy per unit volume of the F e - C o rich particles), the particle elongation
l/d and the coercivity increase as the Ti content increases. The increase in
coercivity is also accompanied by a decrease in the volume fraction p and the
particle diameter d. If we can assume that the coercivity dependence on the
packing fraction is given by,

He(p) = He(0)(1 - p ) ,

an increase in Hc is to be expected if p decreases. The increase in the coercivity


due to the titanium addition can also be partly attributed to an increase in the
elastic and surface energy associated with the spinodal decomposition which
increases the smoothness, perfection and regularity of the particle s p a c i n g - s e e
e.g. electron micrographs of alnico 8 (fig. 22(a), (b)) obtained by De Vos (1966, 1969),
and those obtained by Granovsky et al. (1967) and Takeuchi and Iwama (1976).
While Ti inhibits the a ~ y transformation Vallier et al. (1967) have shown that
it restricts the temperature ranges for homogenization and thermomagnetic
treatment so th~it the choice of the heat treatment schedule is critical. De Vos
(1966, 1969) has suggested that the necessity to give titanium containing alnicos an
isothermal anneal in a magnetic field is due to the fact that the Ti 2+ ion has a large
radius and a low rate of diffusion so that longer times are required for the phase
separation and particle elongation to occur. Marcon et al. (1971) have shown, for
alnicos containing more than 28% Co, that it is practically impossible to avoid the
precipitation of the deleterious y phase during thermomagnetic treatment unless
appropriate amounts of a-stabilizing elements such as silicon or titanium are added.
Unfortunately titanium inhibits the formation of the columnar grain structure.
However, Gould (1964), McCaig (1964) and Harrison and Wright (1967) have
shown that the addition of small quantities of sulphur, selenium or tellurium
facilitates the formation and growth of columnar grains. Thus the addition of
titanium to the alnicos, although beneficial, means that much more careful control
of the composition and heat treatment of the alloys is required. It is for this
reason and the higher cobalt content that the high coercivity alnicos 8 and 9 are
generally more expensive than the other alnicos. The effects of additions of Ti, S,
Nb, and other elements on the magnetic and metallurgical properties of the
158 R.A. McCURRIE

alnicos have been investigated by Clegg (1966, 1970), Palmer and Shaw (1969),
Higuchi (1966), Hoffmann and St/iblein (1970) and Wright (1970).

10. Dependence of the magnetic properties on the angle between the direction of
measurement and the preferred or easy axis of magnetization

The variation of the coercivity with angle 0 to the preferred axis in alnico 9 and
ticonal 900 (single crystal) has been measured by McCurrie and Jackson (1980) who
observed small maxima at 0 = 60 ° followed by a rapid decrease to zero at 90 ° (fig. 29).
According to Shtrikman and Treves (1959) the reduced coercivity hc as a function of
the angle of m e a s u r e m e n t 0 and the reduced radius for infinitely long cyclindrical
rods with shape anisotropy only is given by

Hc 1.08(1- 1.08S 2)
hc = 2XMs- $2[(1 _ 1.08S_2)2 ( 1 - 2.16S -2) sin 2 0] 1/2'

where S = R/Ro (Ro = (4"n't.zoA/JZs)1/2) and 0 is the angle between the easy axis of
magnetization (i.e., the axis of the cylinder) and the direction of m e a s u r e m e n t of He.
Comparison of the results shown in fig. 29 with the theoretical curves obtained from
the above relation shown in fig. 30 suggests that the observed angular variation of Hc
in these alloys is due to magnetization reversal by the curling mechanism. The
coercivities of alnico 9 and ticona1900 are - 0 . 2 5 H A (HA is the anisotropy field) which
are in good agreement with the values to be expected if the magnetization reverses by
curling (Shtrikman and Treves 1959). The angular variation of the r e m a n e n c e
coercivity Hr (Hr is the reverse field required to reduce the r e m a n e n c e to zero)
showed even m o r e pronounced maxima at 0 ~ 80 ° followed by a rapid decrease to
zero at 0 ~ 90 °. Unfortunately no detailed theory of the variation of Hr with 0 is
available for comparison, though Kneller (1966, 1969) has suggested that Hr(O)/Hc(O)
varies between the limits:

1 (0 = 0 °) <~H,-(O)/Hc(O) ~< o~ (0 = 90°).

Paine and Luborsky (1960) have measured the angular variation of the intrinsic
coercivity of alnico 5DG, columnar alnico, which was found to be almost
independent of the angle up to about 40 ° after which Hc decreased and at 0 = 90 ° it
was about half the value at 0 = 0 °. For a specimen of alnico 9 the coercivity
decreased with the angle 0 and at 0 = 90 °, it was about a quarter of the value at
0 = 0 °.
Bate (1961) found that for partially aligned y-Fe203 particles Hr(O)/Hc(O)
increased from 1.25 at 0 = 0 ° to 2.3 at 0 = 90 °.
The dependence of the magnetic properties on the crystal orientations (i.e.,
crystallographic texture) in columnar alnico 5-7 has been investigated in con-
siderable detail by McCaig and Wright (1960), Makino and Kimura (1965) and
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 159

3,0 AHrT
tXHrA
10 I
• HcT 0-9
i 2.5
E 0"8 ~ S < 1
O'7 ~ 3
,~, 2.0
0.6
"r" j¢ S=1"47
o,+
e- 1"5

..r"
~, 1.0 0"3

>

o= 0.5
0~
O o.1 S--6,0 _ . ~ \
+ I I I I 0 I I 1 I [ I I I ~1

0 15 30 45 60 75 90 0 10 20 30 40 50 60 70 80 90
Angle 0 ° from easy axis in degrees 0o

Fig. 29. Dependence of the coercivity Hc and Fig. 30. Reduced coercivity hc for an aligned array of
remanence coercivity Hr on the angle 0 between infinitely long single domain cylindrical particles
the preferred axis and the direction of measure- with shape anisotropy only as a function of the angle
ment for columnar alnico 9 (A) and single crystal 0 between the easy axis (i.e. the cylindrical axis) and
alnico 9 (T) (McCurrie and Jackson 1980). the direction of measurement. S is the reduced
particle radius. (S ~ R/Ro where R0 =
(4~A)l/2//z~/2Ms and A is the exchange constant.)
The curve for S < 1 is identical to that for the
Stoner-Wohlfarth coherent rotation theory (after
Shtrikman and Treves 1959).

M o o n (1974), M c C a i g a n d W r i g h t (1960) h a v e m a d e m e a s u r e m e n t s of t h e m a g -
netic p r o p e r t i e s at v a r i o u s angles to the p r e f e r r e d d i r e c t i o n in t h r e e c o l u m n a r
alnico m a g n e t s (49.7 wt % F e - 25% C o - 13.5% N i - 8% A 1 - 3% C u - 0.8% N b )
a n d h a v e s h o w n t h a t t h e d e c r e a s e in r e m a n e n c e with angle is in g o o d q u a l i t a t i v e
a g r e e m e n t with results p r e d i c t e d f r o m t h e S t o n e r - W o h l f a r t h (1948) t h e o r y . F o r an
a l i g n e d a s s e m b l y of e l o n g a t e d p a r t i c l e s t h e r e m a n e n c e at an angle 0 to t h e
p r e f e r r e d d i r e c t i o n of m a g n e t i z a t i o n , i.e., t h e c o l u m n a r a x i s - n o m i n a l l y p a r a l l e l to
the [100] d i r e c t i o n , is given by

B~ : L = J+ cos O.

F o r angles > 7 5 ° t h e m e a s u r e d v a l u e s of Jr ( M c C a i g a n d W r i g h t 1960) are l a r g e r


t h a n t h o s e p r e d i c t e d by t h e a b o v e e x p r e s s i o n (fig. 31). T h e v a r i a t i o n of t h e
c o e r c i v i t y with t h e a n g l e b e t w e e n t h e p r e f e r r e d d i r e c t i o n a n d t h e d i r e c t i o n of
m e a s u r e m e n t is shown in fig. 32 f r o m which it can b e seen that the coercivities are
a l m o s t c o n s t a n t up to an angle of 20 ° . T h e l a t t e r lack of a g r e e m e n t with t h e
160 R.A. McCURRIE

1 0 0 ~ l ~ ~ o C o l u m n a r alnico5
r, .~ - ~-~sw Theory

60 .\
°~
= ' ~ '~ r

E 20 SW Theory ~ , , ~ , v

0 I I I ~ ~ ~ ~--'~R41

0 30 6O 90
Angle e °

Fig. 31. Variation of remanence Br and maximum energy product (BH)max with angle 0 to the
preferred axis in columnar alnico, 5DG (alnico 5-7) and comparison with Stoner-Wohlfarth (1948)
theory (after McCaig and Wright 1960).

100

8O

60

•~- 4O
._>
@
O
(3
20

0
0 30 60 90
Angle e °
Fig. 32. Variation of coercivity ~ with angle 0 to the preferred axis in columnar alnico 5DG (alnico
5-7) and comparison with Stoner-Wohlfarth (1948) theory (after McCaig and Wright 1960).
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 161

Stoner-Wohlfarth (1948) theory is due to incomplete alignment of the particles,


magnetostatic interaction between the particles and to incoherent magnetization
reversal by curling (see Zijlstra, chapter 2 of this handbook). The variation of the
(BH)ma~ product with angle to the preferred direction is shown in fig. 31 from
which it can be seen that the decrease in (BH)r~ax with increasing angle is much
less rapid than that predicted from the Stoner±Wohlfarth (1948) theory.

11. Relationship between magnetic properties and crystallographic texture

The relationship between the crystal texture and the magnetic properties of alnico
8 (33 wt % F e - 36% C o - 1 5 % N i - 7% A 1 - 5 % T i - 4% C u ) h a s been studied by
Higuchi and Miyamoto (1970). The magnets were solution heat treated for 30 min
at 1250°C and then annealed in a magnetic field of 240 kAm -1 for 8 min at 810°C,
after which they were given a two-step anneal at about 600°C. Unlike the previous
studies discussed above the crystal textures of the specimens were determined
from their magnetization curves assuming that changes in the magnetization
occurred by coherent rotation of the magnetization vector. Higuchi and Miyamoto
(1970) showed that there was very good agreement between the measured values
of Br, Hc and (BH)m~x and the values calculated from their model. The highest
observed energy product was 89 kJm -3.
The relationship between magnetic properties and crystallographic texture of
alnico 5 has also been investigated by Moon (1974). In this work measurements of
the orientation (by an X-ray diffractometer) and magnetic properties were made
on several specimens from a series of columnar magnets with a nominal com-
position 50.4wt % F e - 2 4 . 5 % C o - 1 3 . 5 % N i - 8 % A 1 - 3 % C u - 0 . 6 % Nb. The
magnets were cooled from 1250°C at 1.2°Cs -1 in a magnetic field - 2 2 0 kAm -~
parallel to the direction of solidification, annealed at 590°C for 48 h and then
annealed for a further 48 h at 560°C. Moon showed that the energy product
(BH)max decreased rapidly with increase in the standard deviation of the angle 0
between the [100] direction and the axis of the magnet and suggested that small
improvements in the casting techniques and hence the crystallographic texture
might lead to a significant improvement in the (BH)ma× product.

12. Effects of particle misalignment on the remanence and coercivity of the


anisotropic field-treated alnicos

12.1. R e m a n e n c e

Although the magnetic properties of the alnicos depend on the degree of particle
alignment, a considerable deviation from complete alignment can be tolerated
without a very significant deterioration in the magnetic properties. Suppose that
the axes of elongation of the particles in a given specimen all lie within a cone of
162 R.A. M c C U R R I E

0 ~X

Fig. 33. Definition of solid angle dO for determination of r e m a n e n c e of an array of uniaxial single
domain particles whose preferred axes of magnetization lie within the cone with semi-vertical angle 0.

semi-solid angle 0 m a s shown in fig. 33 where 0 m is the maximum deviation


from the preferred axis.
The number of particles with their axes at an angle 0 is proportional to the
solid angle contained within the interval 0 and 0 + dO viz. 27r sin 0 dO. So that the
remanence is given by

IOOm27r sin 0 cos 0 dO


_ Js(1 + cos 0m)
Jr=L" or. (1)
fo 2~- sin 0 dO

so that even when 0m = 30 °, Jr/Js = 0.933. Unfortunately it is difficult to compare


this simple theory with experimental measurements because of the difficulty in
obtaining samples which have their particle axes included in a suitable range of
measurable 0m angles. Furthermore the cosine function is relatively insensitive to
changes in 0m (fig. 34). McCurrie (1981) has suggested that a more accurate

1.0
Jr
Js 0.8

0.6

0-4

0.2

0 I I I I I I I I I

0 30 60 90
Angle e m
Fig. 34. D e p e n d e n c e of r e m a n e n c e on m a x i m u m angle of'deviation 0m of the particle axes from
preferred axis of magnetization (McCurrie 1981).
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 163

estimate of the degree of alignment can be o b t a i n e d by measuring the saturation


r e m a n e n c e at 90 ° to the preferred or easy axis. In this case,

Jr _ (Om -- 1 sin 2Ore)


J~ vr(1 -- COS Om) '

or

J~/Js ~ ~ sin Ore.

These functions, which are shown in fig. 35, are m u c h m o r e sensitive to changes in
0m than the function given in eq. (1).

0.5

Js _~sinem,,,~,/
,,;,/
0'3 / g ~

/,,/\
,~' (.em-
~ _ ~ ' .~Sin2ern)
_ 2
0-1 ' ~1 2c-'~SOm)

0 i ! ! i I l | ! i
0 30 60 90
Angle 0 m
Fig. 35. D e p e n d e n c e of r e m a n e n c e m e a s u r e d perpendicular to preferred axis and the m a x i m u m angle
of deviation 0m of the particle axes from the preferred axis of the magnet (McCurrie 198l).

12.2. Coercivity

T h e coercivity is also decreased when the axes of elongation of the particles are at
angles in the range 0 to 0m to the preferred axis. F r o m the S t o n e r - W o h l f a r t h
(1948) theory of hysteresis in single domain particles it can be shown that for a
single particle whose axis of elongation is at an angle 180-0 to the applied field
(i.e. a demagnetizing field) the coercivity is given by:

Hc = (Dz - Dx)Ms (1 - tan 2/3 0 + tan 4/3 0) 1/2


( 1 + t a n 2/30) , for0<0~<45 ° ,

and

He = }(Dz - Dx)Ms sin 20, for 45 ° ~ 0 ~< 90 ° .


164 R.A. McCURRIE

The variation of the coercivity with angle 0 is shown in fig. 32, where the ordinate
has been plotted in reduced units (% of He at 0 = 0°). From fig. 32 it can be seen
that the coercivity decreases rapidly for 0 up to about 15° after which the decrease
with increasing 0 is m u c h less rapid. For a reasonably well aligned array of
particles in an alnico alloy 0 is unlikely to be greater than about 30 ° at the
maximum.

13. Determination of the optimum volume fraction of the F e - C o rich particles

Since the F e - C o rich particles are required to be elongated single domains they
must not be precipitated in a quantity sufficient to cause coalescence. This means
that the composition of the alloys must be chosen so that the volume fraction of
the F e - C o rich cq phase is surrounded by the non-ferromagnetic or weakly
ferromagnetic Ni-A1 rich ~2 phase. A further restriction is placed on the volume
fraction, p, of the F e - C o rich al phase because the coercivities of the alloys are
reduced by particle interaction effects which are usually assumed to be propor-
tional to ( 1 - p).
If we assume that for a packing fraction p and coherent magnetization reversal,
the coercivity is given by (see for example, Ndel 1947a, Kittel 1949, Kittel and
Galt 1956, Brown 1962, Compaan and Zijlstra 1962):

H e ( p ) = H~(O)(1 - p ) ,

where He(0) is the extrapolated coercivity when p = 0, then the value of p for
which the energy product is a maximum can be determined as follows. Since the
maximum possible coercivity Hc(0) is given by

He(p) = [(Dz - Dx)(1 - p ) M ~ ] / p ,

and taking (Dz - Dx) ~ ½,

1 1
He(p) - 5(1 - p ) M ~ _ (1 - p) Js,
p 2p, o

where Ms~ is the average or overall saturation magnetization of the alloy and Js is
the saturation magnetic polarization of the particles. (Note that if the mag-
netization reversal occurs entirely by curling then Hc is independent of p because
curling produces no interaction fields (Kneller 1969)i) However, since mag-
netostatic interaction domains (section 16) are observed during the demag-
netization of the alnicos it appears that the particles do produce interaction fields
so that some coherent magnetization reversal must also occur. Thus for the
present approximate calculation it seems that the use of the above relation for
He(p) is justified.
Consider an alnico alloy with a rectangular M - H hysteresis loop with an
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 165

intrinsic coercivity H¢(p) as shown in fig. 36(a). The corresponding B - H demag-


netization curve is shown in fig. 36(b). The flux coercivity ,Hc = He(p) because the
M - H loop is rectangular so that the discontinuous magnetization and flux
reversals take place in the same reverse field. The maximum energy product for
this B - H demagnetization curve is given by the area OPQR, i.e.,

(BH)max = (PJs - ½(1 - p)Js)(1 - p)Js/ZI-Lo.

Note that the saturation magnetic polarization of the alloy as a whole is Jm = PJs.
From the above expression we find that

(BH)m,x = (J~/4tzo)(-3p 2 + 4p - 1),

which has a maximum value when

d(BH)max/dp = 0 and d2(BH)m,x/dp 2< O,

i.e., when

6p=4, or p=2.

Thus (BH)max is equal to J~/12/x0. This value of (BH)max is obtainable only when
the intrinsic coercivity, He, satisfies the condition

H~ = (1 - p)JJ2/x0.

For the magnetically annealed alnicos the absolute maximum in the (BH)max
product is obtained when p = 0.6-0.7 which is in good agreement with both the

i ~
Mr=Ms B
Br=Jm=PJs

-H Hc H ?,
/ /
/
/
/
/
Z
/ R
'H He BHc O
(a) (b)
Fig. 36. (a) Ideal rectangular M - H hysteresis loop with intrinsic coercivity iH¢. (b) Corresponding B - H
demagnetization curve for M - H loop in (a). (NB: The flux coercivity BHc=iH¢, the intrinsic
coercivity because the M - H loop is rectangular.)
166 R.A. McCURRIE

above theories. The packing fraction p can be determined by quantitative


metallography (see for example, Hilliard 1962, 1967, 1968, Hilliard and Cahn
1961, Underwood 1970, 1973, Saltykov 1970) of the two phases which are clearly
visible on the replication electron micrographs such as those shown in figs. 22(a)
and (b). Another estimate of p can be obtained from the measured value of the
saturation magnetic polarization, Jm, of the magnet alloy as a whole if it is
assumed that the individual particles have a saturation magnetic polarization Js,
equal to that of FeCo and by using the relationship

Jm = PJ~.

For field heat treated grain oriented alnico 5-7 the saturation magnetic polariza-
tion is approximately 1.4 T. Since Js = 2.4 T for FeCo we find that p = 0.6 which is
very close to the values determined by quantitative metallography (De Vos 1966,
1969, Granovsky et al. 1967, Takeuchi and Iwama 1976) and from magnetic
measurements (Bulygina and Sergeyev 1969, Sergeyev and Bulygina 1970).
Apart from the reduction in the coercivity due to the packing fraction p of the
particles (H~(p)= H~(0)(I-p)), Kondorsky (1952a, b) and Aharoni (1959) have
shown that the critical diameter for infinitely long single domain particles is given by:

Re(p) = Re(0) (1 - p)-l/e,

so that R~(p) increases with the packing fraction.


For materials such as the alnicos in which the coercivity is due almost entirely to
the shape anisotropy of elongated single domain particles of Fe-Co the critical radius
Re(0) (for an infinitely long cylinder) is given by

Rc(0) = (1.08/2Dz )1/2(47"rtzoA/J2s)1,2 ,


i.e.,
Rc(0) = 1.04(4 7"rtzoA/J~) ~/2 ,

since the demagnetization factor Dz for an infinitely long cylinder is 0.5. In the above
relation A is the exchange energy coefficient, Js the saturation magnetic polarization
and/z0 = 47r x 10 -7 Hm -1 is the magnetic constant. A more detailed discussion of the
critical sizes for single domain particles is given by Zijlstra in chapter 2 of this
handbook.

14. Interpretation of the magnetic properties in terms of the Stoner-Wohifarth


theory of hysteresis in single domain particles

The magnetic hysteresis of non-interacting single domain particles with uniaxial


shape anisotropy has been investigated theoretically by Stoner and Wohlfarth
(1948). They investigated the behaviour of a single domain ellipsoid of revolution
when subjected to increasing magnetic fields at various angles to the easy axis of
S T R U C T U R E A N D P R O P E R T I E S O F A L N I C O P E R M A N E N T M A G N E T ALLOYS 167

magnetization of the ellipsoid assuming that changes in magnetization occurred by


coherent rotation of the magnetization vector. The relationship between the
applied field, the easy axis and the magnetization is shown in fig. 37. For a particle
with a saturation magnetization of Ms the total magnetic free energy is

1 2
E = ½1xo(Dz - D x ) M 2 sin 2 ~ + gl.xoDxMs + IxoHMs cos qS. (2)

where Dx and Dz are the demagnetization factors parallel and perpendicular to


the axis of elongation of the particle. The particle is assumed to be a prolate
ellipsoid of revolution so that Dy = D~ and Dy > Dx. Thus in a given field H the
stable positions of the magnetization vector are those which correspond to the
minima of eq. (2). The solutions, for various values of the field and the angle
0, have been tabulated by Stoner and Wohlfarth (1948) who have also plotted
several representative hysteresis loops notably those for 0 = 0, 45 ° and 90 °. For a
survey of the results the reader is referred to Zijlstra, (chapter 2 in this hand-
book). One of the most important conclusions from the Stoner-Wohlfarth theory
is that the coercivity of an aligned assembly of identical non-interacting (i.e.,
infinitely dilute) single domain particles with uniaxial shape anisotropy is

Hc = (D: - D x ) M s ,

i.e., the coercivity is directly proportional to the saturation magnetization and to


the difference in the demagnetization factors perpendicular and parallel to the
easy axis. For a randomly orientated array of such particles the coercivity is
reduced to

Hc = 0.479(Dz - D x ) M s .

The dependence of ( D z - Dx) on the degree of elongation, i.e., the ratio of the
lengths of the x- and z-axes of the particle, is shown in fig. 38. It has already been
mentioned in section 1.2 that in view of the complex microstructures and particle
shapes in the isotropic alnicos 1-4 it is not possible to give a quantitative
comparison or interpretation of their magnetic properties in terms of the Stoner-
Wohlfarth (1948) theory though their magnetic properties can be understood

Ms

Fig. 37. Relationship between the applied field, easy axis and magnetization vector in a prolate
ellipsoidal single domain particle with uniaxial shape anisotropy.
168 R.A. M c C U R R I E

0.5

0"4

Dz
~0-3
I
N

0'2 Dx

0"1

0 I I I I I I I I
1 2 3 4 5 6 7 8 9 10
Dimensional ratio m = a / b
Fig. 38. Difference (Dz- Dx) between demagnetization factors perpendicular and parallel to the
preferred axis of magnetization as a function of the particle elongation or dimensional ratio m = a/b,
where a is the semi-major and b the semi-minor axis of the prolate ellipsoid of revolution.

qualitatively. In any case in the alnicos 1-4 both the al and a2 phases are
ferromagnetic so that magnetization changes by domain wall movement may also
occur.
From the electron micrographs of alnico 5 and alnico 5-7, the precipitated
F e - C o rich particles are rod-like (i.e., they approximate to prolate ellipsoids of
revolution) and have dimensions ~150 nm long and ~40 nm in diameter. From fig.
38 the corresponding value of Dz - Dx is ~0.5 and hence if we take the saturation
magnetization of the particles Ms = 1.7x 106Am -1 which is considerably larger
than the observed value of 60 kAm -1. Although the particles are not fully aligned
this is insufficient to account for the difference in the two values. Since the
coercivity depends on the presence of demagnetizing fields it is clear that particle
interactions cannot be neglected. Many authors (see e.g., N6el 1947a, Compaan
and Zijlstra 1962) suggest that the effects of particle interactions on the coercivity
are best described by the relation

He(p)= H o ( O ) ( 1 - p ) ,
where He(p) is the coercivity for a particle packing fraction p and He(0) is the
coercivity for a packing fraction of zero. Thus if the effects of particle interactions
are included the coercivity of a fully aligned array of single domain particles with
uniaxial shape anisotropy is

Ho= (1-p)(Dz -Dx)Ms, (3)


or

Hc = (1 - p ) ( D z - Dx)M's/p,

where M'~ is the average saturation magnetization of the alloy. The packing factor
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 169

p is usually in the range 0.6-0.7 and (Dz-D~) approaches 0.5 so that the
theoretical value of Hc ~ 300 kAm -~. Thus even if particle interaction effects are
included there is still a very significant difference between the theoretical and
observed coercivities (-~60 kAm-1).
Baran (1959) has derived a more general expression for the coercivity of a fully
aligned array of particles which takes account of the reduction in coercivity due to
particle interactions and the possibility that the matrix is also ferromagnetic, viz.

Hc = p(1 - p)(Dz - Dx)(MI - M2)2/M's, (4)

where p is the volume fraction of the most strongly ferromagnetic phase, M1 and
M2 the saturation magnetizations of this phase and the matrix phase respectively,
M's the overall saturation magnetization of the alloy and Dz and Dx are the
demagnetization factors Of the particles perpendicular and parallel to the axis of
elongation respectively. For the maximum coercivity it is clear from eq. (4) that
the difference M 1 - M 2 should be as" large as possible.
If it is assumed that the alloy is fully heat treated, i.e., sufficient time has been
allowed for all the Fe and Co atoms to diffuse to the precipitated particles, then
the matrix saturation magnetization M2 ~ 0 and the saturation magnetization of
the Fe--Co rich particles M1 = Ms = M's/p, so that eq. (4) reduces to eq. (3).
From the electron micrographs of alnico 8 shown in figs. 22(a) and (b) (De Vos
1966, 1969) the F e - C o rich particles are very long rods and approximate to
prolate ellipsoids of revolution with length to diameter ratios l/d ~-16 cor-
responding to D z - D~--~ 0.5. Quantitative metallography shows that the volume
fraction of the F e - C o rich particles p ~ 0.65. Since there is a very high degree of
particle alignment we may calculate the coercivity from eq. (1). If Ms is taken to
be at least equal to the value for pure iron viz. 1.7 MAre -1 (the saturation
magnetization for Fe-30% Co is about 10% higher than that of pure iron) the
theoretical coercivity as derived from eq. (3) is ~300 kAm -1 which compares very
favourably with the highest experimental value of 180 kAm -~ for alnico 8. The
discrepancy between the observed value of the coercivity and that calculated
according to the Stoner-Wohlfarth (1948) theory suggest that the assumption that
changes in the magnetization occur by coherent rotation is invalid (see Zijlstra,
chapter 2 in this handbook).

15. Interpretation of the magnetic properties in terms of magnetization reversal by


the curling mechanism

According to calculations by Kondorsky (1952a), Brown (1957, 1963, 1969), Frei et al.
(1957) and Aharoni and Shtrikman (1958) magnetization reversal by coherent
rotation in infinitely long cylinders with shape anisotropy only, should occur only
when the cylinder radius R is less than a critical radius R0 = (4zrA)llZ/Izaol2Ms
where A is the exchange constant,/z0 is the magnetic constant 4~- × 10 -7 H m -1 and Ms
the saturation magnetization. Above this radius magnetization reversal occurs by the
curling mechanism (apart from a very small range of radii for which magnetization
170 R.A. McCURRIE

reversal occurs by a process known as buckling). According to Aharoni and


Shtrikman (1958) the dependence of the coercivity on the cylinder radius R is given
by

Hc = 1.08(Ro/R )ZMs/2 , (5)

where it is of course assumed that D z - Dx has its maximum value of 0.5 for an
infinitely long cylinder. The curling mechanism is discussed in chapter 2 of this
handbook by Zijlstra who has also included a diagram showing the configuration of
the magnetic Spins during the reversal process. From measurements on the electron
micrographs of alnico 8 obtained by De Vos (1966, 1969) (see figs. 22(a) and (b)) the
particles have radii ~15 nm. Although the particles are obviously not infinitely long
cylinders their length to diameter ratio l/d ~ 16 so that assuming magnetization
reversals occur by curling we may estimate the coercivity from eq. (5) from which we
find that Hc ~ 2.4 × 105 A m -1 (McCurrie and Jackson 1980).
The highest observed flux coercivity for alnico 8 BHo ~ 1.8 × 105 A m -1 (this is
slightly less than the intrinsic coercivity He) so that the theoretical and observed
coercivities are in reasonable agreement. When the magnetization reversal occurs
entirely by curling there are no particle interaction effects so that the coercivity is
independent of the packing fraction p. However, the appearance of magnetostatic
interaction domains (discussed below) during demagnetization suggests that some
interparticle interaction does occur. From measurements of the angular variation
of the coercivity and rotational hysteresis in columnar and single crystal alnico 9
(see sections 10 and 18) McCurrie and Jackson (1980) have concluded that
magnetization reversal does occur by the curling mechanism and that the quan-
titative results are in reasonable agreement with those predicted on the basis of
the theory proposed by Shtrikman and Treves (1959).

16. Magnetostatic interaction domains in alnicos

Magnetization reversal in the alnicos is further complicated by the formation of


magnetostatic interaction domains which are due to interparticle interactions.
Although it is well established that the alnicos contain single domain particles,
macroscopic surface domain structures known as magnetostatic interaction
domains, can sometimes be observed by the Bitter colloid technique, as shown by
Nesbitt and Williams (1950), Bates (1955), Bates and Martin (1955), Kussmann and
Wollenberger (1956), Schulze (1956), Andr/i (1956), Kronenberg and Tenzer
(1958), Bates et al. (1962) and Iwama (1968). This apparently contradictory
observation can be explained as follows. Consider the demagnetization process of
an alnico magnet. As the magnetization of the particles is gradually reversed we
should expect that as a result of the long-range interaction fields, particles in a
particular region which have already reversed their magnetization will tend to
assist the reversal of those in adjacent regions. Thus in the demagnetized state,
when just half the total magnetization has been reversed, the magnet is effectively
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 171

Fig. 39. Magnetostatic interaction domains in alnico 5 (Bates et al. 1962).

divided up into a domain structure as shown in fig. 39. At the boundaries between
the various regions or interaction domains there are stray magnetic fields and
gradients which attract the colloidal particles and so delineate the domain
structure. These boundaries are of course quite different from the conventional
Bloch-Lifshitz walls (Kittel 1949, Kittel and Galt 1956). Similar magnetostatic
interaction domains have also been observed on elongated single domain magnets
by Craik and Isaac (1960). For detailed discussions of magnetostatic interaction
domains and the mechanism of their formation the reader is referred to the
papers by Craik and Lane (1967, 1969). The effects of particle interactions have
also been discussed by Wohlfarth (1955).

17. Comparison of the N6ei-Zijlstra and Cahn theories of magnetic annealing in


alnico alloys

The improvements in the magnetic properties of alnico alloys by heat treatment in


a magnetic field were first observed in alnico 5 (see section 2) but the following
discussion also includes the effects which are observed when alnicos 6, 8 and 9 are
given an isothermal heat treatment in a magnetic field.
172 R.A. McCURRIE

17.1. Ndel-Zijlstra theory

In the theory of magnetic annealing proposed by N6el (1947b) and later


developed by Zijlstra (1960-1962) the spinodal decomposition at high tem-
peratures results in the formation of very small spherical particles and is con-
sidered to be complete before the elongation of the particles begins. If the
decomposition takes place in a magnetic field the spherical particles develop into
ellipsoids with their axes of elongation parallel to the applied field thereby
reducing their total magnetic free energy. During the thermomagnetic treatment
the particles increase their size and elongation very rapidly because at high
temperatures the rate of atomic diffusion is high. The applied field should of
course be sufficient to saturate the Fe-Co rich particles in order to achieve the
maximum possible alignment during the relatively short thermomagnetic heat
treatment.
The relevant counteracting energies are the interfacial energy F~ between the a~
and a2 phases and the difference in the magnetostatic energy Fm between the
decomposition waves parallel and perpendicular to the applied field. During the
elongation of the particles there is arl increase in the interfacial energy but this is
accompanied by a larger reduction in the magnetic energy of the particles in the
applied field. Thus the elongation is energetically favourable when the ratio Fm/Fs
is large. In the N6el-Zijlstra theory F~ is defined as the product of the interface
tension and the amount of interface per unit volume. Thus the ratio Fm/Fs will
increase in a coarsening structure so that the interfacial free energy is of great
importance in determining the efficacy of magnetic annealing. According to
Zijlstra (1960-1962) the rates of elongation and coarsening of the particles are
proportional to the square of the difference in the saturation magnetic polariza-
tions of the OgI and a2 phases (AJs)2 and inversely proportional to the interfacial
energy Y, i.e., the efficacy of the thermomagnetic treatment is proportional to
(AJs)2/T. This theory explains why thermomagnetic treatment is most effective in
alnico alloys with a high cobalt content (i.e. alnicos 5-9) which raises the Curie
temperature and AJs and lowers the interfacial energy Y. From measurements on
alnico 5 Zijlstra (1960) concluded that 3' ~0.1 Jm -2, a result which confirms an
earlier suggestion by Kittel et al. (1950) who also suggested that since the
interfacial energy is small the particles should develop a shape anisotropy in order
to minimize their magnetic energy in the applied field. Thus according to the
N6el-Zijlstra theory the axes of elongation of the particles should be parallel to
the applied field irrespective of its orientation relative to particular crystallo-
graphic directions in the alloy. This conclusion suggests that the shape anisotropy
for finite annealing time at a given temperature should give the same result for
both polycrystalline and single crystal specimens. However, Zijlstra's (1960)
experiments show that the anisotropy of the single crystal specimens is consider-
ably larger than that for the polycrystalline specimens. Zijlstra (1960) suggested
that this difference could be due to a contribution from magnetocrystalline
anisotropy, but the magnitude of the difference makes this unlikely.
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 173

17.2. C a h n ' s theory

The spinodal decomposition of the alnico alloys into two phases al and Ol2 occurs
because it is accompanied by a reduction in the total free energy FT of the alloy.
The effect of the magnetic field in producing elongated particles with their axes of
elongation parallel to the field direction has been investigated theoretically by
Cahn (1963). He suggested that the effect of the magnetic field is to suppress the
spinodal decomposition waves along the magnetic field direction and that this is
due to the difference in the magnetic energy Fm of the spinodal waves parallel and
perpendicular to the field direction which favours the formation of long rods
parallel to the field direction. In the absence of the magnetic field the spinodally
decomposed system consists of three mutually perpendicular spinodal decom-
position waves, i.e., a cubic array of isotropic particles. Cahn (1963) stated that the
two main sources of anisotropy are the magnetic and elastic energies. The former
favours compositional waves parallel to the internal field direction while the latter
favours compositional waves parallel to particular crystallographic directions.
Waves parallel to the (100) directions in the {100} planes in cubic crystals are
favoured when

2C44- Cl1+ C12>0 ,

where the Cii coefficients are elastic energy constants from the stress tensor. If the
anisotropy in the magnetic energy is much larger than the anisotropy in the elastic
energy then the geometry of the decomposition will be independent of the
crystallographic orientation so that the axes of elongation of the particles will be
parallel to the applied field direction. However, if the anisotropy in the elastic
energy predominates then the axes of elongation of the particles will be parallel to
the crystallographic direction which minimizes the elastic energy. When the elastic
and magnetic anisotropy energies are of comparable magnitude the axes of
elongation of the particles lie between the field direction and the nearest (100)
direction. In particular if the applied field is parallel to a (100) direction then the
axes of elongation of the particles are parallel to the chosen (100) direction.
Cahn (1963) also suggested that for a solid solution with a composition
fluctuation ( x - x 0 ) where x0 is the average composition, the anisotropy in the
magnetic energy is proportional to (OJs/Ox) 2 where Js is the saturation magnetic
polarization of the precipitated particles. The quantity (OJs/Ox) 2 is expected to
vary rapidly with temperature and to be very large near the Curie temperature,
To, so that thermomagnetic treatment will be most effective when the spinodal
decomposition temperature T is at, or just below, the Curie temperature. If the
alloy is cooled to lower temperatures the effectiveness of the field diminishes
rapidly. The anisotropy in the elastic energy depends on two factors: (1) it is
proportional to (d In a / d x ) 2 where 'a' is the stress-free lattice parameter, and (2) it
is proportional to the variation in the elastic energy coefficient with crystallo-
graphic direction; the latter can be approximated by Ay = ]71100]- 71111][. Cahn
174 R.A. McCURRIE

(1963) estimated that the elastic energy is much greater than the magnetic energy
except near the Curie temperature. Thus according to Cahn's (1963) theory a
large ratio of Fm/F-r favours the elongation of the particles when they are heat
treated in a magnetic field. The shape of the particles is established during the first
stages of the decomposition and the magnetic shape anisotropy can be further
increased only by a subsequent increase in their saturation magnetic polarization
by diffusion of atoms between the two phases al and a2 without thereby changing
their shape, though their wavelength is increased by this process.

17.3. Discussion of the Ndel-Zijlstra and Cahn theories

Electron micrographs of alnico 8 (De Vos 1966, 1969) which had been given
an isothermal heat treatment in a magnetic field show that even in the initial
stages of the spinodal decomposition the three (100) decomposition waves are
developed which have practically the same wavelength, a result which disagrees
with Cahn (1963) who suggested that the (100) decomposition waves per-
pendicular to the magnetic field direction are suppressed from the beginning of
the spinodal decomposition. De Vos's (1966, 1969) electron micrographs of
alnico 8 also show that elongated Fe-Co rich particles with their axes of elon-
gation parallel to the (100} directions develop even in the absence of an applied
field; this result is also in disagreement with Cahn's (1963) theoretical predictions.
According to the discussion in section 7.2 of the relationship between the field
direction and the preferred direction of magnetization Heidenreich and Nesbitt
(1952) and Nesbitt and Heidenreich (1952a, b) found that when the field direction was
parallel to a principal crystallographic direction, i.e., [100], [110] or [111], the easy
or preferred direction of magnetization is parallel to the field direction. However,
Heidenreich and Nesbitt (1952), Nesbitt and Heidenreich (1952a, b), Hoselitz and
McCaig (1951), McCaig (1953) and Yermolenko et al. (1964) showed that when
the thermomagnetic field is applied at an angle to a (100) direction, the
direction of easy magnetization lies between the field direction and the chosen
(100) direction. The above authors also found that the magnetic anisotropy (at
room temperature) has a maximum value when the field is parallel to a (100)
direction. Thus from the results presented in section 7.2 and the short summary of
these given above there is strong evidence that in addition to the magnetic and
interracial energies the anisotropy of elastic energy has some effect in determining
the orientation of the elongated Fe-Co rich particles during thermomagnetic
treatment. This conclusion is also supported by the microstructural observations
made by De Vos (1966, 1969). Although Cahn's (1963) suggestion that the anisotropy
of the elastic energy is important in determining the nature of the spinodal
decomposition waves does not agree with the electron microscopic observations of
De Vos (1966, 1969) it is clear that the orientation of the particles is affected by the
anisotropy of the elastic energy.
The experiments performed by Zijlstra (1960-1962) and De Vos (1966,
1969) show that the Ndel-Zijlstra theory of thermomagnetic treatment or mag-
netic annealing is in better agreement with the experimental results than that
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 175

proposed by Cahn (1963). Strong support for this conclusion is provided by the
following three experiments carried out by Zijlstra (1960-1962):
(1) A specimen of alnico 5 (51% Fe, 24% Co, 14% Ni, 8% A1 and 3% Cu by
weight) was heat treated in a magnetic field of 128 kAm -1 parallel to the [100]
direction for 7 h at 748°C and subsequently a part of the same specimen was
annealed for 24h at the same temperature with the field parallel to the [010]
direction. A part of this specimen was then annealed for a further 24 h at 748°C.
Electron micrographs of each of the three thermomagnetic treatments (actually
thermal magnetic anneals) are shown in figs. 40(a), (b) and (c) from which it can be
seen that when the alloy is annealed in a field perpendicular to the initial field
direction, the particles become elongated in the new field direction thus reducing
their magnetic free energy, in agreement with the N6el-Zijlstra theory. The
electron micrographs shown in figs. 41(a) and (b) also confirm that in the initial
stages of the phase separation the Fe-Co rich particles are spherical (fig. 41(a)) and
that they are elongated by heat treatment in the magnetic field-the direction of
elongation being parallel to the field direction as shown in fig. 41(b).
(2) A polycrystalline specimen (disk) of alnico 5 with the same composition as
that given above was heat treated in a non-inductively wound furnace for 7 h at

i~ i~ i~

a b C
t ....!

Fig. 40. Microscopical demonstration of crossed-field annealing of a single crystal: (a) after heat
treatment for 7 hours at 748°C with field along the [100] direction; (b) after subsequent heat treatment for
24 hours at 748°C with field along the [010] direction; (c) after final heat treatment for 24 hours at 748°C
with field along the [010] direction. All three micrographs are made of the (001) plane (after Zijlstra 1960).
176 R.A. McCURRIE

a I b
2pm
Fig. 41. (a) Electron micrograph of an alnico showing that in the initial stages of the phase separation
the Fe-Co rich particles are spherical. (b) Electron micrograph after heat treatment in a magnetic field.
The direction of elongation of the Fe-Co rich particles is parallel to the direction of the field (after Zijlstra
1962).

755°C in the absence of an applied field. After this treatment the specimen was
shown to be isotropic but when it was subsequently heat treated in a magnetic
field of 6 4 0 k A m -t at the same t e m p e r a t u r e it became anisotropic with an
anisotropy energy approximately equal to that of a polycrystalline alnico 5
specimen which had been given a m o r e conventional thermomagnetic heat
treatment, i.e., Ku ~ 8 x 104 Jm -3.
(3) A specimen of polycrystalline Mishima alloy, Fe2NiA1, which is not usually
considered to respond to thermomagnetic treatment was heat treated in a field of
640 k A m -1 for 2 weeks at 725°C after which it became magnetically anisotropic
with an anisotropy energy K ~ 6 × 104 Jm73 which is comparable to that of alnico
5. After this treatment, however, the particles are too large to be single domains
so that the coercivity is correspondingly low ~10 k A m -1 parallel to the preferred
direction, i.e., the field direction.
None of the above three results is to be expected according to the theory
proposed by Cahn (1963). However, it should be emphasized that Zijlstra's
(1960-1962) results were obtained by heat treatment times much longer than those
used in commercial practice so that it seems likely for shorter thermomagnetic
heat treatments (e.g. in cooling of alnico 5) that the orientation of the axes of
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 177

elongation are partly determined by the anisotropy in the elastic energy but there
now seems to be little doubt that the N6el-Zijlstra theory of the thermomagnetic
treatment of the alnicos is in good general agreement with the experimental
results.

18. Rotational hysteresis

The rotational hysteresis energy W~(H) in a field is defined as the energy required to
rotate the specimen through 360 °, i.e.,

Wr(H) = f~'~ F(O) dO.

It can also be conveniently obtained by measuring the area enclosed by the 360 °
clockwise torque curve and the 360 ° anticlockwise torque curve (McCurrie and
Jackson 1980); the rotational hysteresis energy W~(H) is equal to half this area, i.e.,

W~CH)= ½ Fc(O) dO + /'Ac(O) dO ,


JO

where Fc(0) and FAt(0) are the torque curves in a given field for clockwise and
anticlockwise rotation in the applied field. The clockwise and anticlockwise torque
curves corresponding to the maximum value of Wr(H) for single crystal alnico 9 are
shown in fig. 42. McCurrie and Jackson (1980) have measured the rotational
hysteresis energies of alnico 9 and single crystal alnico 9 as a function of the applied
field and their results are shown in fig. 43. Note the very rapid variation in W~(H) for
applied fields close to the coercivity. From the value of the rotational hysteresis
integrals, defined by

j~

where Wr(H) is the rotational hysteresis energy observed in a field H, R can


therefore be obtained from the area under the Wr(H)/Js vs 1/H curve (Bean and
Meiklejohn 1956, Jacobs and Luborsky 1957, Luborsky 1961, Luborsky and
Morelock 1964). McCurrie and Jackson (1980) have concluded that magnetization
reversal in alnico 9 and ticonal 900 occurs by curling. This conclusion is also
supported by their measurements of the angular variation of the coercivity discussed
in section 10.
The rotational hysteresis energy of single crystal alnico 8 as a function of the
applied field has also been measured by Livshitz et al. (1970c). They also observed a
very sharp peak in the rotational hysteresis energy but in contrast to the results
obtained by McCurrie and Jackson (1980) they observed a smaller peak in a higher
applied f i e l d - about 2.5 times the field corresponding to the first sharp peak.
178 R.A. M c C U R R I E

1"00 I

0'751- Ha= 188 kAm -1


H c = 138 kAm -1

0"50

I
.-j
E 0.25 Acw w
0
90 12( 270
0 Angle 0 °
~0.25

~0-50
Ticonal 9 0 0
(100) plane
- 0-75

-1.00 u-
Fig. 42. Torque curves showing rotational hysteresis in a single crystal of alnico 9 (ticonal 900) in an
applied field of 188 k A m -1 (McCurrie and Jackson 1980). T h e rotational hysteresis energy Wr(H)can be
determined by halving the area enclosed between the two curves: C W - c l o c k w i s e rotation; A C W -
anticlockwise rotation.

2"0

1.5
E
• Ticonal 9 0 0

1.0

0-5

0 .P~ I , ~--..-.-~ I
0 2 4 6 8 10
H ~'105Am -1)
Fig. 43. Variation of rotational hysteresis energy of columnar alnico 9 and single crystal alnico 9
(ticonal 900) with applied field (McCurrie and Jackson 1980).
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 179

19. Anhysteretic magnetization

The anhysteretic magnetization of a ferromagnetic material is the magnetization


which results from the application of a static field H and a superimposed
alternating field Ha > H which is gradually reduced to zero while the static field H
is still on. The anhysteretic magnetization c u r v e - s o m e t i m e s referred to as the
ideal magnetization c u r v e - r i s e s much more steeply than the conventional curve
and if the specimen has a low or zero external demagnetization factor the initial
slope of the anhysteretic magnetization curve is known as the anhysteretic
susceptibility Ka. A simple general theory of anhysteretic magnetization curves has
been presented by Ndel (1943) and N6el et al. (1943) but for a short review the
reader is referred to Kneller (1969).
The reciprocal K21 is known as the internal demagnetization factor. Dussler
(1927) has shown that the geometrical demagnetization factor Dg of open
magnetic circuits also depends on the structural details of the material. The
demagnetizing field HD for a uniformly magnetized ellipsoid is given by

HD = D g M ,

but for non-uniformly magnetized particles such as occur in the alnicos it has
been found by experiment that

Ka 1 = D, where D > Og.

The difference D i = AD = D - D g has been attributed to internal interaction


effects arising from the structure of the materials. Unfortunately, Di is not a
constant, characteristic of the material, but also depends on Dg as one would
expect since the shape and structural effects are due fundamentally to mag-
netostatic interactions.
Bulgakov (1950) found that for alnico 5 that the values of Di parallel and
perpendicular to the preferred axis of magnetization were respectively 0.02 and 0.4.
Similar measurements were also made by Gould and McCaig (1954) on alnico 5
who obtained values of 0.01 and 0.5 for D~. They suggested that the precipitated
particles were oriented parallel to the thermomagnetic field and separated by a
non-ferromagnetic matrix, but they stressed that their measurements of D~ did not
enable them to make any really firm conclusions.

20. Magnetic viscosity

In all the above discussions of coercivity theory it has been assumed that after
saturation of a ferromagnetic material in the forward direction the demag-
netization curve represents instantaneous and stable values of B or M as a
function of the reverse applied field H.
However, if a steady reverse field less than the coercivity is applied after
180 R.A. M c C U R R I E

saturation, the magnetization decreases with time t after the application of the
field H. This time dependent change of the magnetization in an applied field is
known as magnetic viscosity. The magnetic viscosity of cast alnico 2 (54% Fe,
18% Ni, 12% Co, 10% A1, 6% Cu) as measured by Street and Woolley (1949) is
shown in fig. 44.
Measurements on the same alloy which had been heat treated at 1250°C for 20
minutes and then allowed to cool at about 2°C per second gave similar results to
the cast alloy. According to Street and Woolley (1949) the magnetic polarization
at time t is given by the empirical relation

J=C-Slnt,

where C is a constant and S is a parameter which depends on the temperature T


and the applied field H.
Since a similar effect also occurs when the applied field is positive, the
phenomenon of magnetic viscosity is a general property of all ferromagnetic
materials, though its presence is not always readily observable. The above
equation suggests that the magnetization reversal by the field H involves the
thermally activated surmounting of energy barriers in the material.
For further details on the magnetic viscosity in the alnicos the reader is referred
to the papers by Bulgakov and Kondorsky (1949), Street and Woolley (1949, 1950,
1956), Street et al. (1952a, b), Phillips et al. (1954) and Barbier (1954).

0"20

_ Alnico 2 ~ 523°K

0-15

-ff
<~
0.10

0"0,'

-- ~ 86°K

-I 30 60 120 300 600 1200


I
Time t in s e c

Fig. 44. C h a n g e in net magnetic polarization as a function of time at various temperatures for alnico 2
(after Street and Woolley 1949) 54% Fe - 18% Ni - 12% C o - 10% Al - 6% Co. Cast alloy cooled from
1250°C held at this temperature for 20 rain and then cooled at 2°Cs -1, followed by 2 h at 600°C. Not in
o p t i m u m p e r m a n e n t magnet state. However, the fully heat treated alloy gave similar magnetic viscosity
results.
STRUCTURE AND PROPERTIES OF ALNICOPERMANENTMAGNETALLOYS 181

21. Temperature dependence of magnetic properties

The effects of temperature on the magnetic properties of the alnicos are com-
plicated and depend on the previous heat treatment of the alloys. Clegg and
McCaig (1957) investigated the temperature dependence of the saturation mag-
netic polarization, the coercivity BHc and the remanence of alnico 5 after various
heat treatments. The temperature dependence of the coercivity for various
specimens is shown in fig. 9. Curves (c) and (g) represent specimens with optimum
permanent magnet properties and may therefore be regarded as typical.
The temperature dependence of the magnetic properties of alnicos has also
been investigated by McCaig (1968), Dietrich (1966a, b, 1967), Clegg and McCaig
(1958), Roberts (1958), Clegg (1955), Pawlek and Reichel (1955), Bulgakov (1949)
and Jellinghaus (1943).The effects of low temperatures on alnico 5 and alnico 5-7
alloys have been studied by Clegg (1955).
In a different type of investigation Bates and Simpson (1955) measured the heat
changes which occur when the alnicos are cycled through a hysteresis loop.
For general discussions of the temperature dependence and stability of the
magnetic properties of permanent magnets the reader is referred to Gould (1958,
1962), Parker and Studders (1962), Schiller and Brinkmann (1970), Heck (1974),
Joksch (1975) and McCaig (1977), Gould and McCaig (1954).

22. Dynamic excitation (AC magnetization)

Very few studies have been made on the effects of applying alternating fields to
alnico magnets. The response of the magnets is complex due to the effects of
magnetic viscosity and eddy currents and therefore depends on the frequency and
magnitude of the applied fields.
Sabir and Shepherd (1974) have studied the behaviour of alnico 5 in sinusoidal
fields in the frequency range 50-350 Hz. The energy loss per unit volume was
obtained by measuring the areas under the dynamic B - H loops and agreed with
direct wattmeter measurements and also with theoretical calculations. Eddy
current losses were found to be negligible at 50 Hz and even at 350 Hz were only
about 10% of the total loss. The total loss at a fixed field was found to be almost a
linear function of the frequency. The dynamic excitation of alnico 5 has also been
studied by Shepherd et al. (1974).

23. Prospects for improvement in the magnetic properties

Since the observed coercivities are only ~0.3HA (the anisotropy field H A =
(Dz- Dx)Ms) even in the most favourable cases there is certainly some room for
improvement in He. This could be achieved by reducing the particle thickness to
avoid curling and to make the particles more regular in shape to prevent parasitic
nucleation of magnetization reversal. However, the attainment of these objectives
is likely to present considerable metallurgical difficulties.
182 R.A. McCURRIE

Since the coercivity depends on the difference in the demagnetization factors Dz


and Dx it can readily be appreciated from fig. 38 that when the length to diameter
ratio (m) of the particles is greater than about 10, the coercivity is about 94% of
the theoretical m a x i m u m for a given Ms and p so that there is little to be gained
by any further increase in rn resulting from more elaborate thermomagnetic
treatment. No way has yet been found to increase Ms of the Fe-Co-rich particles
significantly so it appears that there is little prospect for any further i m p r o v e m e n t
in the coercivities of the best alnicos, 8 and 9.
The p e r m a n e n t magnet properties of single crystal alnico 8 and 9 are of course
significantly better than those of the fully columnar alnico 9 owing to the higher
degree of particle alignment but any prospective i m p r o v e m e n t in their properties
is similarly limited by the factors already discussed.
From table 3 it can be seen that in a highly oriented columnar magnet,
remanences Jr ~ 1.43 T can be achieved so that the m a x i m u m possible energy
product (BH)ma× is

(BH)max = B~/4tXo = J~/41Xo = 400 kJm 3,


provided of course that the intrinsic coercivity Hc satisfies the condition:

Hc > Jd2/z0,
i.e.,

Hc > 570 k A m -~ .

However, in view of the requirements mentioned above there seems to be little


prospect of increasing the coercivity Hc much beyond about 180 k A m 1 without
reducing the remanence.
The best laboratory properties of alnico p e r m a n e n t magnets have been
obtained by Naastepad (1966) for a single crystal of an alloy containing 35 wt %
Fe, 34.8% Co, 14.9% Ni, 7.5% A1, 5.4% Ti and 2.4% Cu which had a coercivity
of 122 k A m 1 and an energy product (BH)max of 107 kJm -3. Unfortunately these
exceptionally high values of the m a x i m u m energy product cannot be obtained on
a commercial scale, but it is possible to obtain a commercial alnico 9 with
BHc = 120 k A m -1 and a (BH)max of 75 kJm -3.
It appears that the alnicos, which have been available since 1931, have now
reached the zenith of their development and there is little prospect that any
further substantial i m p r o v e m e n t in their magnetic properties can be made.

24. Summary

The relatively high coercivities of the alnicos are due to the shape anisotropy of
Fe or F e - C o rich single domain particles which are precipitated in a weakly
ferromagnetic or non-ferromagnetic N i - A I rich matrix.
After cooling from about 1200°C at a controlled rate ~30°Cs -I the isotropic
S T R U C T U R E AND PROPERTIES OF ALNICO P E R M A N E N T M A G N E T ALLOYS 183

alnicos 1-4 are subsequently tempered for several hours at about 600-650°C.
The anisotropic alnicos 5, 6, 7 and the grain oriented alnico 5DG are produced
by controlled cooling from 1250°C in a saturating magnetic field ~200-300 kAm 1
and then tempered for several hours at about 600°C. The high coercivity alnicos 8
and 9 are also cooled from about 1250°C and then annealed isothermally in a
saturating magnetic field ~200-300 kAm 1 for a few minutes at about 820°C after
which they are given a two-stage tempering treatment usually for several hours at
about 650°C and then at 550°C.
The phase separation takes place by spinodal decomposition at 800-850°C as
the alloys are cooled. The final shapes and sizes of the particles are determined in
the very early stages of the spinodal decomposition and microstructural in-
vestigations have shown that the Fe or F e - C o rich particles are elongated parallel
to the (100) directions in the matrix. In the isotropic Fe2NiA1 and Fe-Ni-Al-low
cobalt containing alloys the particles often have mixed rod-like and plate-like
shapes which form a complex three-dimensional interconnected structure in which
preference for particle axis alignment with the (100) directions is not well defined.
Microstructural studies of the anisotropic field treated alnicos 5-9 show that the
precipitated F e - C o rich particles are rod-like, and preferentially aligned parallel
to the field direction during thermomagnetic treatment. In these alloys the higher
coercivities can be attributed to an increase in the shape anisotropy of the
precipitated particles though there is evidence to show that the magnetization
reversal occurs by the incoherent mechanism known as curling, rather than coherent
rotation of the magnetization vector.
In the crystal oriented alnicos 5-7 and 9, the direction of the applied field
during the thermomagnetic treatment should be as close as possible to a particular
(100) columnar axis in order to maximize the remanence, coercivity and energy
product. In order for the thermomagnetic treatment to be effective the Curie
temperature must be above the spinodal decomposition temperature. Further-
more the ratio of (2xJ)2/-/ (where AJ is the difference between the saturation
magnetic polarizations of the cq and c~2 phases and 3' is the interracial energy)
between the F e - C o rich particles and the matrix should be as high as possible so
that the particles can be elongated (by minimization of their magnetic free energy)
in the field direction. All of the above requirements can be satisfied if the cobalt
content is in the range 24-42 wt %. This high cobalt content also has the beneficial
effects of increasing the remanence and the coercivity (He ~ Js). The addition of Ti
to alnicos 8 and 9 lowers the saturation magnetic polarization and Curie tem-
perature of the matrix, but increases the perfection and elongation of the
particles. Titanium also increases the time required for the thermomagnetic
treatment to be effective and it is for this reason that alnicos 8 and 9 are given an
isothermal heat treatment in a magnetic field at about 820°C for a few minutes.
The principal effect of the tempering heat treatment is to increase the difference
•J in the saturation magnetic polarization between the F e - C o rich particles and
the Ni-A1 rich matrix. This increases the shape anisotropy field of the precipitated
particles which is proportional to AJ and hence increases the coercivity. The
shapes and the sizes of the particles do not change significantly during this heat
treatment. The optimum volume fraction of the particles p ~ 0.6-0.7.
184 R.A. McCURRIE

References Bronner, C., J. Sauze, E. Planchard, J.M.


Drapier, D. Coutsouradis and L. Habraken,
Aharoni, A., 1959, J. Appl. Phys. Suppl. 30, 70 S. 1967, Cobalt, No. 36, 123.
Aharoni, A. and S. Shtrikman, 1958, Phys. Bronner, C., J.-P. Haberer, E. Planchard, J.
Rev., 109, 1522. Sauze, J.M. Drapier, D. Coutsouradis and L.
Albanese, G., G. Asti and R. Criscuoli, 1970, Habraken, 1968, Cobalt, No. 40, 131.
IEEE Trans. Magn., MAG-6, 161. Bronner, C., J.-P. Haberer, E. Planchard, J.
Allec, G., 1971, Cobalt, No. 52, 156. Sauze, J.M. Drapier, D. Coutsouradis and L.
Andrfi, W., 1956, Ann. Phys. (Leipzig) 19, 10. Habraken, 1969, Cobalt, No. 42, 14.
Arbuzov, M.P. and A.A. Pavlyukov, 1965, Fiz. Bronner, C., J.-P. Haberer, E. Planchard, J.
Met. Metalloved. 20, No. 5, 724 (Engl. Sauze, J.M. Drapier, D. Coutsouradis and L.
Transl., Phys. Met. Metallog. 20, No. 5, 83). Habraken, 1970, Cobalt, No. 48, 111.
Baran, W., 1959, Tech. Mitt. Krupp, 17, 150. Bronner, C., J.-P. Haberer, E. Planchard and J.
Barbier, J.C., 1954, Ann. Phys. (Paris) 9, 84. Sauze, 1970a, Cobalt, No. 46, 15.
Bate, G., 1961, J. Appl. Phys., Suppl. 32, 239 S. Bronner, C., J.-P. Haberer, E. Planchard and J.
Bates, L.F., 1955, Research, 8, 462. Sauze, 1970b, Cobalt, No. 49, 187.
Bates, L.F. and D.H. Martin, 1955, Proc. Phys. Brown, Jr., W.F., 1957, Phys. Rev. 105, 1479.
Soc. Lond. B68, 537. Brown, Jr., W.F. 1962, Magnetostatic Principles
Bates, L.F. and A.W. Simpson, 1955, Proc. in Ferromagnetism (North-Holland, Am-
Phys. Soc. Lond. B68, 849. sterdam) p. 117.
Bates, L.F., D.J. Craik and E.D. Isaac, 1962, Brown, Jr., W.F., 1963, Micromagnetics (Inter-
Proc. Phys. Soc. Lond. 79, 970. science/Wiley, New York) p. 79.
Bean, C.P. and W.H. Meiklejohn, 1956, Bull. Brown, Jr., W.F., 1969, Ann. New York Acad.
Amer. Phys. Soc., Set. II, 1, 148. Sci. 147, 461.
Belova, V.M., V.I. Nikolaev, S.Yu. Stephano- Bulgakov, N.V., 1949, Dokl. Akad. Nauk
vich and S.S. Yakimov, 1969, Fiz. Tverd. (SSSR) 69, 627.
Tela, 11, 3662. Bulgakov, N.V., 1950, Dokl. Akad. Nauk,
Belozersky, G.N., Yu.N. Grinblat and A.I. (SSSR) 70, 205.
Shapiro, 1971, Fiz. Met. Metalloved. 32, No. Bulgakov, N.V. and E. Kondorsky, 1949, Dokl.
2, 301. (Engl. Transl., Phys. Met. MetaIlog. Akad. Nauk. (SSSR) 69, 325.
32, No. 2, p. 75). Bulygina, T.I. and V.V. Sergeyev, 1969, Fiz.
Berkowitz, A.E., 1969, Constitution of Multi- Met. Metalloved. 27, No. 4, 703. (Engl.
phase Alloys, ch. VII of 'Magnetism and Transl., Phys. Met. Metallog. 27, No. 4, 132).
Metallurgy', A.E. Berkowitz and E. Kneller, Burgers, W.G. and J.L. Snoek, 1935, Physica, 2,
eds. (Academic Press, New York, London) 1064.
Vol. 1, p. 331. Cahn, J.W., 1961, Acta Met. 9, 795.
Betteridge, W., 1938, J. Iron and Steel Inst. Cahn, J.W., 1962, Acta Met. 10, 179.
139, 187. Cahn, J.W., 1963, J. Appl. Phys. 34, 3581.
Bradley, A.J., 1949a, J. Iron and Steel Inst. 163, Cahn, J.W., 1968, Trans. Met. Soc. AIME, 242,
19. 166.
Bradley, A.J., 1949b, Physica, 15, 175. Cahn, J.W. and J.E. Hilliard, 1958, J. Chem.
Bradley, A.J., 1951, J. Iron and Steel Inst. 168, Phys. 28, 258.
233. Cahn, J.W. and J.E. Hilliard, 1959, J. Chem.
Bradley, A.J., 1952, J. Iron and Steel Inst. 171, Phys. 31,688.
41. Clegg, A.G., 1955, Brit. J. Appl. Phys. 6, 120.
Bradley, A.J. and A. Taylor, 1938a, Proc. Roy. Clegg, A.G., 1966, Z. Angew. Phys. 21, 77.
Soc. Lond. A166, 353. Clegg, A.G., 1970, IEEE, Trans. Magn. MAG-6,
Bradley, A.J. and A. Taylor, 1938b, Magnetism 201.
(Institute of Physics, London) 89. Ciegg, A.G. and M. McCaig, 1957, Proc. Phys.
Bronner, C., 1970 IEEE, Trans. Magn. MAG-6, Soc. Lond. B70, 817.
301. Clegg, A.G. and M. McCaig, 1958, Brit. J.
Bronner, C., E. Planchard and J. Sauze, 1966a, Appl. Phys. 9, 194.
Cobalt, No. 31, 63; 1966b, Cobalt, No. 32, Compaan, K. and H. Zijlstra, 1962, Phys. Rev.
124. 126, 1722.
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 185

Craik, D.J. and E.D. Isaac, 1960, Proc. Phys. Gould, J.E., 1971, Cobalt Alloy Permanent
Soc. 76, 160. Magnets (Centre d'Information du Cobalt,
Craik, D.J. and R. Lane, 1967, Brit. J. Appl. Brussels).
Phys. 18, 1269. Gould, J.E. and M. McCaig, 1954, Proc. Phys.
Craik, D.J. and R. Lane, 1969, Brit. J. Appl. Soc. Lond. B67, 584.
Phys. (J. Phys. D. Appl. Phys.) Ser. 2, Vol. 2, Granovsky, B., P.P. Pashkov, V.V. Sergeyev
33. and A.A. Fridman, 1967, Fiz. Met. Metal-
Cronk, E.R., 1966, J. Appl. Phys. 37, 1097. loved. 23, No. 3, 444 (Engl. Transl., Phys.
Dean, A.V. and J.J. Mason, 1969, Cobalt, No. Met. Metallog. 23, No. 3, 55).
43, p. 73. Guillaud, C., 1953, Rev. Mod. Phys. 25, 64.
Dehler, H., 1942, Stahl u Eisen, 47, 983. Haanstra, H.B., J.J. de Jong and J.M.G.
De Jong, J.J., J.M.G. Smeets and H.B. Haan- Smeets, 1957, Philips Tech. Rev. 19, 11.
stra, 1958, J. Appl. Phys. 29, 297. Hansen, J.R., 1955, Proc. Conf. Magnetism and
De Vos, K.J., 1966, Doctoral Thesis, Tech- Magnetic Materials (Pittsburgh, Pa, USA) p.
nical High School (Eindhoven, The Nether- 198.
lands). Harrison, J., 1966, Z. Angew. Phys. 21, 101.
De Vos, K.J., 1969, 'Alnico Permanent Magnet Harrison, J. and W. Wright, 1967, Cobalt, No.
Alloys', ch. 9 in: Magnetism and Metallurgy, 35, 63.
A.E. Berkowitz and E. Kneller, eds. Heck, C., 1974, Magnetic Materials and their
(Academic Press, New York, London) p. 473. Applications (Butterworths, London).
Dietrich, H., 1966a, Cobalt, No. 30, 3. Heidenreich, R.D. and E.A. Nesbitt, 1952, J.
Dietrich, H., 1966b, A. Angew. Phys. 21, 125. Appl. Phys. 23, 352.
Dietrich, H., 1967, Cobalt, No. 35, 78. Heimke, G. and R. Kohlhaas, 1966, Z. Angew.
Durand-Charre, M., C. Bronner and J.-P. Phys. 21, 73.
Lagarde, 1978, IEEE: Trans. Magn. MAG-14, Heimke, G., H. van Kempen and R. Kohlhaas,
797. 1966, IEEE, Trans. Magn. MAG-2, 411.
Dussler, E., 1927, Z. Phys. 44, 286. Higuchi, A., 1966, Z. Angew. Phys. 21, 80.
Ebeling, D.G. and A.A. Burr, 1953, J. Metals, Higuchi, A. and T. Miyamoto, 1970, IEEE,
5, 537. Trans. Magn. MAG-6, 218.
Edwards, A., 1957, Elect. Energy, 1, 146, 178. Hillert, M., 1961, Acta Met. 9, 525.
Edwards, A., 1962, Magnet Design and Selec- Hilliard, J.E., 1962, Trans. AIME, 224, 1201.
tion of Material, ch. 6 of Permanent Magnets Hilliard, J.E., 1967, Determination of Struc-
and Magnetism, D. Hadfield, ed. (Iliffe tural Anisotropy, Proc. 2nd Intern. Con-
Books Ltd., London; Wiley, New York) p. gress for Stereology, H. Elias, ed. (Springer,
191. New York) p. 219.
Fahlenbrach, H., 1954, Tech. Mitt. Krupp, 12, Hilliard, J.E., 1968, Measurement of Volume in
177. Volume, ch. 3 of Quantitative Microscopy,
Fahlenbrach, H., 1955, Naturwissenschaften, R.T. DeHoff and F.N. Rhines, eds.
42, 64. (McGraw-Hill, New York, St. Louis, San
Fahlenbrach, H., 1956, Tech. Mitt. Krupp, 14, Francisco, Toronto, London, Sydney) pp. 45-
2. 76.
Fahlenbrach, H. and H. Stfiblein, 1964, Proc. Hilliard, J.E. and J.W. Cahn, 1961, Trans.
Intern. Conf. Magnetism (Inst. Phys., Not- Amer. Inst. Min. Engrs. 221, 344.
tingham, UK) p. 767. Hoffmann, A. and H. St~blein, 1966, Tech.
Frei, E.H., S. Shtrikman and D. Treves, 1957, Mitt. Krupp, Forsch. Ber. 24, 113.
Phys. Rev. 106, 445. Hoffmann, A. and H. St~iblein, 1967, Z. Angew.
Fujiwara, T. and T. Kato, 1960, J. Japan. Inst. Phys. 23, 182.
Metals, 24, 526. Hoffmann, A. and H. Stiiblein, 1970, IEEE,
Gould, J.E., 1958, Instrum. Practice, 12, 1083. Trans. Magn. MAG-.6, 225.
Gould, J.E., 1962, Magnetic Stability, ch. 10 of Hoffmann, A. and P. Pant, 1970, Tech. Mitt.
Permanent Magnets and Magnetism, D. Krupp, Forsch. Ber. 28, 117.
Hadfield, ed. (Iliffe Books Ltd, London; Honda, K., 1934, Metallwirtschaft, 13, pp. 425-
Wiley, New York). p. 443. 427.
Gould, J.E., 1964, Cobalt, No. 23, 82.
186 R.A. McCURRIE

Hoselitz, K. and M. McCaig, 1949a, Proc. Phys. Koshiba, S. and T. Nishinuma, 1957, J. Jap.
Soc.. Lond. B62, 163. Inst. Metals, 21, 166.
Hoselitz, K. and M. McCaig, 1949b, Nature Koshiba, S. and T. Nishinuma, 1960, J. Jap.
(Lond.) 164, 581. Inst. Metals, 24, 433.
Hoselitz, K. and M. McCaig, 1951, Proc. Phys. Kronenberg, K.J., 1954, Z. Metallk. 45, 440.
Soc. Lond. B64, 549. Kronenberg, K.J., 1960, J. Appl. Phys. 31, 80 S.
Hoselitz, K. and M. McCaig, 1952, Proc. Phys. Kronenberg, K.J., 1961, J. Appl. Phys. 32, 196
Soc. Lond. B65, 229. S.
Iwama, Y., 1968, Trans. Jap. Inst. Metals, 9, Kronenberg, K.J, and R.K. Tenzer, 1958, J.
273. Appl. Phys. 29, 299.
Iwama, Y., M. Inagaki and T. Miyamoto, 1970, Kussmann, A. and J.H. WoIlenberger, 1956, Z.
Trans. Jap. Inst. Metals, 11,268. Angew. Phys. 8, 213.
Iwama, Y., M. Takeuchi and M. Iwata, 1976, Kussman, A. and O. Yamada, 1956, Arch.
Trans. Jap. Inst. Metals, 17, 481. Elektrotech. 42, 237.
Jacobs, I.S. and F.E. Luborsky, 1957, J. Appl. Lange, G., 1968, Deutsche Edelstahlwerke,
Phys. 28, 467. Tech. Ber. 8, 209.
Jellinghaus, W., 1943, Arch. Eisenhfittenw. 16, Lindner, E., R. Wittig and K. Prissier, 1963,
247. Neue Hfitte, 8, 557.
Joksch, C., 1975, Physica, 80B, 199. Livshitz, B.G., B.A. Samarin and V.S. Shu-
Jonas, B. and H.J. Meerkamp van Embden, bakov, 1970a, IEEE, Trans. Magn., MAG-6,
1941, Philips Tech. Rev. 6, 8. 242.
Julien, C.A. and E G . Jones, 1965a, J. Appl. Livshitz, B.G., E.G. Knizhnik, G.S. Kraposhin
Phys. Suppl. 36, 1173. and Y.S. Linetsky, 1970b, IEEE, Trans.
Julien, C.A. and F.G. Jones, 1965b, Cobalt, No. Magn., MAG-6, 237.
27, 73. Livshitz, B.G., V.I. Sumin and A.S. Lileev,
Kittel, C., 1949, Rev. Mod. Phys. 21, 541. 1970c, IEEE, Trans. Magn., MAG-6, 169.
Kittel, C. and J.K. Galt, 1956, Solid State Phy- Luborsky, F.E., 1961, J. Appl. Phys. Suppl. 32,
sics, vol. 3, F. Seitz and D. Turnbull, eds. 171 S.
(Academic Press, New York, London) pp. Luborsky, F.E. and C.R. Morelock, 1964, J.
437-564. Appl. Phys. 35, 2055.
Kittel, C., E.A. Nesbitt and W. Shockley, 1950, Luteijn, A.J. and K.J. de Vos, 1956, Philips
Phys. Rev. 77, 839. Res. Repts. 11, 489.
Kneller, E., 1966, Handbueh der Physik, S. Makarov, E.F., V.A. Povitsky, E.B. Granovsky
Fl/igge, ed., Ferromagnetism, vol. XVIII pt. and A.A. Fridman, 1967, Phys. Status Solidi,
2, H.P.J. Wijn, ed. (Springer, Berlin, Hei- 24, 45.
delberg, New York) p. 512. Makarov, V.A., E.B. Granovsky, E.F. Makarov
Kneller, E., 1969, Fine Particle Theory, ch. VIII and V.A. Povitsky, 1972, Phys. Status Solidi,
of Magnetism and Metallurgy, A.E. Berko- 14, (A) 331.
witz and E. Kneller, eds. (Academic Press, Makino, N., 1962, Cobalt, No. 17, 3.
New York, London) vol. 1, p. 365. Makino, N. and Y. Kimura, 1965, J. Appl.
Koch, A.J.J., M.G. van der Steeg and K.J. de Phys. 36, 1185.
Vos, 1957, Proc. Conf. Magnetism and Mag- Makino, N., Y. Kimura and I. Yamaki, 1963, J.
netic Materials (Amer. I.E.E., Boston, Mass., Jap. Inst. Metals 27, 582.
USA) p. 173. Marcon, G., C. Bronner and R. Peffen, 1971;
Koch, A.J.J., M.G. van der Steeg and K.J. de Cobalt, No. 51, 99.
Vos, 1959, Berichte der Arbeitsgemeinschaft Marcon, G., R. Peffen and H. Lemaire, 1978a,
Ferromagnetismus (Riederer, Stuttgart) p. IEEE, Trans. Magn., MAG-14, 685.
130. Marcon, G., R. Peffen and H. Lemaire, 1978b,
Kolbe, C.L. and D.L. Martin, 1960, J. Appl. IEEE, Trans. Magn., MAG-14, 688.
Phys., Suppl. 31, 84 S. Mason, J.J., D.W. Ashall and A.V. Dean, 1970,
Kondorsky, E., 1952a, Izvest. Akad. Nauk SSSR, IEEE, Trans. Magn., MAG-6, 19l.
16, 398. McCaig, M., 1949, Proc. Phys. Soc. Lond. B62,
Kondorsky, E., 1952b, Dokl. Akad. Nauk SSSR, 652.
82, 365. McCaig, M., 1953, J. Appl. Phys. 24, 366.
STRUCTURE AND PROPERTIES OF ALNICO PERMANENT MAGNET ALLOYS 187

McCaig, M., 1966, Z. Angew. Phys. 21, 66. 1969, J. Appl., Phys. Suppl. 40, 1308.
McCaig, M., 1968, Cobalt, No. 41, 196. Pashkov, P.P., A. Fridman and E. Granovsky,
McCaig, M., 1977, Stability of Permanent 1970, IEEE, Trans. Magn., MAG-6, 211.
Magnets, ch. 5 of Permanent Magnets in Pater, M., R. Bloch and E. Krainer, 1963, Z.
Theory and Practice (Pentech Press, London, Angew. Phys. 15, 261.
Plymouth) p. 155. Pawlek, F. and K. Reichel, 1955, Z. Metallk.
McCaig, M. and W. Wright, 1960, Brit. J. Appl. 46, 308.
Phys. 11, 279. Pfeiffer, I., 1969, Cobalt, No. 44, 115.
McCurrie, R.A., 1981, J. Appl. Phys. 52, Phillips, J.H., R. Street and J.C. Woolley, 1954,
7344. Phil. Mag. 45, 505.
McCurrie, R.A. and S. Jackson, 1980, IEEE Planchard, E., R. Meyer and C. Bronner, 1964a,
Trans. Magn. MAG-16, 1310. Z. Angew. Phys. 17, 174.
Mishima, T., 1932, Ohm, 19, 353. Planchard, E., C. Bronner and J. Sauze, 1964b,
Moon, J.R., 1974, J. Phys. D: Appl, Phys. 7, Proceedings of the Journ6es Internationales
1233. des Applications du Cobalt (Brussels, 9-11th
Naastepad, P.A., 1966, Z. Angew. Phys. 21, June, 1964)pp. 134-511.
104. Planchard, E., C. Bronner and J. Sauze, 1965,
N6el, L., 1943, Cahiers Phys. 17, 47. Cobalt, No. 28, 132.
N6el, L., 1947a, Comptes Rendus Acad. Sci. Planchard, E., C. Bronner and J. Sauze, 1966a,
(Paris) 224, 1550. Z. Angew. Phys. 21, 63.
N6el, L., 1947b, Comptes Rendus Acad. Sci. Planchard, E., C. Bronner and J. Sauze, 1966b,
(Paris) 225, 109. Z. Angew. Phys. 21, 95.
N6el, L., R. Forrer, N. Janet and R. Battle, Povitsky, V.A., E.B. Granovsky, A.A. Frid-
1943, Cahiers Phys. 17, 51. man, E.F. Makarov and P.P. Pashkov, 1970,
Nesbitt, E.A., 1950, J. Appl. Phys. 21, 879. IEEE, Trans. Magn., MAG-6, 215.
Nesbitt, E.A. and H.J. Williams, 1950, Phys. Povolotskii, E.G., Ya.M. Dovgalevskii and
Rev. 80, 112. V.K. Baitina, 1963, Metal Science and Heat
Nesbitt, E.A. and R.D. Heidenreich, 1952, J. Treatment, 11, 631.
Appl. Phys. 23, 366. Ritzow, G., 1963, Neue Hiitte, 8, 282.
Nesbitt, E.A. and H.J. Williams, 1955, J. Appl. Ritzow, G. and W. Ebert, 1957, Deutsche
Phys. 26, 1217. Elektrotechnik, 11, 527.
Nesbitt, E.A. and H.J. Williams, 1957, Proc. Roberts, W.H., 1958, J. Appl. Phys. 29, 405.
Conf. Magn. and Magn. Materials (Amer. Sabir, S.A.Y. and W. Shepherd, 1974, Proc.
I.E.E., Boston, Mass., USA) p. 184. IEE, 121, No. 8, 907.
Nesbitt, E.A., H.J. Williams and R.M. Bozorth, Saltykov, S.A., 1970, Stereometric Metallo-
1954, J. Appl. Phys. 25, 1014. graphy (Metallurgizdat, Moscow) 3rd ed.
Nicholson, R.B. and P.J. Tufton, 1966, Z. Shepherd, W., H. Gaskell and P. Rashid, 1974,
Angew. Phys. 21, 59. IEEE, Trans, Magn., MAG-10, 50.
Oliver, D.A. and J.W. Shedden, 1938, Nature Schiller, K., 1968, Deutsche Edelstahlwerke,
(Lond.) 142, 209. Tech. Ber. 8, 147.
Paine, T.O. and F.E. Luborsky, 1960, J. Appl. Schiller, K. and K. Brinkmann, 1970, Dauer-
Phys. Suppl. 31, 78 S. magnete, Werkstoffe und Anwendungen
Palmer, D.J. and S.W.K. Shaw, 1969, Cobalt, (Springer, Berlin, Heidelberg, New York).
No. 43, 63. Schulz, L.G., 1949, J. Appl. Phys. 20, 1030.
Pant, P., 1974, Proc. 3rd Europ. Conf. on Hard Schulze, D., 1956, Exptl. Tech. Phys. 4, 193.
Magnetic Materials (Amsterdam, 17-19 Sep- Schwartz, L.H., 1976, Ferrous Alloy Phase
tember, 1974) H. Zijlstra, ed. (Bond voor Transformations, ch. 2 in: Applications of
Material en Kennis, P.O. Box 9321, Den M6ssbauer Spectroscopy, R. Cohen, ed.
Haag, The Netherlands) p. 186. (Academic Press, New York, San Francisco,
Parker, R.J. and R.J. Studders, 1962, Per- London) vol. 1, pp. 58-61.
manent Magnets and their Applications Sergeyev, V. and T.Y. Bulygina, 1970, IEEE,
(Wiley, New York, London). Trans. Magn., MAG-6, 194.
Pashkov, P.P., A.A. Fridman, Ye.B. Granov- Shtrikman, S. and D. Treves, 1959, J. Phys.,
sky, V.V. Sergeyev and R.Ya. Larichkina, Radium, 20, 286.
188 R.A. McCURRIE

Shtrikman, S. and D. Treves, 1966, J. Appl. (Amer. Soc. Metals, Metals Park, Ohio,
Phys. 37, 1103. USA, 1973) 8th ed., vol. 8, pp. 37-47.
Sixtus, K.J., Kronenberg, K.J. and R.K. Tenzer, Vallier, G., C. Bronner and R. Peffen, 1967,
1956, J. Appl. Phys. 27, 1051. Cobalt, No. 34, 10.
Snoek, J.L., 1938, Probleme der Technischen Van der Steeg, M.G. and K.J. de Vos, 1964, Z.
Magnetisierungskurve (Springer, Berlin) p. 73. Angew. Phys. 17, 98.
Snoek, J.L., 1939, Physica, 6, 321. Van Wieringen, J.S. and J.G. Rensen, 1966, Z.
Stfiblein, H., 1963, Tech. Mitt. Krupp, Forsch. Angew. Phys. 21, 69.
Ber. 21, 171. Wittig, R., 1962, Z. Angew. Phys. 14, 248.
Stfiblein, H., 1968, Tech. Mitt. Krupp, Forsch. Wittig, R., 1966, Z. Angew. Phys. 21, 98.
Ber. 26, 1. Wohlfarth, E.P., 1955, Proc. Roy. Soc. Lond.
Steinort, E., E.R. Cronk, S.J. Garvin and H. A232, 208.
Tiderman, 1962, J. Appl. Phys. Suppl. 33, Wright, W., 1970, Cobalt, No. 48, 115.
1310. Wright, W. and R. Ogden, 1964, Cobalt, No.
Stoner, E.C. and E.P. Wohlfarth, 1947, Nature 24, 140.
(Lond.) 160, 650. Wyrwich, W., 1963, Z. Angew. Phys. 15, 263.
Stoner, E.C. and E.P. Wohlfarth, 1948, Phil. Yamada, O., 1955, Z. Phys. 142, 225.
Trans. Roy. Soc. (Lond.) A240, 599. Yermolenko, A.S. and Ya.S. Shur, 1964, Fiz.
Street, R. and J.C. Woolley, 1949, Proc. Phys. Met. Metalloved. 17, No. l, 31 (Engl. Transl.,
Soc. Lond. A62, 562. Phys. Met. Metallog. 17, No. 1, 31).
Street, R. and J.C. Woolley, 1950, Proc. Phys. Yermolenko, A.S., E.N. Melkisheva and Ya.S.
Soc. Lond. B63, 509. Shur, 1964, Fiz. Met. Metalloved. 18, No. 4,
Street, R. and J.C. Woolley, 1956, Proc. Phys. 540 (Engl. Transl., Phys. Met. Metallog. 18,
Soc. Lond. B69, 1189. No. 4, 63).
Street, R., J.C. Woolley and P.B. Smith, 1952a, Yermolenko, A.S. and A.V. Korolyov, 1970,
Proc. Phys. Soc. Lond. B65, 461. IEEE, Trans. Magn., MAG-6, 252.
Street, R., J.C. Woolley and P.B. Smith, 1952b, Zijlstra, H., 1956, J. Appl. Phys. 27, 1249.
Proc. Phys. Soc. Lond. B65, 679. Zijlstra, H., 1960, Doctoral Thesis, University
Sucksmith, W., 1939, Proc. Roy. Soc. Lond. of Amsterdam, The Netherlands.
A171, 525. Zijlstra, H., 1961, J. Appl. Phys. Suppl. 32, 194
Takeuchi, M. and Y. Iwama, 1976, Trans. Jap. S,
Inst. Metals, 17, 489. Zijlstra, H., 1962, Z. Angew. Phys. 14, 251.
Tenzer, R.K. and K.J. Kronenberg, 1958, J. Zijlstra, H., 1966, Z. Angew. Phys. 21, 6.
Appl. Phys. 29, 302. Zingery, W.L., W.B. Whalley, E.B. Romberg
Underwood, E.E., 1 9 7 0 , Quantitative and F.W. Wheeler, 1966, J. Appl. Phys. 37,
Stereology (Addison-Wesley; Reading, Mas- 1101.
sachusetts, Menlo Park, California, London. Zumbusch, W., 1942a, Arch. Eisenhfittenw. 16,
Don Mills, Ontario). 101.
Underwood, E.E., 1973, Applications of Quan- Zumbusch, W., 1942b, Electrotechnik-Mas-
titative Metallography, in: Metals Handbook chinenbau, 60, 522.
chapter 4

OXIDE SPINELS*

S. KRUPI(~KA AND P. NOVAK


Institute of Physics
Czechoslovak Academy of Sciences
Prague
Czechoslovakia

* For accounts of specifically non-microwave and microwave ferrites see chs. 3 and 4, Vol. 2, by P.I.
Slick and J. Nicolas. For an account of garnets, see ch. 1 Vol. 2, by M.A. Gilleo.

Ferromagnetic Materials, Vol. 3


Edited by E.P. Wohlfarth
© North-Holland Publishing Company, 1982

189
CONTENTS

1. C r y s t a l s t r u c t u r e a n d c h e m i s t r y . . . . . . . . . . . . . . . . . . . 191
1.l. D e s c r i p t i o n of t h e c r y s t a l s t r u c t u r e . . . . . . . . . . . . . . . . 191
1.2. C a t i o n s e n t e r i n g t h e o x i d e spinels . . . . . . . . . . . . . . . . 194
1.3. T h e r m o d y n a m i c s t a b i l i t y . . . . . . . . . . . . . . . . . . . 196
1A. C r y s t a l field a n d c o v a l e n c y . . . . . . . . . . . . . . . . . . . 202
1.5. C r y s t a l e n e r g y , c a t i o n d i s t r i b u t i o n a n d site p r e f e r e n c e s . . . . . . . . . 206
1.6. O r d e r i n g a n d d i s t o r t i o n s . . . . . . . . . . . . . . . . . . . 211
2. M a g n e t i c o r d e r i n g . . . . . . . . . . . . . . . . . . . . . . . 216
2.1. E x c h a n g e i n t e r a c t i o n s in spinels . . . . . . . . . . . . . . . . . 217
2.2. M a g n e t i c o r d e r i n g : t h e o r y . . . . . . . . . . . . . . . . . . . 221
2.3. M a g n e t i c o r d e r i n g : e x p e r i m e n t . . . . . . . . . . . . . . . . . 224
2.3.1. Spinels w i t h o n e m a g n e t i c s u b l a t t i c e o n l y . . . . . . . . . . . 224
2.3.2. Spinels w i t h m a g n e t i c i o n s in b o t h s u b l a t t i c e s . . . . . . . . . . 226
2.3.3. T h e effect o f d i a m a g n e t i c s u b s t i t u t i o n s . . . . . . . . . . . . 227
2.3.4. T e m p e r a t u r e a n d field d e p e n d e n c e s . . . . . . . . . . . . . 229
2.4. S p i n w a v e s . . . . . . . . . . . . . . . . . . . . . . . . 231
3. M a g n e t i c p r o p e r t i e s . . . . . . . . . . . . . . . . . . . . . . . 233
3.1. A n i s o t r o p y a n d m a g n e t o s t r i c t i o n . . . . . . . . . . . . . . . . . 233
3.1.1. I n t r o d u c t o r y r e m a r k s . . . . . . . . . . . . . . . . . . 233
3.1.2. M i c r o s c o p i c origin: a n i s o t r o p y . . . . . . . . . . . . . . . 235
3.1.3. M i c r o s c o p i c origin: m a g n e t o s t r i c t i o n . . . . . . . . . . . . . 238
3.2. M a g n e t i c a n n e a l a n d r e l a t e d p h e n o m e n a . . . . . . . . . . . . . . 245
3.2.1. O r i g i n . . . . . . . . . . . . . . . . . . . . . . . 245
3.2.2. E x a m p l e s o f local a n i s o t r o p i c c o n f i g u r a t i o n s . . . . . . . . . . 247
3.2.3. K i n e t i c s , m a g n e t i c a f t e r - e f f e c t s . . . . . . . . . . . . . . . 249
3.2.4. S u r v e y of e x p e r i m e n t a l r e s u l t s . . . . . . . . . . . . . . . 251
3.3. D y n a m i c s of m a g n e t i z a t i o n . . . . . . . . . . . . . . . . . . 256
4. O t h e r p h y s i c a l p r o p e r t i e s . . . . . . . . . . . . . . . . . . . . . 260
4.1. E l e c t r i c a l p r o p e r t i e s . . . . . . . . . . . . . . . . . . . . . 26O
4.1.1. D C e l e c t r i c a l c o n d u c t i v i t y a n d S e e b e c k effect . . . . . . . . . . 261
4.1.2. D i e l e c t r i c c o n s t a n t . . . . . . . . . . . . . . . . . . . 275
4.1.3. O p t i c a l a n d m a g n e t o o p t i c a l s p e c t r a . . . . . . . . . . . . . 276
4.2. M e c h a n i c a l a n d t h e r m a l p r o p e r t i e s . . . . . . . . . . . . . . . . 284
4.2.1. I n f r a r e d s p e c t r a . . . . . . . . . . . . . . . . . . . . 284
4.2.2. E l a s t i c c o n s t a n t s . . . . . . . . . . . . . . . . . . . . 285
4.2.3. H e a t c a p a c i t y . . . . . . . . . . . . . . . . . . . . . 287
4.2.4. T h e r m a l c o n d u c t i v i t y . . . . . . . . . . . . . . . . . . 288
4.2.5. T h e r m a l e x p a n s i o n . . . . . . . . . . . . . . . . . . . 289
Appendix: Intrinsic magnetic properties . . . . . . . . . . . . . . . . . 291
References . . . . . . . . . . . . . . . . . . . . . . . . . . . 298

190
1. Crystal structure and chemistry

The spinel structure is one of the most frequently encountered with the MM~X4
compounds. X represents oxygen or some Chalcogenic bivalent anion (S2-, Se 2 ,
Te ~ ) which may be partly substituted by proper monovalent anions (F , I-, Br-);
M, M' are metallic ions (or a combination of them) whose valencies have to fulfil
the electroneutrality requirements.
Due to the large electronegativity of oxygen the ionic type of bonds prevails in
almost all oxide spinels. As a consequence, the electrical resistivity is usually high
and we are allowed to classify these compounds as insulators even though
sometimes the term low-mobility semiconductors would apply. This is not the case
with the other bivalent anions (S, Se, Te) possessing considerable lower elec-
tronegativities. They behave more like semiconductors or even metals and,
therefore, they are treated separately (see Van Stapele, ch. 8).

1.1. Description of the crystal structure

The spinel structure is called after the mineral spinel MgAI204. It is formed by a
nearly close-packed fcc array of anions with holes partly filled by the cations. There
are two kinds of holes differing in coordination: tetrahedral (or A) and octahedral (or
B). From all these sites available in the elementary cube containing 32 anions, only 8
of the A-type and 16 of the B-type are occupied by cations. The geometry of the
occupied interstices may be seen from fig. 1, where the primitive cell containing 2
formula units MM~X4 is shown. The symmetry is cubic and corresponds to the space
group O7(Fd3m). A small displacement, defined by a single parameter u, of the
anions from their ideal position is allowed along the corresponding body diagonal
(fig. 2) which enables a better matching of the anion positions to the relative radii of A
and B cations.
The coordinates of ions in spinel are summarized in table 1. For the ideal
close-packed anion lattice u equals 2. In real oxide spinels usually u ¢ ~ and in the
model of hard spheres it should increase linearly with (rA-- rB)/a (fig. 3). The local
symmetry of the cation sites is cubic in the A positions and trigonal in B, the
trigonal axis being one of the body diagonals (fig. 1); note that the trigonal
symmetry is due to both the configuration of neighbouring cations and the
distortion of the anion octahedron if u ¢ ~. Each of the four body diagonals

191
192 S. KRUPI(~KA A N D P. N O V ~ K

/ I // I /
// I ,~"/X " / I / /

~C" l~ f,'" t"-",_,tT/'~ I

a
2 z ~

I I I i~. ~ p r I
r I t I I I I
I I I I I I r
j_L___T___~ia___ . /a . . . . /
I

~_C . . . . . . . . 2.£ . . . . . . . . . ~

L/ 2-

@ A

0 B
Fig. 1. Primitive cell of the spinel structure.

TABLE 1
Coordinates of ions in the elementary cell of spinel structure (in fractions of the lattice
parameter ).
i I 1
8 ions in A-sites 0, 0, 0; ~, ~,

16 i o n s in B - s i t e s _5 5 5~; 5L ~,
8, ~, 7 ~;
7 L7 5g, g;
7 ~,
7 ~,
7 85-

1 1 1 1 1 1
32ionsX u, u, u; u, ~, ~; 8 - u, ~ - u , ~ - u; ~ - u, ~ + u , ~ + u
1 1 1 + U~ 1 ~_}_1 U~ i
~, U, t2; t~, ~, U; ~ + U, ~-- U, ~+U, ~--U

Coordinates of remaining ions are obtained by translations (0, ½,1), (½, O, ½), (½,1~, 0).

ah, /

5 Ir - - - ~,
,I
a(u-~)I3" , -.
~ i 1 ~
1 I i I ~

Q_
4-
Fig. 2. Environment of the anion in spinel.
O X I D E SPINELS 193

2-3 type 4-2 type


• Go 3. 13 Ge 4.
+ AL 3. o .,or)4*
+
0.39
• V 3÷

• C r 2÷

. A Fe 3~


0.38

0.37
Na2~O~

Ag2~o04

0.36
I I I I I I I / I
-0.06 -0.04 -0.02 o ~-~
(2

Fig. 3. D e p e n d e n c e of the p a r a m e t e r u on t h e reduced difference of radii of A and B ions.

belongs to just one of the B cations in the primitive cell in fig. 1. Hence, these
cations are non-equivalent differing in their local symmetry axis and each of them
may be taken as representing one fcc sublattice with the lattice constant a. On the
other hand, the local symmetry of the A positions remains cubic even if u # 3, so
that both sublattices represented by A sites in the primitive cell are mutually
equivalent. When considering some aspects for which the local symmetry is
irrelevant, all the B positions may be treated as belonging to only one sublattice
(octahedral or B) in the same way as the A positions may be unified to form one
tetrahedral sublattice (A).
Both translational and local symmetries corresponding to the O 7 space group
strictly apply only if each sublattice contains only one kind of cations, i.e., if all M
ions in MM~X4 are in tetrahedral and all M' in octahedral positions. The spinel is
then called normal. It was early established by X-ray diffraction (Barth and
Posnjak 1932), and later by neutron diffraction (Hastings and Corliss 1953) as
well, that besides this, another cation distribution exists in many spinels called
inverse: here, one half of the cations M' are in A positions and the rest, together
with the M ions are randomly distributed among the B positions. There are also
many examples of intermediate cases between a normal and an inverse spinel
(mixed spinels). Therefore, in order to characterize fully the spinel structure, a
further parameter is needed describing the degree of inversion. The chemical
194 S. KRUPICKA AND P. NOVAK

formula may then be explicitly written as

MaM~-~[MI-~M~+a]X4, (1)

where the cations on B sites are in brackets. Note that 6 = 1 means a normal
spinel, 6 = 0 an inverse one.
If there are different cations coexisting in equivalent interstices (partial or full
inversion, solid solutions) the symmetry is perturbe~d. Nevertheless, new symmetry
relations may appear when the different cations order regularly (section 1.6). The
symmetry of the spinel structure may also be changed by a spontaneous distortion
due to the cooperative Jahn-Teller effect (section 1.6). It has been pointed out on
the basis of far infrared spectra and calorimetric measurements (Grimes 1972,
1974, Grimes et al. 1978) that in some spinels the octahedrally coordinated cations
might be shifted a little out of their central positions ("off centre ions") which
would also change the space group. The attempts to prove this conclusion by a
direct diffraction study have given controversial results until now (Rouse et al.
1976, Hwang et al. 1973, Thompson and Grimes 1977).

1.2. Cations entering the oxide spinels

The,electroneutrality considerations lead to 3 basic spinel types, according to the


cation valency combinations in MM~O4:M2+M~3+O4 2- (2-3 spinels), M4+M~2+O2-
(4-2 spinels) and M6+M~+O 2- (6-1 spinels). Other possible types may be deduced
from these by formal substitutions (e.g. Nf+~½(M"> + M '3+ giving ~,,l+~,3+c~2-
Iv-to.5 -tvx2.5 '~.J4 ,

M2+~[]~/3 .± . .~.4 12/3


3+
leading to so-called y - M ~ O 3 defect sesquioxide phase, etc.).
The basic binary spinel types are listed in table 2a together with relevant cations
and their ionic radii. Additional spinel types derived by substitutions of the basic
ones are listed in table 2b. It is seen from tables 2a and 2b that practically any
cation with radius within the limits 0.4 to = 1 ~ may be incorporated into the
spinel structure and most of them can occur in both octahedral and tetrahedral
positions. The smallest cations with valency />4, however, are found in the
tetrahedral coordination only, while the large monovalent cations occurring main-
ly in 6-1 spinels are confined to the octahedral sites. Besides the geometrical factors,
the distribution of cations among A and B positions is influenced by many others,
as briefly discussed in section 1.5.
The interstices available for cations in the spinel structure have radii

RA = (u - ~)aX/5- r(O>); R , = (~- u ) a - r(O2-). (2)

In table 3, these are compared with corresponding ionic radii as listed in table 2,
using the experimental values for a and u. The agreement is satisfactory though a
small but rather systematic contraction of the ionic radii, up to a few percent, is
observed. On the other hand, a comparison of experimental a and u values with
the calculated ones based on the eq. (1) (table 4), where the ionic radii from table
2 are inserted for RA, RB, shows that this contraction concerns the lattice
O X I D E SP1NELS 195

J
oo

c5
+
• eq + •

.x. ~ ~ ~ ~
0 ..<N
"0
c~ o
~+
c5 ~ c 5~
~ + + ~
+ e~ + r-I
+ r~ ~'-~
t~

,..<

2 ~- + +
=L
o~ O

t"q c5
+~

"'.c~ +
r~
~h
~ r-- oq
O
+ +
~" ea t~

+
O 3-
+ oz-
O
,'6
©

c5 ~5 E + +~
m ~5 e~ c'4
..~
O~ ©
+
% c5
N~ "7
6c O

g. q~
m E

O.,
y, O
196 S. KRUPICKA AND P. NOV.~A~

TABLE 3
Radii of A and B interstices, R, and corresponding ionic radii RM (all in A).

A B

Spinel R RM R Rra

ZnFezO4 0.715 0.74 0.762 0.785


NiFe204 0.632 0.63 0.775 0.808
MnFe204 0.722 0.766 0.783 0.785
Fe304 0.616 0.63 0.801 0.853
MnCrzO4 0.77 0.80 0.732 0.755
FeV204 0.731 0.77 0.76 0.78
CoFe204 0.653 0.64 0.778 0.787
Mg2TiO4 0.75 0.71 0.745 0.773
Na2WO4 0.62 0.56 1.08 1.16

TABLE 4
Comparison of experimental values of a and u with those calculated using
eqs. (2) and table 2.

Lattice o
parameter (A) u

Spinel Distribution of ions exp. calc. exp. calc.

ZnFe204 ZnZ+[Fe3+] 8.433 8.533 0.3852 0.3853


MnFe204 Mn2+ Fe3+
0.82 rMn3+
oast 0.18Fe2+
0.18Fe3+ 1
a.641 8.50 8.611 0.3848 0.3861
3+ 2+ 2+ 3+
CoFe204 Fe0.ssCo0.12[Co0.88Fea.12] 8.38 8.384 0.3818 0.3808
NiFe204 Fe3+[Ni2+Fe3+] 8.33 8.423 0.3822 0.380
ZnCr204 Zn2+[Cr~+] 8.321 8.453 0.3863 0.3866
MnCr204 Mn2+[Cr~+] 8.437 8.545 0.3889 0.3892
Fe304 Fe3+[Fe2+Fe3+] 8.394 8.543 0.379 0.3777
Co2GeO4 Ge4+[Co~+] 8.3175 8.223 0.375 0.3757
Mg2TiO4 Mg2+[Mg2+Ti4+] 8.441 8.453 0.3875 0.3846
Na2WO4 W6+[Na~] 9.1297 9.255 0.369 0.3635

p a r a m e t e r a a n d h e n c e the whole structure, leaving u u n c h a n g e d . A n overall view


of the g e o m e t r i c a l relations in various types of spinels m a y be acquired from fig.
3; the e x p e r i m e n t a l u values follow quite well the linear d e p e n d e n c e on the
r e d u c e d difference of ionic radii ( r ( B ) - r(A))/a.

1.3. Thermodynamic stability

T h e oxide spinels m a y usually be p r e p a r e d at e l e v a t e d t e m p e r a t u r e s by a direct


solid state r e a c t i o n b e t w e e n the simple oxides. T h e r e l e v a n t t e m p e r a t u r e r a n g e is
a b o u t 800 to 1500°C, d e p e n d i n g o n the type of cations. T h e t h e r m o d y n a m i c
stability of spinels c o m p a r e d to the simple oxides is given by the G i b b s free
OXIDE SPINELS 197

e n e r g y of f o r m a t i o n A G c o n n e c t e d with t h e r e a c t i o n

M O + M~O3 ---)MM~O4. (3)

B y c o m b i n i n g A G with t h e e n t h a l p y of f o r m a t i o n AH, t h e e n t r o p y c h a n g e A S
c o r r e s p o n d i n g to r e a c t i o n (3) m a y b e e s t i m a t e d . I n s t e a d of A G , A H d a t a a r e
s o m e t i m e s u s e d w h e n c o m p a r i n g t h e r m o d y n a m i c stability for v a r i o u s spinels. T h e
values of A G , A H a n d A S for s o m e s e l e c t e d spinels a r e given in t a b l e 5 a n d t h e
A H ' s for 2 - 3 spinels a r e d i s p l a y e d in fig. 4 as d e p e n d i n g on t h e t y p e of d i v a l e n t
cation.
T h e r e l a t i v e l y g r e a t stability of o x i d e spinels is c o n n e c t e d with t h e s t r o n g l y ionic
c h a r a c t e r of t h e b o n d . T h e spinels with a n i o n s possessing s m a l l e r e l e c t r o -
n e g a t i v i t i e s t h a n o x y g e n (S, Se, T e ) a r e m o r e c o v a l e n t a n d b e c o m e also less stable.
This m a n i f e s t s itself in a d e c r e a s e d n u m b e r of such spinels, t h e i r s m a l l e r m u t u a l
solubility a n d m o r e r e s t r i c t e d n u m b e r of c a t i o n s e n t e r i n g t h e s t r u c t u r e . T h e r e is
s t r o n g c o m p e t i t i o n b e t w e e n spinel and, e.g., n i c k e l a r s e n i d e s t r u c t u r e , a n d s o m e
s u l p h o s p i n e l s can b e t r a n s f o r m e d i n t o t h e l a t t e r u n d e r high p r e s s u r e ( B o u c h a r d
1967). T h e o x i d e spinels, on t h e o t h e r h a n d , f o r m solid s o l u t i o n s a l m o s t w i t h o u t

TABLE 5
AG, AH and AS values for selected spinels; after Navrotsky
and Kleppa (1968).

Spinel AG* AH(970 K)t AS:~

1273 K: -8.4
MgAI204 1673 K: -8.4 -8.72 ± 0.29 0

MgFe204 1273 K: -5.3 -4.93 - 0.24 +0.3

Fe304 1000 K: -7.5 1000 K: -5.1 +2.2

1273 K: -9.4
CoFe204 1473 K: -8.2 -5.89 + 0.21 +2.7

NiFe204 1273 K: -6.0 - 1.22 ± 0.22 +3.8

ZnFe204 1400 K: -6.5 -2.67 ± 0.23 +2.7

FezTiO4 1073 K: -7.2

Co2TiO4 1073 K: -6.7 -6.04 ± 0.19 +0.7

Co2GeO4 - 14.1 ± 0.4

* Equilibrium measurements.
t Calorimetric measurements.
:~High temperature values calculated at the temperature
for which AG is taken.
198 S. KRUPI(~KA AND P. NOV~id{

aH
i i i

/-,'<cal 7
L m-~r_/ o
-~4 I I \\ /I I
i \ i t [
i x / I
I I \~1 I I

-12 i~/." ~ ]
o ,", , !
I \ I I
-I0 ~ ~ ~ ]
I \ I ]

-6 • i Ill Xi\\\ .~'l'~ll ll' fill?


I -+ - - - --~ ~ "%% \ It %1 I ] II
4---1"------ \ [] I
I \ t t t I I/
-4. / '~. ,,', , ,/I
/ \ ~ 'l I ii
\\ .l'i ~ I If

o , , , I ",t\1,~ ',,t I !,i'~ I,"",.


~,, , ~, ,
I "\
,,
I
\. t I1 / \

2
• MAt20` ", ".~' ,'
4 + MFe204 ~ f'~t,
A MMn204 ~
Mg 2+ Mn 2+ Fe 2. Co 2+ Ni 2` Cu 2+ ZD 2+ C 2+

Fig. 4. Enthalpy of formation AH for 2-3 spinels (Navrotsky and Kleppa 1968, Jacob and Walder-
raman 1977).

restriction and the competing structures are of other t y p e s - i n the first place
olivine and phenacite structures. They usually appear if one of the cations is very
small (Si4+, Ge4+). The situation may be elucidated on the basis of the hypothetical
phase diagram in fig. 5. We may infer from it that under normal conditions all
germanates are spinels except Zn2GeO4 having the phenacite structure, while the
silicates are olivines. Most of the latter ones can be transformed into spinels under
high pressure and/or elevated temperatures. Both olivine and phenacite structures
are less densely packed compared with spinels, the typical volume change being
- 8 % and +19% for the olivine--> spinel and olivine-->phenacite transformation,
respectively. The thermodynamic relations among olivine, spinel and phenacite
structures in silicates and germanates are discussed by Navrotsky (1973a, b, 1974,
1975), Navrotsky and Hughes (1976) and Reinen (1968). The AG values for both
groups of oxides are shown in fig. 6.
With large cations, e.g., large divalent cations (Ca 2*, Ba 2+, etc.) in M2+M~>O4,
the spinel structure becomes also unstable and other structures (hexagonal)
appear. The mutual relations to spinel have not yet been established.
Hitherto no attention was paid to the chemical stability of the oxide spinels
against the oxidizing or reducing influence of the atmosphere. This may be
justified if the incorporated cations have fixed valencies that cannot be changed
OXIDE SPINELS 199

F . . . . . .

i~Geq
[co~Geq
jn2Geq I ,Fe~SJO~ iI A+~' '
L . . . . . •

. . . . . 1 ' N~S~Oj ' Z ISjO ;


, , ylic%j ',
,In92vq ', l'J I'---~---~' '
, f ~ ~ ',n~2sJq',
. . . . . J L . . . . . 3
I

L. . . . "

F . . . . . 31

z~Geq!
L . . . . . A

Temperature
Fig. 5. Phase relations between spinel, olivine and phenacite; after Navrotsky (1973).

i i i i , ,

Hg
16

14

~9

~8

I I I I I
-I0 -8 -6 -4 -2 0
Germanates a G ~kcaL/moLJ
Fig. 6. Gibbs free energy of transformation olivine-spinel in germanates and silicates; after Navrotsky
and Hughes (1976).
200 S. KRUPICKA AND P. NOV/~d~

TABLE 6
Valencies and ground states of 3d" ions.

Number of
3d electrons
Valency 0 1 2 3 4 5 6 7 8 9 10

1+ Cu
2+ Ti V Cr Mn Fe Co Ni Cu Zn
3+ Sc Ti V Cr Mn Fe Co Ni
4+ Ti V Cr Mn Fe Co
5+ V

Ground term 1S 2D 3F 4F 69 58 5D 4F 3F 2D IS

within wide limits of the oxygen pressure. In the case of transition metal ions,
however, this certainly is not a plausible approach, because these ions usually
exist in several valency states (see tables 2a and 6). As a consequence the
thermodynamic stability has then to be considered for both the crystal and
gaseous phases in mutual contact. This leads to definite deviations from chemical
stoichiometry of the ideal spinel depending on both the temperature T and the
partial oxygen pressure Poz. If this deviation exceeds some critical value (depend-
ing on T and po~) the spinel structure may become unstable. The situation is
illustrated in figs. 7 and 8. Note that in air the transition metal oxides with
spinel structure melt often incongruently, which makes it difficult to prepare single
crystals of an arbitrary composition from their own melt. Also the fact that at a
given temperature the spinel of a definite composition is in thermodynamical

~~'~4~ hematite[I
i o.o 5 3

600 1400 1200 1000


,pc]
Fig. 7. Phase diagram for system Fe-O (Tretyakov 1967); projection into temperatureiequilibrium
pressure of oxygen plane is shown. The heavy lines define the borders of the stability regions of the
respective phases. The thin lines are the lines of constant 3' in the formula Fe304+~.
OXIDE SP1NELS 201

•~ ÷ •~ ~0~
.~<~ {~. ~ - ~
'~ ~ '~t'-I ~ ~ ~ ~ Z- .~

e- i ~ / \ ~ "~ ~,<~
•. /, <~ s / \ ,~ ~. ~

~'~o~
0 0

<:)
~'~
i
(=3 o .~ ~ ~ ' ~
(:~ i::::) (:b ~ i:::) C) (::) ~ ~ <:)
e~l e',,l ~ ~ co N'- ¢M <::)
0",1 ¢',1 . . . . .

l::i

~ ~ ~ ~

i,-i

<~L~ ~
?
! / i / /111 ~ ~ i

<b ~ ~ I '" " i " ~ --~ ! ' # i ~ LiP 11# ''~


~ ~ - - 7 ~ , --"P-~ , ~L_L_ ~ ' ~.~b.¢',/9 , '

~-+---~<..'~.,~.,.--C7~o~ '< ~, ' ,', ,' ,' l i V ~,:'~.~,"r. 0

, • 21) ~ ~1 I I i '; %
"~ :A<<~-~"~, . , , , ,.x , ~ ! ~ Z : / ' 7 \: I
i.~ - • "Ill '% -- ~--o ~ i ii [
o~ A. " o:~:'~"~ - i ' o ' i ,.s~ll : - ~ ,\ , . ,
o~ " ~1o,...> ~ -< ix. ,\,
,\, ~x ~\,
">~z / \ i "-/o~ ..+ "v-Gb% I - , . ; ; \ " "; PA' %
o~ " * _ ~ , ~." <::> ~ ' - < \ " " i . ~ ,\ ,~ © 0
.#. *\ , c< ..,. "..~.~<..\, " < \ ~r,~
o°O~ ,' " , ~ ; ~* ~ ~ , \ ~,\:\I\
°O~<z .__ ' ~ , "~. '~ ' ~ "-l~'i 9--'

0
.-~
l.i.. il [ • ~
£
o
Q:: ~.. J ~J_
ba

.
202 S. KRUPICKA AND P. NOV~d,~

equilibrium only if the partial oxygen pressure has a definite value, is very
important for the production of these materials. On the other hand, when cooling
the material rapidly or under equilibrium conditions in a controlled atmosphere to
room temperature (or more precisely to temperatures at which the ionic diffusion
is ineffective) we can preserve the original spinel structure even at this tem-
perature under non-equilibrium condition, e.g. in air.

1.4. Crystal field and covalency

With respect to the magnetic properties the interest is primarily in transition


metal ions, particularly those of 3d" group (table 6). The outer d-electrons of
these ions may be regarded as practically localized in almost all oxide spinels so
that the crystal (or better ligand) field theory applies. From the energy spectra
determined by this theory the low lying levels are decisive for the magnetic
behaviour. Note that the origin of the ligand field splitting of levels is to be seen in
both the electrostatic crystal field and the covalency between the cation and the
surrounding anions (ligands). Both these effects also contribute to the stabilization
of cations in the given surrounding (crystal or ligand field stabilization energy =
lowering of the ground level with respect to the ground level of the free ion).
Although the local symmetry is trigonal in octahedral sites (point group D3d)
and cubic in tetrahedral ones (point group Td), the prevailing part of the ligand
field in both cases is cubic (point group Oh and Td, respectively). As a rule, the
cubic ligand field strength, which is characterized by the cubic field splitting
parameter A = 10Dq, corresponds to the so-called intermediate ligand field: it is
weaker than the correlation between equivalent d-electrons but definitely stronger
than the spin orbit coupling, i.e.,

/~correl > A > AL_ S .

The only exceptions exhibiting characteristics of the "strong" crystal field (low
spin state) are Co 3+, possibly Ni 3+ and the ions from the 4d n group.
The splittings of the ground state terms of ions d" in the intermediate octa-
hedral cubic fields are shown in fig. 9; in the tetrahedral case the sequence of
levels is reversed which corresponds to A < 0. In table 7 the experimental data
on the splitting parameter A in both B and A positions for various ions and
spinels are summarized; for comparison and/or completion the data of some other
oxides are included.
The trigonal field in the octahedral positions leaves the orbital doublet in fig. 9
unsplit while the triplets T are split into a doublet and singlet. It depends on the
type of ion and possibly, on other circumstances which of them is lower. The
singlet is usually lower for Fe 2+ and the doublet for Co 2+. Note that this does not
automatically apply to other structures, e.g. garnets. The splitting by the trigonal
component of the ligand field in spinels is about one order of magnitude smaller
than that in cubic field. Unfortunately, there are only few experimental data
available; the examples are given in table 8. Sometimes they have been estimated
OXIDE SPINELS 203

i 6Dq ' 4Dq 7"29(3)


I

-4Dq -6Dq
k i T2y(3) *I
d ~and d~ E9(2)
d~and d 9

A~9(1)

~9(3)

t-I
/ 12Oq
~ 2Dq 729(3) t -I i 5Dq

Ii -6Dq I
~g(3)
* T~9(3) i - 12Dq
d2and d 7 v A2g(O
d 3 and d 8

Fig. 9. Splitting of the ground multiplets of 3d" ions in octahedral field. The orbital degeneracy is
given in parenthesis.

in a more or less indirect way, e.g., using the temperature dependence of the
nuclear quadrupole splitting which, for ions with L > 0 , reflects the spatial
distribution of the electronic cloud (EibschLitz et al. 1966).
In the presence of the Jahn-Teller ions (see section 1.6) the local symmetry may
be tetragonal in both types of sites. This symmetry splits both the E doublets and
T triplets. The latter is split similarly as in a trigonal field, i.e., into a doublet and
a singlet. If the symmetry is orthorhombic, the T level is additionally split so that
only singlets appear. Note that due to partial disorder (inverse spinels, solid
solutions) the local fields often possess fluctuating components with lower sym-
metries.
In early work on d-level splitting in insulators only the effect of electrostatic
crystal field was considered. This leads among others to the conservation of the
centre of gravity of the split d-levels which makes it easy to determine the crystal
field stabilization in terms of 10 Dq (Dunitz and Orgel 1957, McClure 1957). This
was frequently used in evaluating the relative preferences of cations for A and B
positions in spinels (section 1.5). It appeared, however, that this simple picture is
inadequate because it entirely neglects the covalency whose effect has been shown
to be of comparable magnitude. Due to covalent mixing of states the centre of
gravity of 3d levels is generally shifted upwards, while the energy of the (p, s)
valency band is lowered (fig. 10). As a consequence, a contribution to the
stabilization energy appears which is difficult to estimate (Goodenough 1963,
Blasse 1964).
204 S. KRUPICKA A N D P. NOV/kK

e-, ..,.d
© ~ cq

t.-,

.t::l II II

~D

° °
~', t"q
t'q ¢q ¢q

,..., II II

-'T
I I I I I I I I [
E I I

©
v~ ©.--,

.% +
+ + + ~ +

~D

%
re) 0"5
o
OXIDE SPINELS 205

[~ ~ ~.~

t"q

r.i dd

Z
~5 ~ .~ ,~ ~5

t"¢3 r¢3
206 S. KRUPICKA AND P. NOV~J(

TABLE 8
Trigonal field splitting of energy levels of Cr 3+ and Fe 2+ ions.
Parameters v, v' are defined in the usual way: v =
(t2gA] Vtrigltzga) - (t2gE[Vtriglt2gE); I)' = (t2gE] VtriglegE). The split-
ting of the ground state triplet T2g of Fe 2+ is ~v, while for Cr3+ the
splitting of 4T 1 is ~'~'I/9+ /)t and that of 47"2is -~v/2 (excited states).

Ion Spinel o (cm-1) v' (crn 1) Ref.

Cr3+ MgA1204 - 200 - 1700 1


Cr 3+ ZnGa204 -650 -1100 2
Cr3+ Li0.sGazsO4 -400 2400 3
Fe z+ GeFezO4 - 1145 4

1. Wood, D.L. et al., 1968, J. Chem. Phys. 48, 5255.


2. Kahan, M.H. and R.M. Macfarlane, 1971, J. Chem. Phys. 54,
5197.
3. Szymczak, H. et al., 1975, J. Phys. C8, 3937.
4. Eibsch/itz, H. et al., 1966, Phys. Rev. 151, 245.

2p
2s

without ~¢ith
covaLency covatenoy

Fig. 10. Effect of the covalency on the energy levels of transition metal oxide (schematically).

A p r o m i s i n g d i r e c t m e t h o d f o r d e t e r m i n i n g t h e r e l a t i v e p o s i t i o n s of t h e d l e v e l s
in A a n d B sites of f e r r i m a g n e t i c spinels s e e m s to b e t h e p h o t o e m i s s i o n m e t h o d
c o m b i n e d w i t h t h e m e a s u r e m e n t of e l e c t r o n spin p o l a r i z a t i o n ( A l v a r a d o et al.
1975, 1977).

1.5. C r y s t a l energy, c a t i o n distribution a n d site p r e f e r e n c e s

T h e far l a r g e s t c o n t r i b u t i o n to t h e crystal e n e r g y in o x i d e spinels is t h e C o u l o m b


e n e r g y of t h e c h a r g e d i o n s ( M a d e l u n g e n e r g y ) :
OXIDE SP1NELS 207

Ec = -(e2/a)AM, (4)

where e is the charge of electron, a the lattice p a r a m e t e r and AM the Madelung


constant. AM m a y be expressed as a function of the mean electric charge qA of the
cations in A positions and of the oxygen p a r a m e t e r u. It was calculated by several
authors (De Boer et al. 1950, Gorter 1954, D e l o r m e 1958, T h o m p s o n and Grimes
1977b) with slightly differing results. H e r e we give the formula based on the
generalized Ewald method used by T h o m p s o n and Grimes (1977b):

AM = AM(qA, U) = 139.8 + 1186A, -- 648332,


--(10.82 + 412.2A, - 1903A ])qA + 2.609q~, (5)
where au = u - 0 . 3 7 5 . The dependence of AM on qA for different values of the
oxygen p a r a m e t e r u is given in fig. 11. With increasing AM the stability of the
spinel increases. Therefore, owing to its dependence on qA, the Coulomb energy
will generally play an important role in the equilibrium distribution of cations
among A and B positions, even though in some cases other energy contributions
may become important. The Born repulsion energy is difficult to estimate in a
direct way and it is usually deduced from the simple oxide data (Miller 1959). The
polarization energy appears as a consequence of the deformation of the spherical
electron cloud of ions in the local electric field in the crystal. A t t e m p t s to calculate
it in the classical way lead apparently to an overestimation (Smit et al. 1962, Smit
1968), so that only qualitative conclusions are usually used (Goodenough 1963). In
the quantum mechanical picture it is difficult to distinguish this effect f r o m

o~
138LL o 0.3F5
\pS \ i 0.3F8
0.381

134 - " 0,384


0.38F
0,390

130 -

2 %
Fig. 11. Dependence of Madelung constant on the average ionic charge qA of A-site ions for several
values of the parameter u.
208 S. KRUPICKA AND P. NOV.Ad(

covalency. The last relevant contribution is the ligand field energy treated in the
preceding paragraph.
As already mentioned in section 1.2, the spinels may have various degrees of
inversion, according to the formula

MaM~-a [M>aM~+a]04.

If the energy difference for two limiting cases 6 = 0, 8 = 1 is not very large, we
expect the distribution of cations to be random at high temperatures (i.e. 6 = ½)
due to the prevailing influence of the entropy term - T S in the free energy. When
the temperature is lowered, the spinel tends to be more or less normal or inverse
depending on the sign and amount of energy connected with the interchange of
cations M, M' in different sublattices. The equilibrium distribution is determined
by the minimum of the Gibbs free energy, i.e.,

dG dH dS
d6 - d ~ T~=0. (6)

If we restrict ourselves to the configurational entropy of cations and supposing


total randomization in both sublattices, S may be approximated by

S = Nk[-8 In 8 + 2 ( 6 - 1) l n ( 1 - 8 ) - (8 + 1) 1n(6 + 1)]. (7)

Defining further AP = d H / d 6 we find

8(1 + 6) _ exp(-AP/RT) (8)


(l - 8) 2

which determines the equilibrium value 6 at the temperature T. Generally, Ap


depends on 8 and frequently a linear expression

A p = 14o + 1-116 (9)

is being used to describe the experimental results (fig. 12). Here, H0 and H0 +/-/1
may be interpreted as energies connected with the interchange of ions M, M' from
different sublattices in the case of completely inverse and normal distribution,
respectively. It may be shown that the linear dependence (9) is obtained when the
short-range interactions between pairs of ions are considered (Driesens 1968).
Sometimes, an entropy term - T S o is added to the exponent in eq. (8) (Driesens
1968, Reznickiy 1977). The representative values of /-/0, H1 are given in table 9.
When Ap ~< 5 kcal/mol, a mixed type spinel is usually observed; otherwise the
energy difference between normal and inverse structure is sufficient to attain the
one o r the other in practically pure form. The achievement of the equilibrium
depends on the rate of cation migration. As this is a thermally activated process,
the time constant changes exponentially with temperature and if T ~< 500 K, the
OXIDE SPINEI~ 209

o.J
+8 o
o
Ng Fe20~ o ~ J f

0 ~
0.2

I I
FO0 900 1100 13~00 r ~
Fig. 12. Dependence of the degree of inversion 6 on temperature. Data for MgFeaO4 are taken from
Pauthenet and Bochirol (1951) (A), Kriessman and Harrison (1956) (rT), Mozzi and Paladino (1963) (©)
and Faller and Birchenall (1970) (+), those for MnFe204 are from Jirfik and Vratislav (1974). The full
lines correspond to formulae (8) and (9) with Ho, H1 given in table 9. For MnFe204 also the time t,
necessary for establishing the equilibrium distribution of ions at given temperature, is shown.

•m i g r a t i o n b e c o m e s t o o slow to allow any o b s e r v a b l e c h a n g e in cation d i s t r i b u t i o n .


S o m e t i m e s , t h e t e m p e r a t u r e d e p e n d e n c e of t h e m i g r a t i o n r a t e is so s t e e p that,
d u e to t e c h n i c a l r e a s o n s , t h e 6 v a l u e s can b e v a r i e d o n l y in n a r r o w limits. This is
t h e case, e.g., of t h e M n - F e spinels (Simgovfi a n d Simga 1974).
T h e ~ v a l u e s a r e usually d e t e r m i n e d by diffraction m e t h o d s ( B e r t a u t 1951, Jirfik
a n d V r a t i s l a v 1974), m e a s u r e m e n t s of t h e s p o n t a n e o u s m a g n e t i z a t i o n ( P a u t h e n e t
a n d B o c h i r o l 1951), o r by M 6 s s b a u e r effect ( S a w a t z k i et al. 1969). It is i m p o r t a n t
to r e a l i z e t h a t t h e e n e r g y A P c o n c e r n s t h e crystal as a whole. T h e m a i n
c o n t r i b u t i o n s to A P c o m e f r o m t h e M a d e l u n g e n e r g y (4), B o r n r e p u l s i o n e n e r g y ,
210 S. KRUPICKA AND P. NOV~K

TABLE 9
Values of H0 and H1 for some spinels.

Spinel T (K) g /40 (kcal/mol) /-/1 (kcal/mol) Ref.

MgFe204 573-1473 0.1 < 6 < 0.28 3.83 -8.63 1


MnFe204 603-1443 0.763< 6 < 0.94 0.4 -8.1 2
MgMn204 T < 1050 0.78 < 6 < 0.99 10.3 - 18.4 3
NiMn204 298-1213 0.74 < 6 < 0.93 3.64 -9.6 4

1. Average value of /40, //1 taken from Pauthenet and Bochirol (1951), Kriessman and Harrison
(1956), Mozzi and Palladino (1963).
2. Jirfik, Z. and S. Vratislav, 1974, Czech. J. Phys. B24, 642.
3. Manaila, R. and P. Pausescu, 1965, Phys. Status Solidi 9, 385.
4. Boucher, B., R. Buhl and M. Perrin, 1969, Acta Crystallogr., B25, 2326.

and further, from polarization and ligand field effects. T h e attempts to explain the
observed degree of inversion on the basis of only the M a d e l u n g energy have not
been successful. T h e only conclusion we m a y draw on the basis of fig. 11 is a
prediction that the stability of the normal structure increases with increasing u for
2-3 spinels and vice versa for 4 - 2 spinels. This m e a n s that the 2-3 spinels with
largest bivalent cations (Mn 2+, Cd2+), as well as 4 - 2 spinels with smallest
tetravalent cations (Si 4+, G e 4+) are expected to be normal in a g r e e m e n t with what
has been f o u n d experimentally. In other (intermediate) cases such reasoning does
not lead to satisfactory results which shows that other energy contributions
b e c o m e important in controlling the equilibrium distribution of cations. T h e
problem with all energies included is too involved to be solved without serious
simplifications. O n the basis of systematic studies of cation distributions in various
spinels it has been recognized early, however, that some regularities exist in them
pointing to the possibility to connect the distribution to individual site preferences
of cations. In such a case, the energy 2xP in eq. (8) could be expressed as a
difference

AP = P(M)- P(M') (10)


of individual preference energies P of cations M and M'. O n c e P ( M ) was k n o w n
for all relevant cations, the distribution of ions in arbitrary spinel could be
predicted. T h e attempts were m a d e to determine P ( M ) s f r o m the ligand field
stabilization (with or without covalency effects) (Dunitz and Orgel 1957, M c C l u r e
1957, Blasse 1964) and taking also M a d e l u n g and Born repulsive energies into
account (Miller 1959, G o o d e n o u g h 1963). T h o u g h the latter p r o c e d u r e suffers
from approximations that do not fully c o r r e s p o n d to the real situation, the results
agree at least qualitatively with the experimental data. A n alternative a p p r o a c h
was chosen by N a v r o t s k y and Kleppa (1967) based on empirical distribution data
and t h e r m o d y n a m i c relations. T h e results of both procedures are c o m p a r e d in fig.
13. N o t e that the reliability of these data is better if they are applied to a definite
class of spinels, e.g., the 2-3 type. T h e approximative character of the concept of
OXIDE SPINELS 211

3O /
P(H)
[k~at/~ot] • Phen0 mendo gioaL ~heory [26,21]/• /
+ Determined from / +
20 thermodynamic data [39] / /

• / /
10
/ \/\./jz
0 , + i I [ '

/ / I I :
÷?~+~+--+ I :
Lri4t
I
E
',
,

-10 /
/ :'Mn,
'
4
• GO 4*
r--.~ 2, i

Fig. 13. Preference energies P. The cations are arranged according to increasing empirical octahedral
site preferences (+). The approximate positions of several other ions within this sequence are
indicated, as judged from the experimental distribution data.

individual preference energies may be seen among others from the fact that it
requires /-/1 = 0 in eq. (9) in contradiction with experiment (table 9). Additional
difficulties arise if the cations do not possess a fixed valency and/or in the presence
of some short-range or long-range order.

1.6. O r d e r i n g a n d d i s t o r t i o n s

It follows from the preceding discussion that the spinels are often found with
other than normal distribution of cations. When more than one kind of cations is
present on an equivalent sublattice, a tendency generally exists to decrease the
internal energy by ordering the cations. In most cases such ordering would be only
short range, if any, but when the ratio between the numbers of different cations is
suitable and if the corresponding energy gain is sufficiently large, a long-range
order may develop. As a rule, some symmetry elements are lost and in most (but
not all) cases the appearance of the superstructure destroys the overall cubic
symmetry. The basic types of such ordering in spinels together with some of their
characteristics are summarized in table 10. In all cases the main contribution to
212 S. KRUPI(~KA AND P. NOV~G(

TABLE 10
Types of ordering in spinels.

Symmetry* of
Type Sublattice Characteristics ordered phase Example

1:1 A Every ion M surrounded by


four ions i~ and vice versa F743m Li0.sFe0.5[Cr2]O4

The rows of B-ions in [i10]


1: 1 B and [110] have succession P41 32 Zn[LiNb]O4
a -M'-/('I'-M'- while those in
[101] etc.-M'-M'-M'-M'-

Succession of (001) planes


1: 1 B of B sublattice occupied Immb Zn[LiSb]O4
fl alternatively by M' and (Fe304)
l~I' ions

In the [110] and [110]


i :3 B rows of B sites each P43212 Fe[Li0.sFeLs]O4
fourth ion lVI', others M'

* Symmetry of the ordered phase is taken from C. Haas, 1965, J. Phys. Chem. Solids 26, 1225.

t h e d r i v i n g f o r c e is b e l i e v e d to h a v e a M a d e l u n g c h a r a c t e r a n d t h e s t a b i l i z a t i o n of
t h e s u p e r s t r u c t u r e s h o u l d t h u s i n c r e a s e with i n c r e a s i n g d i f f e r e n c e b e t w e e n t h e
c h a r g e s of t h e i n e q u i v a l e n t c a t i o n s ( t a b l e 11).
A n analysis, w h i c h is s i m i l a r to t h e o n e of e q u i l i b r i u m d i s t r i b u t i o n of c a t i o n s
(section 1.5) c a n b e m a d e c o n c e r n i n g t h e t e m p e r a t u r e d e p e n d e n c e of o r d e r i n g . A t
sufficiently high t e m p e r a t u r e s t h e e n t r o p y t e r m stabilizes t h e d i s o r d e r e d state.
W h e n t h e s a m p l e s a r e q u i c k l y c o o l e d , this state m a y f r e e z e as t h e m i g r a t i o n of

TABLE 11
Energy gained by ordering of cations (in units of eZ/a). The
values which are taken from Blasse (1964), De Boer et al. (1950)
and Gorter (1954) cannot be taken too literally. Being based on
the point charge model they largely overestimate the real
values.

Type of ordering
Difference of
ion;s valencies 1:1; fl; B 1:3; B I:I;A

1 1.0 0.7 0.5


2 4.0 2.8 2.0
3 9.0 6.3 4.5
4 16.0 11.2
5 25.0
OXIDE SPINELS 213

ions (which is necessary for the superstructure to appear) is negligible below


400°C. Consequently the spinels with the same composition may have profoundly
different properties depending on their thermal history.
A similar situation to that described above exists in magnetite Fe3+[Fe2+Fe3+]O4.
H e r e the superstructure arises due to the ordering of the d-electrons of Fe 2+ and
Fe 3÷ ions located at B sublattice. The activation energy is much lower compared
to the case of ordering of different ions, the phase transition is sharp and appears
at ~120 K. This transition attracted much attention during the last years (Evans
1975, Buckwald et al. 1975, Iida et al. 1977) and it turned out that it is much more
complicated than originally proposed by Verwey and Haaijman (1941).
In the oxide-spinels another type of ordering, called the cooperative J a h n -
Teller effect, is relatively frequently encountered. The necessary condition for this
effect to appear is the presence of transition metal ions which have an orbitally
degenerate electronic ground state (for the orbital degeneracy of electronic states
see fig. 9). The interaction between the degenerate states and the lattice vibrations
leads to an effective coupling between electronic states on different cations. When
this coupling is sufficiently strong and the concentration of active cations exceeds
a certain critical value, the electronic states order and simultaneously a structural
phase transition from cubic to lower symmetry appears. The phase with lower
symmetry is stable only below a critical t e m p e r a t u r e Tc. In most cases the system
undergoes a first order transition at To, though a second order transition was also
observed. The dependence of the order p a r a m e t e r c / a on the reduced tern-

,0 o - o"-~
o

ok
L
E .3

2. o5
0.15 2.85 ;4

,5 o.2s

I I /%
o o.2s o',s Ozs

Fig. 14. Temperature dependence of the ordering parameter se in some spinels exhibiting the
cooperative Jahn-Teller effect. ~: is defined as the ratio (c/a - 1)r/(c/a - 1)0. The data are taken from
McMurdie et al. (1950) (Fe0.15Mnz8504), Ohnishi et al. (1959) (CuFe204, CuFel.7Cro.304)and Pollert and
Jirfik (1976) (CrMn204).
214 S. KRUPIOKA AND P. NOVAK

perature for some spinels is shown in fig. 14. A typical dependence of the lattice
parameters c and a on the concentration of active cations is displayed in fig. 15.
In the past much experimental and theoretical effort has been exerted in order
to understand various aspects of the cooperative Jahn-Teller effect in spinels. For
a review see Englman (1972), Gehring and Gehring (1975). In the B site, there are
two ions (namely Mn > and Cu 2+) both having doubly degenerate ground state of
Eg type, which exhibit the Jahn-Teller effect. The corresponding distortion is
always tetragonal with c/a > 1. With the active cations in the A sublattice the
situation is more varied. The cations with the triplet ground state TI(Ni>),
Tz(CU 2+) as well as those with the doublet state E (Fe > ) give rise to the
Jahn-Teller ordering, though in the latter case the effect need not be very much
pronounced. Tetragonal symmetry with both c/a > 1 and c/a < 1 may occur and in
mixed spinels Fe3-xCrxO4 and NiFexCr2 ~O4 an orthorhombic deformation was

I I

9.4 0

9.3
0

2+ 3+ 3+
9.2
o

9.1
9
~. Ucu h

8.9 + Cltetray

o C o

8.8 u VI/3 o
o

8.7 o

8.6 o
o []
o
8.5 z~mo [] D ° r l
A z~
++
+ +

8.4f + -I- .t-


8.3

82 +
+
+

0 0:2 ' ' 0:0 ' 018 ' ;


X

Fig. 15. Room temperature values of the lattice parameters in the spinel system ~VInl+2xCr2_2xO 4
(Holba et al. 1975).
OXIDE SPINELS 215

observed. Data concerning the cooperative Jahn-Teller effect in the binary spinels
are summarized in t a b l e 12, w h i l e t h e v a l u e s of t h e c r i t i c a l c o n c e n t r a t i o n of active
c a t i o n s i n s o m e m i x e d s p i n e l s a r e g i v e n in t a b l e 13. W h e n t h e c o n c e n t r a t i o n o f t h e
a c t i v e c a t i o n s is less t h a n t h e c r i t i c a l , t h e m a c r o s c o p i c s y m m e t r y is c u b i c , b u t still
the local deformations connected with the Jahn-Teller effect persist (Krupi~ka et

TABLE 12
Cooperative Jahn-Teller effect in binary spinels.

Distribution Active
Spinel of ions ion c/a Tc (K) Ref.

Mn304 normal Mn 3+ 03) 1.162 1443 1-5


CdMn204 normal Mn 3+ 03) 1.191 670 2, 6, 7
ZnMn204 normal Mn 3+ 03) 1.142 1298 1, 2, 8
MgMn204 0.78 < 6 < 0.79 Mn 3+ 03) 1.15 1200 3, 9
CoMn204 normal Mn 3+ 03) 1.15 1173 2, 3, 4
FeMn/O4 8 - 0.15 Mn 3+ 03) 1.045-1.064 470 2, 10
CrMn204 inverse Mn 3+ 03) 1.049 569 11
7-Mn203 uncertain Mn3+03) 1.16 12
CuFe204* 0.06 < 6 < 0.24 Cu 2+ 03) 1.06 633 14-16
CuCr204 normal Cu 2+ (A) 0.913 903 13, 14
CuRh204 normal Cu2+(A) 0.91 833 12
NiCr204 normal Ni 2+(A) 1.04 275 17
NiRh204 normal Ni 2+ (A) 1.04 360 18, 19
FeCrzO4 normal Fe 2+(A) 0.986 135 20
FeV204 normal Fe 2+ (A) 1.014 127 21

* Samples rapidly quenched from high temperatures are cubic.

1. Romein, F.C., 1953, Philips Res. Rep. 8, 304, 12. Sinha, K.P. and A.P.B. Sinha, 1957, J. Phys.
321. Chem. 61, 758.
2. Sinha, A.P.B., N.R. Sinjana and A.B. Bis- 13. Bertaut, E.F. and C. Delorme, 1954, Compt.
was, 1957, Acta Crystallogr. 10, 439. Rend. 239, 504.
3. Miyahara, S. and K. Muramori, 1960, J. Phys. 14. Ohnishi, H. and T. Teranishi, 1961, J. Phys.
Soc. Jap. 15, 1906. Soc. Jap. 16, 35.
4. Wickham, D.G. and W.J. Croft, 1958, J. 15. Bertaut, E.F., 1951, J. Phys. Rad. 12,
Phys. Chem. Solids 7, 351. 252.
5. Brabers, V.A.M., 1971,J. Phys. Chem. Solids 16. Prince, E. and R.G. Treuting, 1956, Acta
32, 2181. Crystallogr. 9, 1025.
6. Robbrecht, G.G. and C.M. Henriet-Iseren- 17. Tsushima, T., 1962, J. Phys. Soc. Jap. 17,
tant, 1970, Phys. Status Solidi 41, K43. Suppl. B-l, 189.
7. Day, S.K. and J.C. Anderson, 1965, Philos. 18. Miyahara, S. and S. Horiuti, 1964, Proc. Int.
Mag. 12, 975. Conf. Magnetism, Nottingham, p. 550.
8. Nogues, M. and P. Poix, 1972, Ann. China. 7, 19. Horiuti, S. and S. Miyahara, 1964, J. Phys.
308. Soc. Jap. 19, 423.
9. Irani, K.S. et al., 1962, J. Phys. Chem. Solids 20. Whiple, E. and A. Wold, 1962, J. Inorg. Nucl.
23, 711. Chem. 25, 230.
10. Finch, G.I., A.P.B. Sinha and K.P. Sinha, 21. Rogers, D.B. et al., 1963, J. Phys. Chem.
1957, Proc. Roy. Soc. A242, 28. Solids 24, 347.
11. Holba, P., M. Nevfiva and E. Pollert, 1975,
Mater. Res. Bull. 10, 853.
216 S. KRUPI~KA AND P. NOV~,K

TABLE 13
Critical concentration of the Jahn-Teller ions in some spinel systems.

Active c/a
Spinel ion for x > xt xt Ref.

Mn2+ [Mnz~Cr2-2x]O4
3+ 3+
Mn3+(B) >1 0.40 1
Mny2+ Fel-y[Fel-yFea+y-zxMn2x]04
3+ 2+ 3+ 3+
Mn3+(B) >1 0.48 2
Co2+[Co~+-2xMn~+]04 Mn3+(B) >1 0.6 3
Mg2÷[Al~-+2xMn~+]O4 Mn3+(B) >1 0.59 3
Zn x2+Ge4+x[MnZ+zxMn~+]O4 Mn3+ (B) >1 0.6 3
Co2+[Crz3+zxMn23+]O4 Mn3÷(B) >1 0.57 5
ZnZ+[Cr3-+2xMn3x+]O4 Mn3+(B) >1 0.55 4
Zn2+f[ Fe 23+
- 2 x Mn3+~¢-~
2xP.J4 Mn3+(B) >1 0.65 4
3+ 2+ 2+ ' 2 + 3+
Fel-sCu8 [Cu2xNli-2x 8Fe1+8104 Cu2+(B) >1 0.32 3
3+ 2+ 3+ 2+ 2+
Fe1-~Cu~ [Fel+sM~-2x-sCu2x]O4 Cu2+(B) >1 0.37 3
M = Mg, Co
Cu2x+M2-+x[Cr3÷]O4 Cu2+ (A) <1 0.5 3
M = Mg, Co, Zn, Cd
Ni~+Zn~+~[Cr3+]O4 Ni2+(A) >1 0.6* 6
Fe~+Co~2x[Cr3+]O4 Fez+ (A) <1 0.7¢ 7

* extrapolated to 0 K
?for T - 9 0 K

1. Holba, P., M. Nevf'iva and E. Pollert, 1975, Mater. Res. Bull. 10, 853.
2. Brabers, V.A.M., 1971, J. Phys. Chem. Solids 32, 2181.
3. Goodenough, J.B., 1963, Magnetism and the Chemical Bond (Wiley, New York, London).
4. Shet, S.G., 1969, Thesis, Univ. of Poona.
5. Mansour, B. et al., 1973, Compt. Rend. 277, 867.
6. Kino, Y., B. Liithi and M.E. Mullen, 1972, J. Phys. Soc. Jap. 33, 687.
7. Arnon, R.J. et al., 1964, J. Phys. Chem. Solids 25, 161.

al. 1964, K u b o et al. 1969, K o z h u h a r et al. 1973). M o r e o v e r , the i n t e r a c t i o n


b e t w e e n the J a h n - T e l l e r ions seems to lead to a clustering of these ions (Blasse
1965, K r u p i r k a et al. 1968, B r a b e r s 1971) at least in s o m e spinel systems.

2. Magnetic ordering

Spinels r e p r e s e n t a classical e x a m p l e of a crystal structure allowing a special type


of m a g n e t i c o r d e r called f e r r i m a g n e t i s m . Neglecting m i n o r differences we m a y
consider all o c t a h e d r a l l y c o o r d i n a t e d cations to form o n e sublattice only (B) a n d
similarly the t e t r a h e d r a l sites to c o m p o s e the o t h e r sublattice (A). T h e s e sublat-
tices are crystallographically i n e q u i v a l e n t a n d if they b o t h c o n t a i n p a r a m a g n e t i c
ions in sufficiently high c o n c e n t r a t i o n the f e r r i m a g n e t i s m (i.e. the antiparallel
o r i e n t a t i o n of the two sublattice m a g n e t i z a t i o n s MA, M s ) m a y occur p r o v i d e d the
i n t e r s u b l a t t i c e e x c h a n g e i n t e r a c t i o n s are a n t i f e r r o m a g n e t i c a n d some require-
m e n t s c o n c e r n i n g signs a n d strengths of the i n t r a s u b l a t t i c e i n t e r a c t i o n s are
OXIDE SPINELS 217

fulfilled (see section 2.2). I n fact, spinels were the first materials where the
existence of such magnetic ordering was recognized and for which the first
molecular field theory was elaborated by N6el (1948). Until the discovery of
ferrimagnetism the magnetic properties of the few magnetic spinels known,
mainly those of magnetite, were classified as ferromagnetic. It was difficult,
however, to understand the low magnetic moments and some other peculiarities
of these materials, e.g., the deviations from the Curie-Weiss law (Serres 1932,
Kopp 1919). The departures from a normal ferromagnetic behaviour were
excellently explained by the N6el theory. If the magnetic cations occupy only one
of the sublattices their magnetic moments usually order antiferromagnetically,
though there are also some few cases of pure ferromagnetism.

2.1. Exchange interactions in spinels

As already pointed out by N6el the cations in spinels are mutually separated by
bigger anions (oxygen ions) which practically excludes a direct contact between
the cation orbitals, making any direct exchange at least very weak. Instead of it,
super exchange interactions appear, i.e., indirect exchange via anion p orbitals
that may be strong enough to order the magnetic moments. It is well known that
apart from the electronic structure of cations this type of interactions strongly
depends on the geometry of arrangement of the two interacting cations and the
intervening anion. Both the distance and the angles are relevant. A survey of the
main super exchange interactions in spinels is given in fig. 16. Usually only the
interactions within the first coordination sphere (both cations in contact with the
anion) are supposed to be of importance and the others are often disregarded (see
further). In the N6el theory of ferrimagnetism the interactions are taken as
effective inter- and intra-sublattice interactions A-B, A - A and B-B. The type of
magnetic order depends on their relative strength.
The theory of super exchange interactions (Anderson 1959, 1963) and the
semi-empirical rules of Goodenough (1958) and Kanamori (1959) yield some
predictions concerning the sign and strength of the super exchange interactions.
As they were originally formulated for the 180° and 90° configurations a direct
application is possible only for the B-B interaction with the angle M-O-M' ~ 90°
(table 14a). The case of the A-B interaction with the angle ~125 ° is more
complicated. The usual way is to interpolate between the 180° and 90° case

6.
AB(nn) BB(nn) AB(nnn) BB(nnn) AA(nn)
Fig. 16. Main super exchange interactions in spinel structure.
218 S. K R U P I ( ~ K A A N D P. NOVi~d~[

T A B L E 14a T A B L E 14b
Prediction of the Prediction of the exchange interaction b e t w e e n cations in A and B
exchange interaction for sublattices.
nn cation pairs in the B
sublattice. angle M - X - M '

d 3 - d3 tt or ]'$ * A B 180 ° 90 ° 125 °*


d 5 - ds ]'~,weak
d 8_ d 8 d2 d3 i'J,weak ]'$ or 1't w e a k uncertain
d ~- d 8 ~'~,w e a k d5 d3 ~t weak t~,m e d i u m (weak) I'$ (uncertain)
d 5 - d8 I'1'or 1"$ d5 ds ]'~strong ]'~m e d i u m ~'~
d 3- ds ]'~weak d5 d8 ]'J,strong 1'$m e d i u m (weak) ]'$
d7 d3 T]'weak ]'$m e d i u m (weak) T$(uncertain)
* Direct exchange be- d7 ds ~'~strong ~'~m e d i u m (weak) ]'~,
tween cations orbitals; it d7 d8 ~'~strong ]'~ or 1]'weak ]'~,
is of c o m p a r a b l e mag-
nitude with the s u p e r * Interpolation.
exchange which is ferro-
magnetic here.

supposing that the change is smooth. This is indicated in table 14b for pairs of
cations with the non-degenerate orbital ground state. If signs of the 180 ° and 90 °
interactions are opposite such interpolation does not give a definite result and the
experiment is to be consulted. The A - A interactions are weak and do not
influence the type of magnetic order as far as there is sufficient number of
magnetic B ions.
Quantitatively, the exchange interactions may be characterized by the ap-
propriate exchange constants (integrals); usually the Heisenberg type isotropic
exchange,

~ex = -- 2J~jS~Sj, ( 11)

between nearest neighbours is sufficient even though biquadratic exchange and


sometimes anisotropic terms or exchange between more distant ions are also
considered. The exchange integrals may be deduced from various experiments:
directly from the paramagnetic resonance spectra and/or optical spectra of pairs
of ions in diamagnetic host crystals; from the dispersion relations of magnons
measured by inelastic neutron scattering; by analyzing the magnetic data, parti-
cularly the temperature dependence of the total magnetization and/or sublattice
magnetizations. The most detailed results (obtained by EPR and optical in-
vestigations) have been published for Cr3+-Cr 3+ pairs in octahedral positions (fig.
17, tables 15(a) and (b)).
The results obtained indirectly in magnetically concentrated systems have to be
regarded with some caution because of the simplifications usually made (only
nearest neighbour isotropic exchange, eq. (11), is considered, usually the molecu-
lar field approximation is used, often the collinear spin structure is assumed even
when it is not fully justified etc.). Some of these results are listed in table 16.
OXIDE SPINELS 219

o Oxygen ions

Bj

B2

Bo
Fig. 17. Labelling of the B sites for table 15a.

TABLE 15a
Bilinear exchange integrals for Cr3+ in ZnGa204 spinel
according to Henning (1980). For labelling of the sites
see fig. 17.

Exchange
Pair Distance (,~) integral (cm-1)

Bo-B1 2.947 -11.1 (5)


Bo-B2 5.104 -0.470 (5)
Bo-B3 5.894 -0.610 (5)
Bo-B4 5.894 -0.400 (5)
Bo-B5 6.589 -0.225 (5)
Bo-B6 8.335 -0.275 (5)

TABLE 15b
Bilinear (J) and biquadratic (j) exchange integrals for nn Cr3+ pairs.

Distance
System B0-B1 (A) J (cm-1) j (cm-1) Ref.

Lio.sAlzsO4 2.79 -25 1.3 1


MgA1204 2.86 14.2 (10) 2.1 (7) 2
Lio.sGaz.sO4 2.90 -13.5 (5) -0.6 (13) 3
ZnGa204 2.95 11.1 (3) -1.7 (4) 4

1, Szymczak, H. et al., 1973, Proc. Int. Conf. Magnetism, Moscow


1973, vol. 5, p. 425.
2. Henning, J.C.M. and H. van den Boom, 1977, Physica B + C 86-88,
1027.
3. Gutowski, M., 1978, Phys. Rev. B18, 5984.
4. Henning, J.C.M. et al., 1973, Phys. Rev. B7, 1825.
220 S. K R U P I C K A AND P. N O V A K

I e'
q

b.,

O
,,-.¢
b,, b.,
0a ~ =
ui
~_o ~:~

<~. o
<
<

0) 2

~4,d t--: od

~D t'--

.,= t'--
~. tq. c
t"- t",l t~ ,4 '--~ r" t-,-
¢-,: t ~
',1 ' ~
I I I I
e'. t"- . ~
.=_.
~D + + ~+~+ ~m+l
t"-

ca ~., ~+< ÷ -
¢-i< m
cc
O O ( la ~ (

©
C

% ¢~ ~t-.- e,'; • 0a
;:,..
~+,4' ~5

~5
~e ,q < ~= ..-
.,I
OXIDE SPINELS 221

2.2. Magnetic ordering: theory

If both sublattices contain magnetic ions, the collinear ferrimagnetic arrangement


(fig. 18(a)) occurs provided (i) the A - B interaction is antiferromagnetic and, (ii)
each of the A - A and B-B interactions is either ferromagnetic, or antifer-
romagnetic but weak compared with the A - B interaction. This is a typical
magnetic structure for many spinels, particularly inverted ferrospinels. The ap-
proximate behaviour of a ferrimagnetic spinel follows from the molecular field
type theory given by N6el (1948). The molecular fields hA, hB acting on magnetic
ions in A and B sublattices are sums of two molecular fields corresponding to
inter-sublattice and intra-sublattice exchange, respectively:

hA = n(aMA -- MB),
n> 0 (12)
hB = n(--MA + flMB),

w h e r e MA, MB are the sublattice magnetizations and n, na, n/3 are the molecular
field coefficients for A-B, A - A and B-B interactions respectively; they are
related to the corresponding exchange integrals by,

n = NA 1 ~ --2Jo/(gAgBtz~),
j~A(i~B)
na = N A1 E -- 2 J i j / ( g A l Z B ) 2 , (13)
jEA(i~j, iEA)

nfl = NB 1 E -- 2Sij/(gBlzB)2 ,
j~B(i~j,i~B)

where NA(NB) is the number of A(B) sites in the volume to which the mag-
netization is referred.
For a description of the statistical behaviour of magnetic moments in both
sublattices we have now two equations:

IMAI = MAoBsA(gASAmBIH + hAI/kT),


IMBI= MBoBss(gBSB/~s]H + hsl/k T) ,

T
a b c d
Fig. 18. Basic types of the spin ordering in spinels: (a) Collinear ferrimagnetism (N6el configuration);
(b) Canted ferrimagnetism (Yafet-Kittel configuration; (c) Antiferromagnetism; (d) Spiral structure.
222 s. KRUPICKA AND P. NOV,~K

that have to be simultaneously solved to obtain /]4fA, Ma and the total mag-
netization M = M A + MB as functions of temperature and magnetic field. Bs(x) is
the Brillouin function. The main conclusions drawn from the solution are as
follows:
(i) In the paramagnetic region a hyperbolic dependence of 1/X on T,

1T- O ;~
X CA + Ca T- O" (14)

is obtained instead of the Curie-Weiss law. In eq. (14)

0 = [CACB/(CA+ C B ) ] n ( - 2 + O~CAICB+/3CalCA),

0 '= [CACa/(CA+ Ca)]n(2 + a +/3),

CA, Ca are Curie constants for the A and B sublattices respectively

CACa n2[Ca(1 + a ) - CB(1 +/3)12.


( - (CA + Ca) 3

(ii) The paramagnetic Curie temperature Tc following from eq. (13) for
1/X = 0, depends on all interactions present, viz.

Tc = ½n{CAC~ + CB/3 + [(CAa -- CB/3)2 + 4CACB] 1/2} ; (15)

(iii) For T < Tc, the temperature dependence of MA generally differs from
MB(T) which leads to a variety of possible types of temperature dependence
M(T), depending on a and/3 (see fig. 19). Some of them are very different from
the "normal" behaviour of ferromagnetics (non-monotonic curves, compensation
point, etc.);
(iv) At T = 0 the collinear ferrimagnetic arrangement (called often the Ndel
configuration) with both sublattices magnetically saturated is predicted to be
stable if c~ > --MB(O)/MA(O) and/3 > --MA(O)/MB(O).
In other regions of the (c~/3) plane the simple theory predicts one or both
sublattices to be unsaturated (partly or fully disordered). This conclusion has
proved to be not correct due to the oversimplifications included in the model: (1)
only two sublattices were considered instead of 6 (six cations in the primitive cell),
and (2) the magnetic moments in each sublattice were a priori assumed to be
parallel.
The removal of the restriction (1) by Yafet and Kittel (1952) led to the so-called
triangular arrangement (fig. 18(b)) in which the spins in one of the sublattices (say
B) are canted and divided into two groups B1, B2 with opposite canting angles; the
corresponding sublattice magnetization MB = MB1 + MB2 is then antiparallel to the
spins in the other sublattice A. This configuration should have replaced the Ndel
solution with one sublattice unsaturated. As shown by Kaplan (1960) and Kaplan
OXIDE SPINELS 223

/'4B

Ns

"K "~5

~g

Of

c~p = i
Paramagnetic
/
~ y~///~/////////,,
. / /
Subtattice B
region unsaturated

Fig. 19. Division of (aft) plane into regions with different magnetic behaviour. For the ordered regions
the type of the M ( T ) dependence is indicated.

et al. (1961), h o w e v e r , it d o e s not r e p r e s e n t a state with m i n i m u m e x c h a n g e


e n e r g y in cubic spinels, w h e r e o t h e r m o r e c o m p l e x c o n f i g u r a t i o n s of t h e spiral
t y p e p o s s e s s i n g l o w e r e n e r g y exist (fig. 20). In t h e i r t h e o r e t i c a l study, which is u p
to n o w t h e m o s t c o m p l e t e o n e c o n c e r n i n g t h e g r o u n d s t a t e m a g n e t i c c o n f i g u r a t i o n
in spinels within t h e classical limit, e v e n r e s t r i c t i o n (2) was r e m o v e d . O n t h e o t h e r
h a n d t h e A - A i n t e r a c t i o n s w e r e n e g l e c t e d a n d only n e a r e s t n e i g h b o u r inter-

8/9 2 4 6 8 10 12 14

E
o ~I/E
I
' .......... ,
EL
Ua I
-20 K,TTEC

-60 LOWER BOUND ~ ~ ~

- 80 ~ ~ ~ ' ~

Fig. 20. Dependence of the normalized energy ~ = E / J ~ S A S B N of various spin configurations in spinel
on the parameter u = 4JBBSB/3JABSA; after Kaplan et al. (1961).
224 S. KRUPICKA AND P. NOVfid(

actions were taken into account. They also found an improved criterion for the
stability of the N6el configuration as expressed by the critical value uo = ~ of a
single parameter u = 4JBBSB/3JABSA describing the relative strength of the B-B
and A - B interactions. For u t> u0 the collinear structure is expected to change
continuously into a ferrimagnetic spiral (fig. 18(d)). Nevertheless, the Yafet-Kittel
type configurations may become stable in tetragonally deformed spinels (Mn304,
CuCr204). Note that in this case 5 generally different exchange integrals between
nearest neighbour cations have to be considered (including JAA which is usually
taken ~0).
Even the collinear ferrimagnet may become canted under the influence of a
strong magnetic field whose torque is sufficient to compete with the exchange
torques. The effect was theoretically studied in the molecular field approximation
(Tyablikov 1956, Schl6mann 1960) and it was found that two critical fields
Ha = n ( M B - MA), /-/2 = n(MB + MA) exist so that the overall magnetization M is
constant and equal to ]MB- MA] for H < Ha and equals MA + M~3 for H > H2. In
the intermediate region M linearly increases with H. As the molecular fields are
rather high for most of the ferrimagnetic spinels, this effect is expected to become
observable only in the region of the compensation point, i.e., when MA --~MB.

2.3. Magnetic ordering: experiment

The results included in this section are mostly based on direct neutron diffraction
determinations of the magnetic structure and completed by magnetic measure-
ments.

2.3.1. Spinels with one magnetic sublattice only


If only the A sublattice is occupied by magnetic ions, an antiferromagnetic
arrangement appears in all known cases. As JAA is small, the corresponding N6el
temperature, TN, is l o w (table 17). Two sublattices A1, A2 with mutually
antiparallel spin alignment were found to be identical with the crystallographical
ones (section 1.1) (Roth 1964).
The situation for spinels with magnetic ions in the B sublattice only is more
complex. As pointed out by Anderson (1956) the octahedral sublattice possesses a
peculiar topology which prevents a long range order to be stabilized by nearest
neighbour (n.n.) antiferromagnetic interactions only. Hence the interactions be-
tween more distant ions have to be considered. A number of different arrange-
ments may then appear as discussed by Plumier (1969), Aiyama (1966), Akino and
Motizuki (1971) and Akino (1974). The complexity may be somewhat reduced if
the symmetry of the crystal structure is lowered either by the cooperative
Jahn-Teller effect or due to the magnetic interactions. Note that large single ion
anisotropy of some ions (Co >, Fe E+, Ni 2+) may strongly influence the resulting
magnetic structure; this is believed to be the case, e.g., for normal 4-2 spinels
G e 4+ [M2+]O42-, M = Fe, Co, Ni (Plumier 1969).
In few cases the ferromagnetic ordering occurs. Only typical examples with
magnetic ions in B sublattice together with the relevant data are listed in table 18.
O X I D E SPINELS 225

T A B L E 17
Spinels with magnetic ions in the A sublattice only. The
A - A interactions seem to be enlarged by the presence of
the transition metal ions Co 3+, Rh 3+ in the B sublattice
even though their spin is zero.

Spinel Tr~ (K) ~ga(K) Ref.

Mn2+Al~+O4 * 6 1
Mn2+Rh~+O4 15 -25 2
Fe2+A13+O4* . 8 1
Co2+Al3+O4 * 4 1
Co 2+Rh 32+O4 27 - 30 3
Co2+Coz3+O4 46 -50 2
Ni2+Rh3+O4 18 -20 2, 3
CuZ+Rh~+O4 25 -70 2

* 5 to 15% inverse spinel.

1. W.L. Roth, 1964, J. Phys. Rad. 25, 507.


2. G. Blasse, 1963, Philips Res. Rep. 18, 383.
3. G. Blasse and D.J. Schipper, 1963, Phys. Lett. 5,
300.

T A B L E 18
Spinels with the magnetic ions in the B sublattice only.

Crystallographic
Spinel structure Magnetic structure ~9a (K) TN or Tc (K) Ref.

Mg2+[Cr~+]O4 cubic at 10 K, weakly complex antiferromagnet -350 14 1


tetragonal at 4.2 K
Mg2+ [ V 3 + ] 0 4 cubic at 77 K, antiferromagnet -750 45 1
tetragonal at 4.2 K
Zn 2+[Mn 3+]O4 tetragonal antiferromagnet -450 200* 2
Zn2+[Fe3+]O4 cubic complex antiferromagnet -50 9 3
Gen+[Fe~+]O4 cubic antiferromagnet - 15 10-11 4
Ge 4+[Ni2+]O4 cubic antiferromagnet -6 15 4, 5
C u + [NI•2+ 4+
1/2Mn3/2]O4 cubic ferromagnett -20 150 6
Cu +rM~2+
[ N 1/2Mn 4+ 3/2J10 4 cubic ferromagnet -75 57 6

* Different N6el temperature TN = 50 K was reported in ref. (7).


t Magnetic moments of Ni 2+ antiparallel to the Mn 4+ moments.

1. Plumier, R., Theses, Paris 1968.


2. Aiyama, Y., 1966, J. Phys. Soc. Jap. 21, 1684.
3. K6nig U. et al., 1970, Solid State Commun. 8, 759.
4. Blasse, G. and J.F. Fast, 1963, Philips Res. Rep. 18, 393.
5. Bertaut, E.F. et al., 1964, J. Phys. (France) 25, 516.
6. Blasse, G., 1966, J. Phys. Chem. Solids 27, 383.
7. Gerard, A. and M. Wautelet, 1973, Phys. Status Solidi a16, 395.
226 S. K R U P I C K A A N D P. N O V i ~ K

2.3.2. Spinels with magnetic ions in both subIattices


The A - B interaction was always found to be antiferromagnetic. For many inverse
or almost inverse ferrospinels this interaction is predicted to be strong (table 14b)
and, accordingly, a collinear ferrimagnetism is expected to appear (section 2.2).
This was confirmed experimentally-representative data are given in table 19.
T A B L E 19
Examples of spinels with the collinear ferrimagnetic spin arrangement.

riB(A) nB(B) nB
System theor, neutrons theor, neutrons theor, neutrons experiment

Fe3+ [Fe3+Fe2+]O4 5 5 9 9 4 4 4.03-4.2

Fe3+[Ni2+Fe3+]O4 5 5 7 5 (Fe
2.3 (Ni3+)
2÷) 2 2.3 2.22-2.40

Fe3+ [Li~.sFe~]O4 5 - 7.5 - 2.5 - 2.6

Mn0.98Fez0204* 5 4.99 9.74 9.72 4.74 4.73 4.52-4.84

* Partially inverse spinel with 6 = 0.13 (see Jir~ik, Z. and J. Zajf~ek, 1978, Czech. J. Phys. B28, 1315).
A s s u m e d ionic distribution M n 2+ 3+ [Mn o.
0.85Feo.z5 3+13Fe3+ 2+13] 04.
1.74Feo.

When the A - B interaction becomes comparable with one (or both) of the
intra-sublattice interactions the collinear structure is destabilized and complicated
structures with canted spins are observed. Spiral structures were reported in
several chromites and Yafet-Kittel-like structures in manganites, vanadates and
some chromites (see table 20). In some of these compounds transitions between
various spin arrangements have been observed. Note that the non-collinear
structures may be also very sensitive to substitutions (Vratislav et al. 1977).

T A B L E 20
Examples of spinels with the non-collinear spin structure.

nB nB Low temperature
System riB(A) nB(B) N6el exp. magnetic structure Ref.

Mn304 4.34 3.64 (B1) 2.54 -1.88 Y-K-like structure 1


3.25 (Bz) in the B sublattice

MnCr204 4.3 2.52 (B1) 1.02 -2.28 spiral in both A 2


2.8 (B2) and B sublattices

CoCr204 2.44 5.88 3.44 0.2 spiral in the B 3


sublattice

MnV204 4.24 2.2 -2.04 -2.74 Y - K structure in 3


the 13 sublattice

1. Jensen, G.B. and O.V. Nielsen, 1974, J. Phys. C7, 409.


2. Vratislav, S. et al., 1977, J. Mag. Magn. Mat. 5, 41.
3. Plumier, R., 1968, Thesis, Paris.
OXIDE SPINELS 227

A special group is formed by the inverse 4-2 spinels, examples being the
titanates M2+[Ti4+M2+]O2 or stannates M2+[Sn4+M2+]O2- (for summary of
experimental data see Landolt-B6rnstein 1970). In these compounds the spin
moments are compensated at 0 K as far as the arrangement is strictly collinear.
Nevertheless, finite magnetic moments may appear due to the unequal orbital
contribution of the M 2+ ions in A and B positions. The measurements on
ulv6spinel Fe2TiO4 indicate further a possibility of a weak ferromagnetic moment
to appear due to a small canting of spins (Ishikawa et al. 1971).

2.3.3. The effect of diamagnetic substitutions


In principle, diamagnetic ions may enter either one of the sublattices only, or
both, depending on the relative site preferences of the ions present. The diamag-
netic substitutions in A sublattice probably represent the most clear-cut case, the
Zn substituted ferrites being a well known example. Because of JAA = 0 only the
A - B interactions are effectively weakened. If the number of substituted ions is
not too high, however, (usually up to 30 or 40%) the overall ferrimagnetic
arrangement is not destroyed for T ~ 0 K even though some loosely bound spins
may become locally canted or disordered at higher temperatures T < Tc. For
larger substitutions the A - B interactions may become comparable to, or even
weaker than the B - B interactions. In this way the collinear ferrimagnets often
change to canted ones. Both local canting and long-range non-collinear structures
were reported (e.g. Piekoszewski et al. 1977, Zhilyiakov and Naiden 1977 and the
references therein), the interpretation of experimental results being, however,
often controversial. Finally, above a certain critical concentration of diamagnetic
ions both sublattices are practically decoupled and only the B sublattice becomes
magnetically ordered (usually antiferromagnetically).
While the diamagnetic substitution as a rule lowers the ferrimagnetic Curie
temperature, the saturated magnetic moment at 0 K may increase. This is shown
for Zn substituted ferrites in fig. 21. If the ferrimagnetic order could be retained
even when approaching ZnFe204 the magnetic moment at 0 K should increase
steadily up to the limit 10/ZB per molecule; the breakdown in the vicinity of 50%
substitution may be interpreted in terms of spin canting mentioned above. For a
more detailed description the statistical methods (Gilleo 1958, 1960, Rosencwaig
1970) originally developed for diamagnetically substituted garnets, may be used.
Diamagnetic substitution in the B sublattice weaken both A - B and B-B
interactions simultaneously and hence the collinear ferrimagnetic structure need
not be so strongly affected, only Tc is always decreased. The examples often
quoted are the A1 substituted ferrites (for experimental data see Landolt-B6rn-
stein 1970). As the net magnetization in simple ferrites is parallel to the sublattice
magnetization MB, the B substitutions lower the saturated magnetic moment and
a compensation point (M = 0) may appear. Note that the change in JBB/JAB and
MA/MB ratios generally modify the form of the M versus T curve (see section
2.3.4). The lowering of the sublattice magnetic moments found in some neutron
diffraction experiments (e.g. K6nig et al. 1969) may be due to the local spin
canting.
228 S. KRUPICKA AND P. N O V ~ K

, , / 10
..-i~~ na
..>?.~C';"
M .-'S:"~JS:" 8
o~'g-
""" ~ o
o~ ' ' ; ' "
o~ . ~ - ' ~ o °\ 6

/o..~/
% , /o.:>~o- \o %
co ",o. %
Ni" - -
\ 2

o ob ot~ oi~ o.8 ~.o


MFe204 x ZnFe204
Fig. 21. Dependence of the saturated magnetic moments of ZnxMl-xFe204 ferrites on x. Data are
taken from Gorter (1954) (©) and Sobotta and Voitl~inder (1963) (+).

~"'+'°'~.o.~. X = '0 o' experi'men~al


50 ~ o.. u + at 90° K
O's J %\ o a~ RT
emulg I o~
40 i. %

0--'°-oooQ
"O,o\ ° \
so \
~ + ° - o . o 0.50 % ½
"13.0 0 \ I
,or-----+.. "%. \ ~o4
}I,_ U.O,~/
,~o,~°--o0.60
~
% ~'0
\ \1
+ ~0~0..~ o 0~0 '~" 0.~.0 ~
~o. o ~ "o,.~
-----÷000.00.75 o-~.o.~ ~O_o~O.~
0~--O--o.._o~2oZll
~--+o-o-o-o-o.o_o_o_o:oO~Oo~-/[
+ o-o-o-o-o-o~ ~'~'O~o

-,0 ~__._~_ ,~ o..O>° /


-~-o........... -- / I
.._...~. _ x =/.00 ~_...o/ I
-20 ~ + ' ° ~ 0 ~ 0 ~ 0 " 0 ~ 0 ~°~
0 0.2 0.4 0.6 0.8 1,0
%
Fig. 22. Temperature dependence of the saturated magnetization of NiFe2-xVxO4 sYstem, after Blasse
and Gorter (1962).
OXIDE SPINELS 229

2.3.4. Temperature and field dependences


The constraint of antiparallel orientation of the sublattice magnetizations reduces
the freedom of the individual m o m e n t s in a collinear ferrimagnet which results in
deviations from the normal statistical behaviour as expressed by the Brillouin type
t e m p e r a t u r e dependences of MA and MB. The same is of course true for the
overall magnetization M = MB -- MA for which the different characteristic types of
M = M ( T ) curves were predicted by the N6el theory (see fig. 19). All of these
types were found experimentally. In fig. 22 the curves for the system of V
substituted nickel ferrites are shown; all 3 types of M ( T ) dependence appear
when the V content is gradually increased. A lot of neutron diffraction, M6ss-
bauer and N M R data are available concerning also the t e m p e r a t u r e course of

1.0 1.0
~" -o ~ % o
"-, ~'o
0.8
-,.\ MB/M8 (0)
0.8
\~ xo
,xo
0.6 0.6

0.4 0.4
~0

0.2

i I i
0.2
't
o12 o,~ o'6 o.8 7.0 012 o'.~ o'.6 o18 1.o
r/rc
i

1.0
-° ~'~o~
Msms(# ",oN,°
0.8
o\
0.6 k o~

~\°~ o
0.4 "~"b~'o

0.2 ~o

o'.2 o'.4 o16 o.'8 ,.o


r/re
Fig, 23. Comparison ot experimental and theoretical temperature dependencies of Ms, MA and MB for
the Li ferrite (Prince 1965). Full curves were obtained by taking the biquadratic exchange into account.
Dashed lines correspond to the simple N6el model with the zero intra-sublattice exchange. Good
agreement of the N6el model with the experimental Ms(T) may be also achieved if appropriate
intra-sublattice interactions are assumed. Then, however, calculated Ma(T) and MB(T) for 0.4~
T/Tc ~ 0.9 are considerably smaller than those determined directly by neutrons (Prince 1964).
230 S. KRUPICKA AND P. N O V ~ K

individual sublattice magnetization (K6nig et al. 1969, Sawatzki et al. 1969, Prince
1964, Yasuoka 1962). Note that their interpretation in terms of molecular field
coefficients usually leads to an overestimation of JAA; this difficulty may be
removed, however, when biquadratic exchange is taken into account (see fig. 23).
In spinels with non-collinear magnetic structure the canting angle represents the
necessary additional degree of freedom for the system of magnetic moments to
make their statistical behaviour Brillouin-like (within the limits of molecular field
theories). For this it is irrelevant whether the non-collinearity is due to competing
exchange interactions or induced by a strong external magnetic field. It holds also
for all canted structures that the canting angle depends on the applied magnetic
field, even for T ~ 0 K. As a consequence the net magnetization increases with
increasing magnetic field even when technical saturation has been reached. This
behaviour was experimentally found for many ferrimagnetic spinels possessing
non-collinear magnetic structures, such as manganites, chromites and others,
including systems with diamagnetic substitutions. Examples are given in fig. 24.
Jacobs (1959) analyzed the data assuming a triangular spin configuration. He
found for this special case that the increase of saturation magnetization at low
temperatures may be related to the molecular field coefficient n/3 for the B-B
interaction by the simple equation

AM = HintS. (16)

The behaviour in the paramagnetic region was found to be similar in all


ferrimagnetic spinels, and corresponds to the predictions of the N6el theory (eq.
(]4)).

560 [ . . . . . .

520 L/" ° ~ ° - - ° - - °-- °~°~°~°--°


,~. Mn Fe2 0,; ,7F°K
24O

200 za z~ . . . . - - z a

160 _~/,~.... ~ ~
Zx ~'
~ 4.2OK
(Mnj/c,'Aq
120 -I

80 a

[] u~d u~ 4.2 K

o 2'0 ~'o ~'o 8'0 ~'oo ~o ~4o


(koe)
Fig. 24. Dependence of the magnetization on an applied magnetic field for three spinel systems
(Jacobs 1959, 1960). While there certainly exists a non-collinear spin arrangement in Mn[FeCr]O4 and
Mn[Cr2]O4, the spin structure of MnFe204 is not yet unambiguously determined.
OXIDE SPINELS 231

2.4. S p i n w a v e s

The simplest quantum mechanical approach to the ferrimagnetism in spinels uses


the two sublattice model (Kaplan and Kittel 1953). As a consequence two magnon
b r a n c h e s - o n e acoustical and one o p t i c a l - a r e obtained. An important result of
this model is that for small values of the wave vector k, the energy of the
acoustical magnons is quadratic function of k,

E = if0 + ~ k 2 , (17)

with

_ 2JAAS 2 + 4JBBS~ - I I J A B S A S B a2 '


16ISA - 2SB[

where a is the lattice constant, JAA, JBB and JAB are the exchange integrals.
The dispersion relation (17) yields the well known r 3/2 dependence of mag-
netization and specific heat at low temperatures

M(T)/M(O) ~- 1 - ~'(3/2)0 3/2 , (18)

C~ -~ ( 1 5 / 4 ) k B [ M (O)/ ge~p.B];',(5/2)O 3/2 ,

with

0 = [g~dM(O)]3/2kBr/(4~),

M (O) = Nola.B(gASa -- 2gBSB),

ge~ = ( g A S A - g B S B ) / I S A - SBI ,

kB is the Boltzmann's constant, gA, gB are the g factors of magnetic ions in A and
B sublattices respectively, No is the number of A cations, and ~'(x) is the Riemann
~" function.
More sophisticated spin wave calculations in spinels (Kaplan 1958, Kowalewski
1962, Glasser and Milford 1963) take into account that there are six sublattices of
cations (section 1.1). Accordingly six magnon branches appear. The quadratic
form (17) of the dispersion relation for acoustical magnons still applies and
therefore both magnetization and the specific heat should follow the r 3/2 depen-
dence at low temperatures.
Experimentally most attention was directed towards magnetite Fe304 where
magnon dispersion curves were determined by several authors (Watanabe and
Brockhouse 1962, Torrie 1967, G r o u p e de diffussion des neutrons 1970) (fig. 25)
using neutron scattering. Some experimental data are also available for M n - F e
(Wegener et al. 1974, Scheerlinck et al. 1974), Co (Teh et al. 1974) and Li (Wanic
1972) ferrites. By fitting the spin wave theory to these experiments the exchange
232 S. KRUPICKA AND P. NOV~K

120 ~ , L , , , , ,

E
/~0v]
1o0
90 ~ * /-+~ (5)a.a(6) -

50 o .+
/
40 o/+
30 % /

I
20 0/4-
10 4 /
o 0.2 & o'6 o18
a__Eo[ool]
2IF
Fig. 25. Dispersion curves of spin waves in magnetite (six branches). Data taken from Groupe de
diffusion des neutrons (1971) (+) and Watanabe and Brockhouse (1962) (©).

integrals may be estimated. Both the sign and the magnitude of the inter-sublattice
exchange integral obtained by such procedure agree with the expected ones
(compare tables 14, 16). The same cannot be said about the intra-sublattice
integrals-e.g, the ferromagnetic B-B coupling and rather large values of JAg
were found in Fe304. Such contradictory results may be connected either with
approximations made in the spin wave calculations (consideration of only isotro-
pic bilinear exchange between nearest neighbours, introduction of effective
exchange integrals etc.) or with insufficiency of the G o o d e n o u g h - K a n a m o r i rules.
The T 3/2 dependence of the magnetization and specific heat predicted by the
spin wave theory was observed in several ferrite systems (Kouvel 1956, Heeger
and Houston 1964). An example of the results obtained is shown in fig. 26.
Several authors reported the observation of spin wave resonance in single
crystal (Ivanov et al. 1972, Baszynski and Frait 1976, Sim~ovfi et al. 1976) and
polycrystalline (Gilbart and Suran 1975) spinel ferrite thin films. From the
resonance fields the value of constant @ in eq. (17) may be determined. This
procedure is somewhat obscured by the unclear way in which the spins are pinned
at the surfaces.
OXIDE SP1NELS 233

T(K)
.~ 50 100 158 200 250

(Mc/s) ~ ' ' '

o 800 1ooo 2400 3200 4000


r ½ (K3,~)
Fig. 26. Temperature dependence of the Mn55 NMR frequency ~ in MnFe204 (Heeger and Houston
1964). ~, is propotional to the sublattice magnetization.

3. Magnetic properties

3.1. A n i s o t r o p y a n d magnetostriction

3.1.1. Introductory r e m a r k s
The anisotropy constants are usually defined with respect to the free energy F of
the system. For cubic symmetry,

F = Fo + K l s + K2p + K3s 2 + . • • , (20)

where

2 2 2 2 2 2. ~2 2~2
S = 0/la2-1- 0/10/3"}- 0 / 2 0 : 3 , p = tXl0/2~ 3 ,

a l , 0/2, 0/3 are the direction cosines of magnetization, and K1, K2, K 3 . . . are the
familiar anisotropy constants. In addition, expressions corresponding to the
tetragonal symmetry and sometimes-also to the orthorhombic one may be
relevant to spinels,

F = Fo+ K l a n + K z a ~ + K3(a4+ 0 / 4 ) + . . . (tetragonal),


(21)
F = Fo + K l a ~ + K ~ ( a ~ - 0/~) + . • • (orthorhombic).

There are several complications connected with such definitions of anisotropy


constants. First, as a rule, the anisotropy is determined from measurements at
constant temperature and external stress so that the experimentally determined
K 1 , / £ 2 . . . refer to the Gibbs potential and not to the Helmholtz free energy. The
difference between the two sets of constants is due to the magnetostriction. The
234 S. K R U P I O K A AND P. N O V / i d ~

detailed analysis as well as the formulae connecting Ki and /£~ are given, e.g., by
Carr (1966). For cubic crystals,

/£1 = K1 + h2(Cll - - C12) - - 2 h 2 c 4 4 - 3hoh3(C~l + 2 c 1 2 ) -}- " • " ,


(22)
g 2 = K 2 - 3 h l h 4 ( C l l - c12) - 1 2 h 2 h s c 4 4 + • • • ,

where Cll, c12, c44 are the elastic constants and coefficients hi characterize the
magnetoelastic coupling. To define hi the Gibbs free energy G is to be expanded
in powers of the stress tensor 0-,

G=Go-A~-½~g~, (23)
g being the elastic compliance tensor. The components of the tensor A are further
written as series in powers of m ,

Ai~ = h0 + hl(O~ 2 - - 1) + h3s + h4(o¢ 4 + 2s/3 - 13) + " ",


(24)
A i j = h2ofio~j + hsoqce].

The last two formulae yield the definition of the hi. The magnetostriction e in a
direction specified by the direction cosines /31, /32, /33 is related to the tensor A
through the relation

e = ~'~ Aij/3i/3j. (25)


/,j

Combining eqs. (24) and (25) the connection between the parameters hi, h2 and
the commonly used magnetostriction constants A100, Aln is established,

3-100= 2 h i 3 ; /~111= 2h2/3 . (26)

One point to note in eq. (22) is the t e r m - 3 h o h 3 ( c n + 2 c 1 2 ) which corresponds to


the contribution of the isotropic strain. This correction depends upon the choice
of the unstrained volume, in particular it vanishes if the state of zero volume
stress is defined by setting h0 = 0. Some authors take as unstrained volume the
volume of a hypothetical crystal with magnetic interactions switched off. Isotropic
strain term may then contribute substantially to anisotropy, e.g., for synthetic
magnetite Birss (1964) estimated this contribution to be - 2 8 % . The disadvantages
of such an approach are: (i) constant h0 may be estimated only indirectly; (ii) it is
difficult to determine h 3 with sufficient precision. It seems therefore more con-
venient to refer the free energy to the volume of the real (magnetized) crystal, in
which case the contribution of the isotropic strain vanishes identically.
The contribution of the anisotropic strain to magnetocrystalline anisotropy of
spinels is, with few exceptions, small. The relevant data for several spinels are
given in table 21.
OXIDE SPINELS 235

TABLE 21
Magnetostrictive contribution AK~ to the first anisotropy con-
stant.

Temperature K1 x 10-5 AK1x 10 s


System (K) (erg/cm3) (erg/cm3)

Fe304 300 - 1.1 - 0.24


Li0.sFe2.504 300 -0.9 0.004
(disordered)
NiFe204 300 -0.7 0.053
TiFe204 77 5.8 52

The second complication in defining the anisotropy constants is connected with


the possible dependence of the magnitude of magnetization on its direction with
respect to the crystallographic axes (anisotropic magnetization). Effectively it
leads to the dependence of anisotropy constants on external magnetic f i e l d - for a
corresponding analysis see, e.g., Aubert (1968). Up to now the experimental results
in spinels were analyzed without taking this contribution into account.
Finally, note that in the presence of the relaxation effects (section 3.3) the
anisotropy measured by static methods generally differs from the one determined
by FMR.
In the following two subsections the microscopic origin of the anisotropy and
magnetostriction will be discussed. A survey of the experimental data will be
given in the Appendix.

3.1.2. Microscopic origin: anisotropy


The dominant source of the magnetocrystalline anisotropy in spinels is to be
sought in the interplay of the ligand field, spin-orbit coupling and the exchange
interaction of the magnetic ions. The magnetic dipole-dipole energy may also
contribute in tetragonal or orthorhombic spinels.
In most cases the single ion model (Yosida and Tachiki 1957, Wolf 1957) is
sufficient to describe semiquantitatively the magnitude and temperature depen-
dences of the anisotropy constants K1, K2. In this model the magnetic ions
contribute additively to the macroscopic anisotropy effects; their interactions are
approximated by effective fields (ligand and exchange) and the anisotropy appears
as a result of the dependence o f their individual energy levels on the direction of
magnetization. T o deduce the low lying levels (i.e. those which may be thermally
populated) the properly chosen effective Hamiltonian is usually used. For ions
with orbitally non-degenerate ground states this reduces to the familiar form of
the spin Hamiltonian (S ~<~):

1 4 4
= IXBHexgS+D[S}-½S(S+ 1)] +~a[S~ + S,+ S z4 - ~S(S
1
+ 1)(3S 2 + 3 S - 1)]
+ ~6F[7S~- 6S(S + 1)S~ + 5 S ~ - 6S(S + 1) + 3S2(S + 1)2]. (27)

This Hamiltonian applies to a cubic crystal field (axes x, y, z) with an axial (e.g.
236 S. KRUPI(~KA AND P. NOV/~G(

trigonal or tetragonal) component along axis (. It may be used, for the 6S ions
Fe 3+ and Mn 2+ (S = I) and also for Mn a+ and Fe z+ (S = 2) in octahedral coor-
dination supposing that the splitting by the crystal field of lower symmetry results
in an orbital singlet ground state. For S ~<-~the quartic terms with constants a and
F can be dropped out of the spin Hamiltonian. The possible anisotropy of the g
tensor corresponds to the anisotropic exchange interactions.
For ions with orbitally degenerate ground state the effective Hamiltonian
contains the terms depending on the orbital momentum and in general its form is
more complicated compared with eq. (27). The ions, which in cubic symmetry
have the orbital triplet lowest, were treated, e.g., by Slonczewsk! (1958), Baltzer
(1962) and Novfik (1972). As an example we quote the effective Hamiltonian
corresponding to Co 2+ ion situated at the B site of the spinel structure (only the
most important terms are retained):

E{ = oqA6Sz + a±A(lxSx + lySy) + DI~ + g~BtIexS. (28)


Here ! = - a L is the pseudoangular momentum (1 = 1), z is the trigonal axis, the
term Dl 2 reflects the presence of the trigonal crystal field. For Co 2÷ ion D is
negative; this corresponds to a situation where the trigonal field splits the ground
orbital triplet into lower doublet and higher singlet. The orbital moment is not
quenched but it is confined to lie along the local trigonal axis. Due to the
spin-orbit interaction the spin is then also coupled to the trigonal axis (see fig. 27),
which is the source of the large single ion anisotropy.
To obtain the single ion contribution ki(T) (i = 1, 2 . . . . ) to the cubic anisotropy
constants Ki the free energy per ion is to be evaluated using the partition function
relevant to the energy spectra of the effective Hamiltonian (Wolf 1957, Yosida
and Tachiki 1957, Slonczewski 1958). For microscopic theory to be compatible

~rig

% fl

g •¢.__.~t
xx/
Exci~ed
sfa~e

~xGround

0 :El2 0 YC
a b
Fig. 27. Origin of the anisotropy due to the B site Co2÷ ion. (a) Coupling of the spin of Co 2+ to the
local trigonal axis. Co) Dependence of the lowest two levels of Co2+ on the angle between mag-
netization and the trigonal axis (assuming /zBH~x>>spin--orbit coupling). The effect of higher order
terms is indicated by weak lines.
OXIDE SPINELS 237

with the macroscopic description put forward in section 3.1.1 (i.e. with /(i
corresponding to the Gibbs free energy) the terms depending on strains must be
added to the Hamiltonian (see section 3.1.3, eq. (30)). Note that averaging over
inequivalent sites must be performed when calculating ki(T). In the cases where
the Hamiltonian (27) applies the contribution to the cubic first anisotropy constant
is (with the F term neglected):

kI(T) = at(y) + 7[DE/(kT)]t(y), (29)

where y = exp[-glzBHex/(kT)], r(y) and t(y) are functions depending on S (Wolf


1957), and 3/ is a constant which equals 4 for ~"--- [111] (four non-equivalent axes)
and - ~ for ~--- [100] (three non-equivalent axes). The latter may be, e.g., the case
in the presence of the local Jahn-Teller distortions.
For T ~ 0

k~(O) = -½S(S - 1)(S - 1)(S - ~)a + 23,S(S - ½)(S - ~)D2/(glXBHex),


so that for ions with $ 4 3 the a term does not contribute to anisotropy. The
temperature dependence (29) may be fitted to the experimental curve KI(T) to
evaluate the constants a, D, gHex (see figs. 28(a) and (b)). On the other hand these
values may be compared with those deduced from an independent experiment
(most often E P R in doped diamagnetic crystals) or estimated on the basis of the
ligand field theory.
In many ferrimagnetic spinels, in particular ferrites, the first term in eq. (29) was
found to predominate (Yosida and Tachiki 1957, Wolf 1957). The second term
becomes important if either S ~<3 or the ratio D2/gtZBHex is of the same order of
magnitude as a. The representative values of the parameters D, a and F are given
in table 22. The one ion character of corresponding contributions to the aniso-
tropy is illustrated in figs. 29(a) and (b).
The anisotropy contribution of ions with the orbitally degenerate ground state
is considerably greater as a rule. The most often encountered example is that of
Co 2+ in octahedral positions mentioned above. It is well known that even very
small concentration of Co in magnetite or in other ferrites substantially influences
their anisotropic behaviour (figs. 30(a) and (b)). The analogous model is believed
to apply also to V 3+ ion in the B sublattice. For Fe 2+ ion the trigonal field in the B
site usually splits the ground triplet into lower singlet and higher doublet, so that
the conventional spin Hamiltonian (27) may be used.
The ions Mn 3+ and Cu E+ which possess doubly degenerate ground state Eg in
octahedral symmetry exhibit a strong Jahn-Teller effect (section 1.6). This,
combined with presence of the local strains, leads to lowering of local symmetry
and to splitting of the ground doublet so that eq. (27) applies as well.
From ions in tetrahedral sublattices which have an orbitally degenerate ground
state the case of Ni 2÷ ion was discussed in detail (Novfik 1972, Pointon and
Wetton 1973). The single ion anisotropy is strong here though the spin-orbit
coupling contributes to it only through the second order terms of perturbation
theory.
238 S. KRUPIOKA AND P. NOV~d4

I i i I i I I I I

o.oO/°~°~
-s~

-I0~

- 150 :< ::;;i;:'o,:


[ I I I I I I I
0 100 200 300 400 500
T(K)

T(K)
oo too 20o 3oo___

-to 3~/ x~(Fe3+)


I//

~ -30t~ ° o o o Experiment
x = 1.55

-5o~ b
Fig. 28. Two examples of the determination of the spin Hamiltonian parameters by fitting the
theoretical Ka(T) dependence to the experimental data. (a) Li0.55Fe2.4504 (Follen 1960). S is the order
parameter, i.e., S = 1 (0.1) corresponds to ordered (almost disordered) Li ferrite. Resulting spin
Hamiltonian parameters are a(Fe~+) = 0.024 cm 1, a(Fe~+)= -0.012 em 1. (b) MnL55Fel.4504 (Krupi6ka
and Novfik 1964). Parameters a(Fe3+) = 0.012cm -I, tzBHex(Mn~+)=28cm a, D(Mn3+)= _1.77cm-1.
The D value is to be taken as the effective one due to the non-collinear spin structure; the corrected value
is D = -2.8 cm 1.

3.1.3. Microscopic origin: magnetostriction


In p r i n c i p l e t h e m a g n e t o s t r i c t i o n m a y b e u n d e r s t o o d u s i n g H a m i l t o n i a n s a n a l o -
g o u s t o eqs. (27) a n d (28), t o w h i c h t h e e l a s t i c e n e r g y of t h e c r y s t a l is a d d e d . T h e
p a r a m e t e r s of t h e e f f e c t i v e H a m i l t o n i a n s a r e t a k e n to b e f u n c t i o n s o f strain, b u t
u s u a l l y o n l y c o n s t a n t t e r m s a n d t e r m s l i n e a r in strain a r e r e t a i n e d . F o r i o n s w i t h
O X I D E SPINELS 239

0.1 0.15 0.2 6"


-5

• 4.2K
• 77K

-10

K,

-15 i i i

"-E"
o

E
0.1

0.05
, °/°/°/
0 0.3 0'.4-
\o ~ _ . 1 0.2 ,~

-0.05
Mn3+
o

b \
-0.
Fig. 29. Additive character of the single ion contributions to anisotropy. (a) Dependence of KI on the
inversion degree 6 in MgFe204 (Arai 1973). (b) MgxMn0.6Fez.4-xO4 (Gerber and Elbinger 1970).
Anisotropy contributions AK~(FeB 2+
) and AKI(MnB 3+
) are shown as functions of concentration A~ in the
molecule of Fe2+ and Mn3+ at temperature T ~ 0.

an orbital singlet g r o u n d state the r e l e v a n t effective H a m i l t o n i a n m a y b e w r i t t e n


in t h e form (Callen 1968):

Y~ = Y((e = O) + ~,~'
~_, [ S ~ O (gH¢x)eix,., + S aeix~,
011) Se, ix,]

+ fourth o r d e r terms + 1 ~ cm,e 2gix . (30)


tx

H e r e YC(e = 0) is the spin H a m i l t o n i a n (27). T h e q u a n t i t i e s OD/Oe,,, form a fourth


r a n k tensor, which in cubic (trigonal) local s y m m e t r y has 9 (18) i n d e p e n d e n t
components.
240 S. KRUPI(~KA AND P. NOV,~K

25
-ff-~ -

co .,j ° M~ ~NiZn
15 \
_ ~~MgFe, FeNiZn
10

0 50 100 150 200 250 300


T(K)

~0

(~m-,)
5O
,,,, ~izn
40,
',, .... \
30 Ni ~. X', \ \
20
10

s~ 16o I~o eoo 2so so0


T (K)

Fig. 30. Co 2÷ contribution to Kl(a) and K2(b) for various ferrites of composition Mel-xCoxFe204 (Broese
van Groenou et al. 1968). The symbols indicate the host ferrite as follows: Fe: Fe304, Mn: MnFe204,
MnTi: MnTio.1sCoo.15Fei.704, MgMn: (Mgo.TsMno.25)Mno.29Fel.7104,MgFe: (Mgo.9~Feo.09)Coo.olFel.9904,Ni:
NiFe204, NiZn: Nio.67Zno.33Fe204, NiZnFe: (Nio.46Zno.29Feo.25)Fe204, Co: CoFe204.
OXIDE SPINELS 241

g-.

0 =8~dZ~o_' o '
2x
,,0.001 0.003 3 -0.005

g--, -40

~- -80
7

•~ -120

- 160

C
Fig. 30c. Dependence of AK1 and AK2 on cobalt content in ordered samples of Lio.5-xl2Fezs-xlzCOx04
(Seleznev et al. 1970) (1) 4.2 K; (2) 77 K; (3) 300 K.

There is an important distinction when calculating magnetostriction in com-


parison with the determination of a n i s o t r o p y - namely in cubic systems the terms
bilinear in spin ( - - O D / O e ~ K ) contribute to magnetostriction constants already in
the first order of the perturbation theory. The same holds for the dipole-dipole
interactions, which consequently must also be considered. On the other hand the
bilinear terms are as a rule much larger than the fourth order terms, and the latter
may therefore be neglected.
Given the form of the spin Hamiltonian (30) the calculation of the mag-
netostriction proceeds similarly as the one of anisotropy. Broadly speaking two
types of contributions may be distinguished: (i) the single ion terms, (ii) two ion
terms. To the second category belongs the contribution of the dipole-dipole
interaction (table 23); this may be determined without difficulty once the lattice
geometry, magnitude of the magnetic moments and the elastic constants are
known.
Much more complicated is the reliable calculation of the single ion contribution.
Here the main obstacle is the determination of the derivatives OD/Oe~,,. In
principle relevant information may be obtained from E P R under pressure;
however, no such experiments have been yet performed in spinels. Moreover, an
assumption must be made that the local elastic properties are the same as the
macroscopic ones. An empirical way of determining the single ion contributions to
magnetostriction in spinels was worked out by Arai and Tsuya (Arai and Tsuya
1973-1975, Arai 1973). These authors used the measurements of magnetostriction
242 S. K R U P I C K A A N D P. N O V f i d (

T A B L E 22
Values of the spin Hamiltonian parameters.

( a - F ) x 102 a × 102
System Ion (cm -I) (cm -1) D (cm -~) Ref. Method

ZnGa204 Fe~÷ 3.53 4.44 -0.2442 1 EPR


ZnAl204 Fe 3+ 4.71 5.75 -0.3402 2
MgA1204 Fe 3+ 4.58 4.77 -0.2467 3
Li0.sAlz504 Fe~+ _+1.66 1.0" 0.104 4
ZnAI204 M n ~+ - 0.075 5
ZnGa204 Cr 3+ - - 0.524 6
ZnAI204 Cr 3+ - - 0.891 7

MgxMn0.6Fez4-x O4 Fe~+ 2.0 8


Mn 3+ 3 Anisotropy
Fe~+ (3.5,-4.0)

MgFe204 Fe 3+ 2.9 9
Fe~+ - 1.8

Li0.sFe2.504 Fe~+ 2.4 10


Fe 3+ - 1.2

* Preliminary result.

1. Krebs, J.J. et al., 1979, Phys. Rev. B20, 6. Kahan, H.M. and R.M. Macfarlane, 1971,
2586. J. Chem. Phys. 54, 5197.
2. Gerber, P. and F. Waldner, 1971, Helv. 7. Drumheller, T.E. and K. Locher, 1964,
Phys. Acta 44, 401. Helv. Phys. Acta 37, 626.
3. Brun, E. et al., 1961, C.R. Colloque 8. Gerber, R. and G. Elbinger, 1970, J. Phys.
A M P E R E 10, 167. C3, 1363.
4. Folen, V.J., in Paramagnetic Resonance, 9. Arai, K.I., 1973, Rep. Res. Inst. Electr.
vol. 1, ed. W. Low (Academic Press, New C o m m u n . T o h o k u Univ. 25, 79.
York, 1963), p. 68. 10. Folen, V.J., 1960, J. Appl. Phys. 31, 166S.
5. Soulie, E. et al., 1973, Solid State C o m m u n .
12, 345.

T A B L E 23
3 3
Magnetoelastic coupling constants L = -gA~00(cll - ca2) and M = -gAulC44; the
calculated dipole~zlipole contribution is to be subtracted from the experi-
mental value to obtain a contribution due to other mechanisms (single ion).

L (cm-I/molecule) M (cm-1/molecule)

Calculated Calculated
dipole~tipole Measured dipole-dipole Measured
System coupling coupling coupling coupling

MnFe204 - 14 15 6.5 -4
Fe304 - 12 12 8 -69
CoFe204 - l1 300 7.5 - 150
NiFe204 - 10 27 7 19
OXIDE SPINELS 243

on Mg ferrites with different degree of inversion to determine the single ion


contributions of Fe 3~ ion in both A and B sublattices (the dipole-dipole inter-
action was taken into account by subtracting its contribution f r o m the experimen-
tal results). Supposing that the Fe 3+ contribution is the same in various spinels the
single ion contributions of Cu 2+ (Arai and Tsuya 1974) and Ni 2÷ (Arai and Tsuya
1975) ions were d e t e r m i n e d from m e a s u r e m e n t s on CuFe204 and Nil-xZnxFe204
spinels respectively. These results are s u m m a r i z e d in table 24.
Ions with an orbitally d e g e n e r a t e g r o u n d state have a m o r e direct coupling
between the lattice and magnetic m o m e n t s and the corresponding contribution to
magnetostriction is large. Similarly, as with anisotropy, most attention was
d e v o t e d to the octahedral C o 2÷ ion (Slonczewski 1960, G r e e n o u g h and Lee 1970).
T o calculate the contribution of Co 2÷ ion to magnetostriction the elastic energy
and a term

~_~ V ~ , e ~ , ,
ix,t~,

which describes the change of the ligand field induced by strain is a d d e d to


Hamiltonian (28). T h e r e are only a few i n d e p e n d e n t matrix elements of V,~,;
using s y m m e t r y a r g u m e n t s and making some approximations in the calculation of
the e n e r g y levels their n u m b e r is r e d u c e d to two. Leaving these two p a r a m e t e r s
free and changing the o t h e r p a r a m e t e r s of the effective H a m i l t o n i a n (28) within
reasonable limits a g o o d fit of experimentally observed t e m p e r a t u r e d e p e n d e n c e s
of A100(r), }till(T) m a y be o b t a i n e d ( G r e e n o u g h and Lee 1970) (fig. 31). T h e Ni 2÷
in tetrahedral sublattice gives also a large contribution to magnetostriction and

TABLE 24
Single ion contributions to the magnetostrictive constants
(sublattice fully occupied by a given ion).

Temperature
Ion Site (K) hi00× 10 6 AlllX 106

4.2 71.1 --15.1


A 77 50 - 13.2
300 8.6 12.6
Fe3+
4.2 -79 24.9
B 77 -61 2.3
300 -21.4 -2.5

4.2 -36 -19.6


Ni2+ B 77 -36 - 19.6
290 -33.5 - 18

A 4.2 -1690 -370


Cu2+
B 4.2 350 100
244 S. K R U P I C K A A N D P. N O V t ~ (

T(K)
0 100 200 300 400
, i ,

rO-e

, - 5ol ' 25

_,oo 20
E

N - 150 15
&
%
-200 I0

-250

x 10-e
a
-30C i i i

ibo 2bo 3bo


T(K)

Fig. 31. Contributions Aha, Ah2 of Co 2+ ions to the magnetostriction constants hi, h2 in
CoxMnl-xFe204 ferrites (Greenough and Lee 1970) for x = 0.038 (A) and x = 0.078 (B). Full curves
were calculated using the single ion model Ahl is given for x = 0.038 only.

may be treated similarly as the octahedral C o 2+ ion (Lioliossis and Pointon 1977).
Anomalously large magnetostriction was observed in the spinels Fe3-xTixO4
when x is close to 1 (ulv6spinels) (Isbikawa and Syono 1971a, b, Ishikawa et al.
1971, Klerk et al. 1977). For almost stoichiometric (x = 0.95) ulv6spinel A100=
4.8 × 10-3 at 77 K, which is almost an order of magnitude larger than A100 of Co
ferrite. The explanation is based on the presence of Fe z+ ions in the tetrahedral
sites. These ions would produce a cooperative Jahn-Teller distortion at tem-
perature T~ lying slightly below the Curie temperature. Long-range magnetic order
seems to suppress the cooperative Jahn-Teller effects; nevertheless a soften-
ing of the crystal as T~ is approached is observed (fig. 32). Due to the small value
of c n - c~2 the magnetostriction of Fe z+ in octahedral sites is then enhanced. The
dependences of A100 and Alll for the system Fe3_xTixO4 on the content of Ti are
shown in fig. 33. The distortion of the crystal is due to the exchange striction.
In the discussion of the origin of magnetostriction we have disregarded the term
in Hamiltonian (30) corresponding to the dependence of the exchange on strain
(exchange striction). Though it generally gives a non-zero contribution to the
magnetostriction constants, it is believed that in spinels with the collinear spin
structure it plays only a minor role. In systems with the non-collinear arrange-
ment, however, its presence is of importance as it often leads to a change of the
crystal symmetry below the ordering temperature. The possibility of a symmetry
lowering always exists when the system of spins has a symmetry lower than the
crystal lattice. Experimentally the tetragonal deformation of the cubic lattice
OXIDE SPINELS 245

6 i i i i i

o
-o8~ o
-o'--~o~...oOo. D
o"~-.o.~ o C,,+,+
5
@ o.~.

x 4 °~o° 80
E
o

• BaLsam ~AttenuatJon(c11_ )
3 o Norlaq •
012
Cli -- Ct2 •
60

.£ \o .g,o-~-'-° •
.2 o .%-'~~.~ 40 o~

LU
iiI I°

. ,,," 20

i I i i 0
200 300
Temperature (K)

Fig. 32. Softening of Fe2TiO4 as T~ is approached (Ishikawa and Syono 1971a).

o~ x
A '<10e /
x AIO0
• A.;, /
1000 I
05

~o.+ * /
. i..~.......--~,/_...... 0.8

0 05 x
Fig. 33. Al00and Am vs Ti content x for Fe3-xTixO4 system (Klerk et al. 1977) for the values of T/Tc
indicated in the figure.

below the ordering temperature was observed, e.g., in MnV204 (c/a = 0.99),
MgCr204 (c/a = 0.998) and MgV204 (c/a = 0.994) (Plumier 1969).

3.2. Magnetic anneal and related phenomena

3.2.1. Origin
Magnetic annealing effects appear as the result of some directional ordering of
defects or local anisotropic configurations in the crystal. In spinels, these usually
246 S. KRUPICKA AND P. NOVAdK

are cation vacancies, substituted ions (impurities), anisotropic short-range order


configurations (pairs or groups of ions) or local Jahn-Teller distortions. A typical
example (even though not the simplest one) is an inverted spinel having cations of
two kinds in crystallographically equivalent octahedral positions. In order to
achieve some kind of directional order of the imperfections with respect to the
direction of magnetization two conditions must be fulfilled, however: (i) an
anisotropic coupling exists between the magnetization vector and the local
configurations (defects), and (ii) in some range of T < Tc these imperfections are
allowed to move or to transform into other ones with a lower energy of their
coupling to magnetization. In such a way the magnetic anneal reflects the ability
of the crystal to adapt itself by an inner rearrangement in the lattice to the given
magnetic state and to stabilize this state by lowering its free energy. If the
magnetization is homogeneous a macroscopic anisotropy results (induced aniso-
tropy). For cubic crystals it is described by (N6el 1954, Penoyer and Bickford
1957):

;K, = - F o +-.., (31)


i i,~j

where direction cosines/3i and c~i refer to the orientation of magnetization during
the annealing process and anisotropy measurement respectively. For polycrystals
we obtain by averaging

FKI = K , cos 2 0, (32)

with

0 being measured from the direction of the annealing field.


The anisotropic coupling responsible for local directional order is of the same
origin as that leading to the magnetocrystalline anisotropy: dipole-dipole inter-
actions, single ion anisotropies and anisotropic exchange (and perhaps further
interactions) in pairs or clusters.
Special types of the induced anisotropy may occur in systems with a non-
collinear spin arrangement. In such systems there exist as a rule several distinct
configurations of spins possessing the same exchange energy. For a given direction
of magnetization the configuration with the lowest anisotropy energy is stable.
When the magnetization is rotated, the energy of another configuration may
become lower. Nevertheless, if the relaxation time of the spin system is sufficiently
long the initial configuration is temporarily retained and an induced anisotropy is
observed. Besides the usual terms (31), it may also possess a unidirectional part.
Krupi6ka et al. (1977), (1980) reported such anisotropy in manganese chromites
MnxCr3 xO4 (x I> 1). It may be noted that the relaxation of this anisotropy is
realized by the rearrangement of the spin system itself.
OXIDE SPINELS 247

3.2.2. Examples of local anisotropic configurations

Cation vacancies

Many spinels, particularly those nominally containing Fe 2+, often exhibit devia-
tions from the oxygen stoichiometry (oxygen excess) which leads to the presence
of cation vacancies. If their concentration on B sites with the [111] local trigonal
axis differs from those in other octahedral positions the dipole-dipole interactions
contribute an axial anisotropic term of the form (Yanase 1962):

E dip = AdipNu (/x - ~..L' ) ( ~ 1 0 £ 2 -~- ~20L3 -~ ~'30£1) , (33)

with/x', Ix denoting the average magnetic moments in the [111] sites and the other
sites, respectively. N is the number of B sites per unit volume and Adip--~
4 x 10 -18 erg/cm 3 as estimated on the basis of the geometrical parameters. This
leads to the induced anisotropy of a G type (i.e. F = 0) with

G ( T., T) = ND(16A2dip/k T)(gS)4[M (T)/M(O)]2[M (Ta)/M (O)I2 , (34)

N[] is the number of vacancies per unit volume, Ta and T are the temperatures of
annealing and measurement, respectively, and gS the average magnetic moment
in /xu. In principle, a similar anisotropy might be introduced by an unequal
distribution of non-magnetic substitutional impurities among four types of the B
positions. Note that such an anisotropic distribution of vacancies and/or non-
magnetic ions also causes the distribution of magnetic ions (i.e. Fe 3+, Fe 2+ etc.) to
be non-uniform which results in an active contribution of these ions to the
induced anisotropy via the mechanism of single ion anisotropy or anisotropic
exchange (Krupi~ka and Zfiv6ta 1968).

Co 2+ in B positions

In the range of low concentrations the Co 2+ ions may be viewed as isolated. In the
presence of exchange their low lying levels corresponding to the spin Hamiltonian
(28) are anisotropic and hence their equilibrium distribution among the four types
of B positions will depend on the direction of magnetization. The resulting
induced single ion anisotropy is again of a G type with (Slonczewski 1958):

a = nN[(X/3] o~A[/2) tanh(V'3laA ]/2kT) - (ah)2/4/XBH~x],


(35)
n = (Nm + Nm - Nm- Niil)/N,

where T is the temperature of measurement, N the total number of Co 2+ in the B


sites, N1~1 their equilibrium number in positions with local axis in [111] etc., A the
parameter of spin-orbit coupling and a a numerical factor, 1 < a <3. In the
special case of Co substituted magnetite laA] = 132cm -1 and IXBH~x--320cm -1
248 S. KRUPICKA AND P. NOV_rid(

were estimated (Slonczewski 1958). Equation (35) applies also to other ions for
which eq. (28) may be used. Single ion contributions to the induced anisotropy
arising from the preferential distribution in the B sites are also expected from ions
with a singlet orbital ground state, via the D term in the spin Hamiltonian (27).
These anisotropies are usually weak but in special cases (Fe 2+) they may achieve a
considerable strength (Watanabe et al. 1978).
If the concentration of Co 2+ (or other active ions) is increased these ions cannot
be considered as isolated and contributions of more complicated local configura-
tions including pairs and eventually larger groups of active ions become important
(Iida and Inoue 1962, Iida and Miwa 1966). As an example a cluster containing an
isolated Co2+-Co 2+ pair is shown in fig. 34; note that there are 6 inequivalent
orientations of this pair corresponding to 6 face diagonals (110). If the occurrence
of differently oriented pairs is unequal, an induced anisotropy appears depending
quadratically on the Co 2+ concentration. Experimentally this was found to be
almost a pure F - t y p e anisotropy ( G = 0) (fig. 35). The source of the anisotropy

Fig. 34. The simplest configuration necessary for the description of the Co2+-Co2+ pair contribution to
induced anisotropy (Iida and Inoue 1962, Iida and Miwa 1966).

15 x 105 i i i

10 o o

~E Oe) x
F
(.9

u~ 5

o o~
~°~. ,.~'" , ,

0.05 0.10 0.15 '.20


x
Fig. 35. Induced anisotropy constants F and O for Co-substituted magnetite CoxFe3-xO4,measured at
room temperature after annealing in a magnetic field at 375 K (Penoyer and Bickford 1957).
OXIDE SPINELS 249

may be sought in both the anisotropic exchange and a local modification of the
ligand field changing the single ion contributions of each ion within the pair
(Tachiki 1960).

Mn3+-distorted octahedra

If a small part of octahedral positions is occupied by Mn 3+ ions the overall


symmetry of the crystal remains cubic, but local Jahn-Teller distortions may
appear (section 1.6). Assuming these to be static (or quasistatic) the local ligand
field of Mn 3+ lowers its symmetry which usually becomes tetragonal. The D term
in eq. (27) then yields uniaxial anisotropy of F~type (G = 0) provided a pref-
erential distribution of the distortion axes may be created by magnetic anneal
(Nov~k 1966). Let us note that due to local stresses and other defects often only a
small part of the distortions is expected to be free to move and to take part in the
process (No%k 1966, Krupi~ka et al. 1980).

3.2.3. Kinetics, magnetic after-effects


In order to achieve thermodynamic equilibrium in the magnetic annealing process
(by establishing a local directional order) some rearrangements in the lattice are
necessary. Hence any change of the imposed conditions (change in the direction
of magnetization and/or its spatial distribution in the magnetic specimen) will be
followed by a relaxation process leading to accommodation of the magnetic
system to the new situation. This manifests itself in relaxation of the induced
anisotropy and in various so-called after-effects summarized, e.g., by Krupi6ka
and Z~v6ta (1968).
The microscopic mechanism underlying the relaxation processes mentioned
above usuallly possesses the character of a diffusion process (migration of ions
and/or vacancies, electron hopping) and it is thermally activated. The relevant
relaxation times then depend exponentially on temperature according to the
Arrhenius relation
~- = a e x p ( - O / k T ) ; (36)
the constant A depends on the type of the diffusion process and O is the
activation energy. Except for a few simple cases when isolated ions or vacancies
rearrange, two or more relaxation times usually exist. The distribution of relax-
ation times is often continuous due to some disorder in the lattice. In practice, the
so-called logarithmic distribution of r (i.e. log r uniformly distributed between
limits log rl, log ~'2) is considered as a good approximation. Large dispersion of
relaxation times has been found in processes involving electron hopping and/or
reorientation of Jahn-Teller distortions. The diffusion may be usually described as
a selfdiffusion process, i.e., without taking into account the driving forces due to
magnetization; this is illustrated in fig. 36. Then

[Ei - Ejl = Io,j - oj l o0 o,


from which eq. (36) follows provided [Ei -Ei[ ~ kT.
250 S. KRUPICKA AND P. NOVd~¢

-- ~l ~ -U-]Oj-
:/
I I

Ei

_ _ i i

Fig. 36. Dependence of the particle energy on the position r relevant to the relaxation process.

If the diffusion in question concerns the ions, the activation energies Q are
typically between 0.7 and 2.5 eV depending on the type of ions and type of the
process. Generally, the presence of cation vacancies makes the diffusion process
faster, the factor Q is lowered and the preexponential factor A in eq. (36)
becomes inversely proportional to the vacancy concentration within certain limits.
A detailed analysis of the vacancy assisted diffusion of Co 2+ ions in Co substituted
magnetite was made by Iida and lnoue (1962) and by Iida and Miwa (1966). In
particular, one relaxation time should exist for isolated Co 2+ ions (the G term)
and two different times for reorientation of Co2+-Co 2+ pairs (the F term). The
effect of the size of the migrating ion on the diffusion rate and the activation
energy was calculated by G e r b e r (1968) on the basis of a simple model.
When the local anisotropic configurations refer to anisotropic distribution of
electronic charge (e.g. the distribution of Fe z+ and Fe 3+ valency states) a migration
of electrons is sufficient for the rearrangement process. The activation energy is
then 4 0 . 7 eV and it is often very low (<0.1 eV). The mechanism of the electron
motion is similar to electrical conductivity (hopping). The spinels with coexisting
Fe 2+ and Fe 3+, or Co 2+ and Co 3+ ions are examples of materials possessing this
type of relaxation. A special case is represented by the local Jahn-Teller dis-
tortions. They are connected with the ordering of partly filled orbitals and their
reorientation does not need any diffusion even though some thermal activation
is necessary in order to overcome an energy barrier. A p a r t from this process the
distortions are usually stabilized by some defects and the mechanical stresses
connected with them; they reorient only if these defects move. This seems to be
the case of low t e m p e r a t u r e magnetic anneal in Mn-rich Mn ferrites where the
distortions of Mn 3+ octahedra may be stabilized by some extra electrons placed at
the neighbouring ions (Krupi6ka and Zfiv6ta 1968).
Apart from thermal activation an optical process may take place in some
particular cases of electronic processes. Effects similar to photomagnetic anneal
O X I D E SPINELS 251

originally discovered in Si doped Y I G (Teale and Temple 1967) were observed in


Mg, Li, Li-Mn and other spinel ferrites (Holtwijk et al. 1970, Marais and
Merceron 1974, Hisatake and Ohta 1977, Bernstein and Merceron 1977).

3.2.4. Survey of experimental results


Most of the information on induced anisotropy effects comes from torque
measurements. In some cases additional knowledge may be obtained from the
study of disaccommodation and other relaxation effects (see, e.g., Braginski 1965,
Krupi'~ka and Zfiv6ta 1968).

C o 2+ containing spinels

As the anisotropic effect of Co 2+ is strong it usually outweighs other possible


contributions to the induced anisotropy. The magnetic annealing effect of Co 2+
ions was first observed in CoxFe3_xO4 by Kato and Takei (1933) and later on
frequently studied in this system by several authors (Penoyer and Bickford 1957,
Slonczewski 1958, Iida 1960, Iida and Inoue 1962, Iida and Miwa 1966, Palmer
1960, Iizuka and Iida 1966). For low cobalt concentration x the results may be
interpreted in terms of local ordering of both isolated Co 2+ ions and Co2+-Co 2+
pairs in octahedral sublattice (figs. 34 and 35). The presence of vacancies does not
change the magnitude of the induced anisotropy but the rate of ordering ~.-1 was
found to be proportional to the vacancy concentration provided this is >~10 5. For
lower concentrations of cation vacancies the activation energy is increased from
the usual value ~1 eV to ~ 2 eV indicating a more difficult diffusion mechanism.
In Ni-Co (Perthel 1962, Glaz et al. 1980), N i - F e - C o ferrites (Michalk 1968) and
N i - Z n - C o ferrites (Michalowski 1965) similar effects appear. The induced aniso-
tropy is usually lowered due to fluctuations in the local symmetry which primarily
concerns the G term. In table 25 the room temperature values of G and F found
for the system CoxNi0.46_xZn0.29Fe22504+ 7 after magnetic anneal at 350°C are
compared with those reported for CoxFe3-xO4. Let us note that the coexistence of
more kinds of ions yields possibilities for new anisotropic configurations to
appear. In Co substituted Li0.sFe2.504 a substantial reduction of the induced
anisotropy was observed when the B sublattice becomes ordered (lvanova et al.
1979).
In spinels containing both Co 2+ and Co 3+ ions in octahedral sites a preferential

T A B L E 25
Constants of induced anisotropy in Co containing ferrites.

System F (105 erg/cm 3) G(105 erg/cm 3) Ref.

CoxFe3-xO4 101 x 2 92.5 x 1


CoxNi0.a6-xZn0.29Fez2504+~ 0.06 + 100 x 2 7x 2

1. Penoyer, R.F. and L.R. Bickford Jr., 1957, Phys. Rev. 108, 271.
2. Michalowsky, L., 1965, Phys. Status Solidi 8, 543.
252 S. KRUPII)KA AND P. N O V ~

occupation of certain positions by Co 2+ may be achieved by the electron transfer


between Co 2+ and Co 3+ ions. If the concentration of Co ions on equivalent sites is
sufficient for a direct electron exchange Co2+ ~,~-Co3+, the activation energy is low
~ 0 . t e V . This is, e.g., case of cobalt rich C o - F e and C o - F e - N i spinels in-
vestigated by Iizuka and Iida (1966)- see figs. 37(a) and ( b ) - or Co 3+ containing
Co ferrite (Marais et al. 1970). Low temperature torque m e a s u r e m e n t s would be
necessary here to get the full frozen-in induced anisotropy. More often the spinels
with low concentration of both Co 2+ and Co 3+ are quoted in the literature (Sixtus
1960, Marais and Merceron 1959, Mizushima 1965). The Co ions are then separated
by many other ions, e.g. Fe 3+, so that the electron transfer involves also these ions
as intermediary. This results in rather high activation energy up to ~0.6 eV.

r , , , , J i , i

-o-o- be¢ore
rec~ucUon
-×-x- al~er
20 ~ o ~
.\ ,,
15
°d,,
' \°i
x
~Z
10
!
×
o
/ ~ttl,~.,,.
0 0.'2 0.'4 O.6 0.8 1.0
X

Fig. 37. The dependence of induced anisotropy on the composition in the system CoxFe3-xO4(Iizuka
and Iida 1966). The reduction decreases the concentration of Co3+ ions.

N i - F e and other iron rich spinels

Induced anisotropy and various after-effects were studied in several systems


containing Fe 2+ ions usually simultaneously with cation vacancies. Typical exam-
ples are iron rich Ni and Mg ferrites including magnetite (Motzke 1962, 1964,
G e r b e r and Elbinger 1964, Wagner 1961). In pure magnetite with a slight oxygen
excess an induced anisotropy of G type was observed (Knowles 1964) possessing a
relaxation time ~ 2 0 s at room t e m p e r a t u r e . This is compatible with the pref-
erential occupation of certain octahedral positions by vacancies. The magnitude
and temperature dependence of G may be brought into reasonable agreement
with the mechanism of a vacancy dipole-dipole contribution (eq. 34). In a m o r e
recent study (Kronm/iller et al. 1974) the corresponding after-effect was shown to
be composed of two processes which were attributed to different distributions of
Fe 2+ and Fe 3+ around the vacancy. Only recently Brtining and Semmelhack (1979)
have also reported F ¢ 0 and directly proportional to the cation vacancy concen-
OXIDE SPINELS 253

tration in their magnetite single crystal. The ratio F / G was practically in-
dependent of vacancy concentration ( F / G = ½).
In mixed spinels, particularly Ni-Fe and Mg-Fe ferrites, the relaxation spec-
trum corresponds to two distinct magnetic annealing processes (fig. 38). One of
them (observable around room temperature) is similar to that discussed above; in
addition to the G term it possesses also an F term proportional to p x ( 2 - x ) , x
and p denoting the concentration of Ni 2+ (or Mg 2+) ions and vacancies, respec-
tively. This indicates the active role of vacancy-Ni 2+ pair ordering. The other
process with the annealing temperature -300°C was usually found to be of pure F
type with F proportional to x2(2- x) 2. It does not depend on p. This points to an
ion pair ordering while the role of vacancies is believed to be similar as previously
discussed for Co 2+ substituted magnetite, i.e., they increase the diffusion rate and
lower the annealing temperature. For historical reasons these processes are
usually denoted as III (lower temperature process) and I in the literature. An
additional effect (II) lying in the region between I and III and reported in some
papers was shown to be due to cobalt impurities. The magnetic annealing spectra
of Mn-Fe spinels are a little more complicated (fig. 39) and were discussed, e.g.,
by Krupi~ka and Vilim (1957), Krupi~ka (1962), Marais and Merceron (1965) and
Braginski and Merceron (1962). In particular, an additional effect labeled as IV
with lower activation energy was observed below room temperature. It was
ascribed to vacancies forming some complexes, e.g., with Mn 3+ and Fe 2+ ions. A
similarly positioned induced anisotropy effect was observed in Ti or Sn substituted
magnetite and ferrites (Knowles and Rankin 1971, Knowles 1974) where for-
mation of some Me 4+ containing complexes (e.g. Ti4+-Fe 2+ pairs) may influence
both the vacancy migration rate and the resulting induced anisotropy. For Ti
substituted magnetite the corresponding anisotropy was found to be of G type
(Knowles 1974). Let us note that by properly adjusting the concentration of Ti 4÷,
Fe 2+ and cation vacancies the effect III may be suppressed so that practically only
IV remains (fig. 40); this is important for the ferrite materials design.

1000 f ~-o-o-

"~ 800 i

/
o

~<- 600

400 /
o o. . . . . ...~o/
200 /o /
o

0 100 200 300 400


T(°C)
Fig. 38. The dependence of induced anisotropy in Ni0.4Fe2.604 on the annealing temperature Ta.
Measurements were performed at -75°C after 10 minutes anneal at 7". (Motzke 1962 and 1964).
254 S. KRUPI(~KA AND P. NOVAK

0.4
i ,

~,o" , i i i

0.3 ,,, r~o 4,


0.2
Pf \ I',°,x
0.1
,vii>:
\/7!,o
o-o-~-n,0 ',,,100 200,°'W,300 400
.,: '~
-200 -100
T(K)

i i
. . . .
/o,,oo-O-'E
5000
0
0~00 ~ 0 _ 0 . 0 #"

o
I
3000 [
0
!
1000 oO.oop
~ 0~ I I

-200 -100 0 I00 200 300 400


T (K)

Fig. 39. Comparison of the disaccommodation spectrum and the induced anisotropy for M n 0 . 2 F e 2 . 8 0 4 . 0 i 4

(Marais and Merceron 1965).

~l~Ip(°lo) / ~

III ~
O.6 - ~ /

0.4 J

0.2 iis/II
~ \k N

-50 0 +50 +I00


Temperature °C
Fig. 40. Disaccommodation spectra for two Ti substituted MnZn ferrites; (A) Mno.68Zno.3-
2+ " 3+ .
Fe0.11Tlo.09Fe18204, (B) Mn0.64Zn0.3Fe~lTio.osFe3~O4;after Knowles and Rankin (1971). The DA maximum
for MnZn ferrites without Ti is indicated by a dashed line.
OXIDE SPINELS 255

A further process was reported to appear at higher temperatures (350°C ~< T, ~<
450°C) in Mn and Zn substituted magnetite and was ascribed to reorientations
involving also the ions in tetrahedral sites, i.e., Mn 2+ and Zn 2+ (Maxim 1969). The
activation energies are rather high here, approximately 2 to 2.5 eV.
In all processes mentioned above the presence of Fe 2+ ions is important because
they may considerably enhance the induced anisotropy by their own contribution.
As they coexist with Fe 3÷ ions in iron rich spinels they may also give rise to
induced anisotropy and after-effects of electronic origin via the redistribution of
valencies Fe2+ ~--Fe 3÷. A C magnetic losses are usually used for study of this effect
(e.g. Kienlin 1957, K6hler 1959). Unless the concentration of Fe 2+ is too low, the
activation energy is rather small, of the order of 10 -2 to 10 1 e V , and the kinetics is
interpreted in terms of the electron hopping mechanism.

Spinels containing M n 3+ and C u 2+

Besides valency redistribution these ions may contribute to the magnetic anneal-
ing process by reorientations of local Jahn-Teller distortions (section 1.6). The
most detailed study was p e r f o r m e d on the MnxFe3 xO4 system (Krupi6ka and
Vilim 1957, Merceron 1965, Zfiv6ta et al. 1966, G e r b e r et al. 1966, Broese van
G r o e n a u 1967, Broese van G r o e n a u and Pearson 1967, Marais and Merceron
1967, Krupi~ka and Zfiv6ta 1968, Y a m a d a and Iida 1968, Iida et al. 1968). The
induced anisotropy measured on two series of samples given as a function of Mn
content (fig. 41) demonstrates convincingly the effect of Mn 3+. The same effect is
reflected in the part denoted as C of the magnetic loss factor spectra (see fig. 42).
The corresponding activation energies lie between ~0.3 eV and 0.5 eV, close to
the activation energy deduced from the electrical conductivity which supports the
interpretation based on reorientation process correlated with the electron motion
(Krupi6ka and Zfiv6ta 1968). The peaks B and A in fig. 42 may be attributed to

1000 , , ,

800 ' ~ x

200 Ij
i i l
1.0 1.2 14 1.6 1.8
X
Fig. 41. Dependence of the induced anisotropy of MnxFe3-xO4+~, on the manganese content (Gerber et
al. 1966). Samples were annealed at 1200°C at 760 mm Hg (a) and 10-2 mm Hg (b).
256 S. KRUPI(~KA AND P. NOV/~'(

0.4 ~ B ~ C

.8~. I I \ I
III I~ ' s+ I \ I
Ill/ ', / \1
A~
lltt I 0.81
/ X \1 '1

o lbo ebo ado r (K)


Fig. 42. Magnetic losses vs T for MnxFe3 xO4+7 system. Figures show manganese concentration x; 1+
and 1- correspond to x = 1 with 3' positive and negative respectively (Z~v~ta et al. 1966, 1968).

Fe 2+~--Fe 3+ valency redistribution and to reorientation of free Mn 3+ distortions


(i.e. not stabilized by electrons or other defects) respectively. The coexistence of
peaks B and C in certain compositional range points to the coexistence of Fe >
and Mn 3+ ions. Relaxation peaks similar to C were also found in other Mn and/or
Cu containing spinels: M n - Z n (Okada and Akashi 1965, Giesecke 1959), Mn-Mg
(Krupi6ka 1960), M n - C u (Krupi~.ka and Z/w6ta 1968). In Mn substituted Li ferrite
(Marais et al. 1972) the magnetic annealing effect of Mn 3+ ions was found to be
enhanced by clustering of these ions.

3.3. D y n a m i c s o f magnetization

Many ferrimagnetic spinels may be treated as collinear two sublatice ferrimagnets.


In such systems the theory of ferrimagnetic resonance (e.g. Keffer 1966) predicts
two resonance frequencies to+, w to occur. Neglecting the effects of anisotropy
and demagnetization field, the lower one is given by
to+ = y e f f H , (37)
with
Ye~ = ( M , - M2)/(M~/y, Mz/y2),

where MI, M2 are the sublattice magnetizations, and yl, 3/2 are respective
gyromagnetic ratios. It is analogous to the simple ferromagnetic mode, except that
the effective value Yen has to be substituted for the gyromagnetic ratio 3/.
The higher resonance frequency to_ is related to the so-called exchange
resonance frequency
toex = n (72M1 - yIM2) , (38)
where n is the intersublattice molecular field coefficient.
OXIDE SPINELS 257

For H ~ AHex = n(M1 - M2),

w_ = Wex-- y2~H, (39)

with

"Year= [T2(M1/3'1)- y1(M2/'Y2)]/(M1/yl- M2/T2),

which demonstrates the dependence of both we× and o) on the strength of the
inter-sublattice exchange. As this is rather strong in most of spinel ferrites ~Oexis
expected to fall into the infrared rather than microwave region. This is presum-
ably one of the reasons why the exchange resonance was not yet unambiguously
detected in spinel ferrites. M o r e o v e r the exchange resonance may be excited only
if Y~ ~ Y2 (Schl6man 1957).
The effective gyromagnetic factor ")/eftin eq. (37) is connected with the effective
spectroscopic splitting factor gen through the relation

Yen = e/(2mc)ge~r = e / ( 2 m c ) ( M l - M 2 ) l ( M 1 / g l - M21g2). (4o)


For S-state ions (Mn 2+, Fe 3+) g is close to its free electron value 2.0023. In other
cases the spin-orbit coupling modifies the value of g. Effective g factors as
obtained from F M R in several spinels are given in table 26.
The width A H of the resonance line is in most ferrites of the order of several
tens Oe. Nevertheless in very pure Li, Mg and Ni ferrites AH--~ 1 - 2 0 e was
observed (Yakovlev et al. 1971). On the other hand, values of A H up to 1000 Oe
may occur in ferrites containing Fe 2+ in the B sublattice. The values of A H in
ferrimagnetic spinels are summarized in table 27.
T h e r e are several relaxation processes which are responsible for the linewidth
in spinel ferrites (for detailed discussion see Sparks 1964, Patton 1975). In
materials containing magnetic inhomogeneities such as surface roughness, grain
boundaries or atomic disorder the two-magnon scattering is important. In many
ferrites, however, the effect of so-called slowly relaxing impurities dominates. This
is supposed to be the case, e.g., for systems containing Fe 2+, Co 2+ or Mn 3+ in the B
sublattice or Ni 2+ ion in the A sublattice. Relevant mechanism requires the
presence of two or more low lying anisotropic levels. The magnetization pre-
cession then modulates the energy separation between the levels. As a con-
sequence, the thermal equilibrium population for each level changes with a period
of precession. The induced transitions between the anisotropic levels then give
rise to the relaxation effect. Note that a similar mechanism applies to the valence
exchange Fe 2+ ~--Fe 3+ (Clogston 1955, Teale 1967). In this case the energy levels
are connected with the hopping of a Fe 2+ ion between the B sites having different
orientation of the local trigonal axis. The linewidth corresponding to the slowly
relaxing impurities is strongly t e m p e r a t u r e dependent exhibiting one or two
maxima (fig. 43).
258 S. KRUPICKA AND P. N O V ~ K

TABLE 26
Values of the g factor from FMR studies.

System T (K) A (cm) g Ref.

Fe304 130 3.35 2.08 1


1.25 2.09
293 3.35 2.17
1.25 2.13

MnFe204 4.2 3.2 2.060 2


77 3.2 2.019 (3)
300 3.2 2.004 (2)

(Mn, Zn)Fe204 4.2-300 1.25-10 2.00 3

CoFe204 295-310 0.37-0.64 2.6 (2) 4


363 1.25 2.7 (3) 5
473 1.25 2.27

NiFe204 293 1.25 2.19 6

Nio.95Fe2.os04 4.2-290 1.25 2.2 7

Nio.75Fe2.2504 4.2 1.25 2.16 7


85 1.25 2.19
290 1.25 2.13

CuFe204 77 1.25 2.44 8


300 1.25 2,09

MgFe204 77 1.25 2.04 9


473 1.25 2.005

Lio.sFezsO4 300 1.25 2.005 10

1. Bickford, L.R., 1949, Phys. Rev. 76, 137.


2. Dillon, J.E. et al., 1955, Phys. Rev. 100, 750.
3. Schl6mann, E., Conf. Magnetism Magn. Materials, Boston, 1956.
4. Assadourian, L. and L. Silber, 1976, AIP Conf. Proc. 29, 684.
5. Tannenwald, P.E., 1955, Phys. Rev. 99, 463.
6. Yager, W.A. et al., 1950, Phys. Rev. 80, 744.
7. Yager, W.A. et al., 1955, Phys. Rev. 99, 1203.
8. Miyadai, T. et al., 1965, J. Phys. Soc. Jap. 20, 980.
9. Kriessman, C.S. and H.S. Belson, 1959, J. Appl. Phys. 30, 170.
10. Schnitzler, A.D. et al., 1962, J. Appl. Phys. 33, 1293.
OXIDE SPINELS 259

TABLE 27
Values of AH in spinel ferrites.

System T (K) h (cm) AH (Oe) Ref. Remark

Lio.sFe2.504 134 6 0.88 1 Ordered


300 6 1.68 state

Mno.4zMgo.61Fel.9504 290 6.1 2.5 2

MnFe204 20 1.25 20 3
300 1.25 80

Mnl.03Fel.9704 15 3.2 173 4 Direction


290 3.2 38 [111]

Mno.asZno.55Fe204 290 1.25 35 5

CuFe204 290 1,25 120 6


300 1,25 60

NiFe204 290 1,25 35 7


290 3,2 2 8

Ni0.95Fe20504 4.2 1.25 30 9


85 1.25 55
290 1.25 50

Nio,vsFe2.2504 4.2 1.25 40 9


85 1.25 120
290 1.25 140

MgFe204 290 6.1 2.3 8 Inversion


= 0.89

Fe304 290 1.25 4300 10 Direction


[100]

1. Remeika, J.P. and R.L. Comstock, 1964, J. Appl. Phys. 35, 3320.
2. Lyukshin, V.V. et al., 1976, Izv. Akad. Nauk SSSR Neorg. Mater. 11,285.
3. Heeger, A.J. et al., 1964, Phys. Rev. 134A, 399.
4. Watanabe, Y., 1974, J. Phys. Soc. Jpn. 37, 637.
5. Gait, J.K. et al., 1951, Phys. Rev. 81, 470.
6. Miydai, T. et al., 1965, J. Phys. Soc. Jap. 20, 980.
7. Yager, W.A. et al., 1950, Phys. Rev. 80, 744.
8. Yakovlev, Y.M. et al., 1971, Fiz. Tver. Tel. 13, 1151.
9. Yager, W.A. et al., 1955, Phys. Rev. 99, 1203.
10. Bickford, L.R. Jr., 1950, Phys. Rev. 78, 449.
260 S. KRUPICKA AND P. NOV./d{

aH
(Oe)
800
i o DI,)

600 \ • Doo]

20O

5'o 1~o 1~o 260 2~o ~bo


T(X)
Fig. 43. Temperature dependence of the FMR linewidth for MsI1.46Fel.5404system (Zfiv6ta and Novfik
197a).

4. Other physical properties

4.1. Electrical properties

4.1.1. D C electrical conductivity and Seebeck effect


Most of the oxide spinels not containing transition metal ions are very good
insulators at room temperature. The A13+ spinels MgA1204, ZnA1204 and others
may be mentioned as examples possessing electrical conductivity of the order of
10-6O-~cm -1 even at temperatures ~900°C (Bradburn and Rigby 1953). This
behaviour may be understood as due to a large energy gap (often of several eV)
between the occupied valence band primarily formed by the oxygen 2p states and
the empty conduction band. The actual values of electrical conductivity and the
activation energies are then usually controlled by impurity levels within the gap.
With the presence of the transition metal ions additional energy levels and/or
narrow bands are introduced usually also lying in the gap. This need not
necessarily change the picture of the electronic charge transport very much
provided that transport within the partly occupied d levels (bands) themselves
does not d o m i n a t e . In particular, many oxide spinels with 3d n ions in the A
positions only possess a very low conductivity (fig. 44). The same is true for pure
stoichiometric spinels having only one kind of ions on equivalent crystallographic
sites such as the normal spinels ZnFe204 and CdFe204, or ordered Lio.sFezsO4. A
considerable increase of the electrical conductivity is usually connected with the
combined effect of disorder and the presence of cations able to change easily their
valency states or actually coexisting with different valencies in the material.
OXIDE SPINELS 261

T °C
600 700 800 900 1000
, '//!
x
104.5
/
R(#) (bV ×
105.0
%/.o/
105.5
/ ./
I× • I ~%'b
.n~ . ~ ° I
!,'o x /
./ ./'.k_~Y"
106.0 /o. / f/
/ / Zv
106.5 //
/
tOzo
17 lh D b
I 0 7 T ( K "1)
Fig. 44. Temperature dependence of the resistance for some Al spinels (Bradburn and Rigby 1953).

Typical examples are bivalent and trivalent Fe or Co ions in the B positions. The
charge transfer may then be effectuated by hopping of electrons or holes between
equivalent ions (valency exchange) which is a rather easy process (see, e.g., Verwey
i951). The sign of the Seebeck coefficient (thermopower) is often helpful in
clarifying the nature of the dominating charge carriers and the mechanism of
their motion. A special case seems to be the vanadium spinels where besides
hopping a nearly band-like conductivity may occur characterized by a drop of
both resistivity and activation energy when the distance between neighbouring
vanadium ions in B sites approaches (or decreases below) certain critical value
(table 28, Rogers et al. 1963). A metallic conductivity was found in LiVzO4 ( V - V
separation ~2.91 A, Rogers 1967).
The most relevant for the scope of this book and perhaps the most often
studied are the electrical transport properties of Fe spinels (ferrites) and their
solid solutions. We shall start with magnetite Fe3+[Fe3+Fe2+]O] - which represents
a rather singular case due to its high R T value of the DC conductivity
(~250 fV~cm -1) and the peculiar character of the conductivity versus temperature
dependence. Other ferrites may be regarded as belonging to systems of solid
solutions MxFe3-xO4 where M stands for (usually bivalent) cations substituting
Fe 2+ in magnetite or for combinations of such ions. Some of the substituting
cations may also exist in higher valency states, e.g., Mn 3+ or Co3+; the relevant
systems are then to be extended to include the mixed spinels with x > 1.
In spite of a large amount of existing experimental data only few of them which
have been obtained on well defined single crystals may be used to draw quan-
262 S. KRUPI(~KA AND P. NOVfid~

TABLE 28
Crystallograplaic and electrical conductivity data for vanadium spinels (after Rogers 1967 and Rogers
et al. 1967).

Activation energy of
electrical conductivity
Formula Lattice constant (A) V-V separation (A) (eV)

Mn[V2]O4 8.522 3.014 0.37


Fe[V2]O4 8.454 2.990 0.25
Mg[V2]O4 8.418 2.974 0.18"
Zn[V2]O4 8.410 2.973 0.16
Li[V;]04 - 2.91 metallic*

* single crystal.

titative conclusions about the n u m b e r of carriers, their mobility, activation


energies etc. The other ones, particularly those related to polycrystalline samples,
are to be taken with caution and may be explored mainly in a qualitative way. The
discussion of the electronic transfer will be limited to conductivity and Seebeck
effect. The evaluation of the Hall mobility is usually difficult due to a large
contribution of the spontaneous magnetization for which no reliable theory seems
to be available. Moreover, it has been argued that in the case of small polaron
hopping which probably is the predominant transport mechanism in ferrites and
related spinels no simple relation exists between Hall and drift mobilities (Adler
1968). It is usually admitted that the drift mobility is rather low, ~0.1 to
1 c m W - l s -1 for magnetite and much lower (10 -4 to 10-Scm2V-Is 1) for com-
positions approaching the stoichiometric ferrites (Klinger and Samochvalov 1977).
As a rule, these values are deduced from conductivity or Seebeck effect
measurements on the basis of some model, and no reliable independent method
has been used for their determination. Even the magnetoresistance experiments
though occasionally reported in the literature have not been interpreted from the
point of view of electrical transport mechanisms (for a review see, e.g., Svirina
1970).

M a g n e t i t e a n d substituted magnetite
The log o- versus T -1 plots for magnetite covering a broad t e m p e r a t u r e region are
shown in figs. 45a and 45b. The Hall coefficient and mobility are displayed in fig. 46.
At elevated temperatures (1500 K > T ~> Tc = 858 K) magnetite exhibits a semi-
conductive behaviour with thermally activated conductivity, which may be fitted
(Parker and Tinsley 1976) by the formula:

~r = A T -~ e x p ( - q / k T ) , (41)

with A = 490 f~-lcm-lK and q = (99-+3)× 10 3eV. Note that eq. (41) can also
explain the m a x i m u m in cr versus T observed at ~1100 K. In the vicinity of Tc
(usually somewhat below) o- begins to depart from eq. (41); in a certain tem-
OXIDE SPINELS 263

103 . . . . . 250 ~ - p ~ . . . . . .

102 ooooo-O--o~ o~ o~.


0.0~ 0
~ 200 "-O~ojO
10~
~(858K) I ~ 1oo~5°r
o

I0-'
O,lS eVJ" X . 50 I
i Lo~/ ternperef;ure transition
10-2 o °"~t ' 4o0 . .800
... 12oo (K)00',6

10-3
o
104

10 -5
\o
10-6 °~o
, , f.o3e~,~o..
10-7
0 i2 14 /6 18 20 22 24
loao/r (K-')
Fig. 45a. Conductivity vs temperature for a single crystal of Fe304 (Miles et al. 1957).

-200
I

-205

"-•-210 ~o

o...-
-215
\2 o
~o
o

-220
-230
o /

-225
[! \ '%/
? '° -235
o NO

~°,,o 4
i - ~ o - o - o . ~ -240
r~ 2 3 4
1ooo/T (K-~)
Fig. 45b. Resistivity vs temperature for several specimens of magnetite (Parker and Tinsley 1976); (1)
Stoichiometric single crystal, (2) oxygen defficient polycrystalline specimen Fe300.3988, (3) stoichiometric
polycrystalline specimen, (4) single crystal according to Smith (1952).
264 S. KRUPICKA AND P. NOV/~K

10~

o~
&
i0 ~
(cmalc)

10

1
bo
///
1o-1 rv

lo-2

10-3
O_ o --0

104 I t I I I

0.50 i i i E

~u
0.20
(cm2/VS)o 1o o~ ["

0.05 0-. 4
I
0.02 I [ I
2
IO00/T (K 4)
Fig. 46. Ordinary Hall coefficient R0 and Hall mobility/zH vs temperature (Siemons 1970).

perature interval it behaves metal-like with negative temperature coefficient and


in the vicinity of room temperature it passes a new maximum. At Tv = 119 K
magnetite undergoes a phase transition (the Verwey transition ) accompanied by a
sudden decrease of conductivity of about two orders of magnitude. Below this it
behaves like a semiconductor with a temperature dependent activation energy, at
least down to ~10 K. Attempts have been made (fig. 47) to fit the or versus T
dependence in this region to Mott's formula;

o" = A e x p ( - B / T W 4 ) , (42)
derived for variable range hopping (Mott 1969).
It was early recognized (Verwey and Haayman 1941) that the transition in
magnetite at 119 K is connected with some kind of electronic charge ordering and
a model was proposed for it based on regular arrangement of Fe 2+ and Fe > ions
in rows parallel to [110] and [110] directions, respectively (Verwey ordering). The
more recent models, partly based on new neutron diffraction, NMR and M6ss-
bauer data either abandon the presumption of definite ionic valencies (Cullen
and Callen 1973) or correspond to more sophisticated ordering schemes of Fe 2+
OXIDE SPINELS 265

~ T(K)
n n I I n I

2 o0%0
0 o
%%
.t%o
-2 "to
%
E o

,o,
%
t~ -6 'oo
c3

oI
-8
q

-10
o

-12
%

-14 ~o
\%%o OoooO

O.2 Oi3 0 .'4 0'e 05


T@ (K)@
Fig. 47. log cr as a function of T -1/4 for Fe304 single crystal, after Drabble et al. (1971).

and Fe 3+ (Hargrove and Kfindig 1970, Fujii et al. 1975, Shirane et al. 1975, Iida et
al. 1976-1978, Umemura and Iida 1979). These models assign the low temperature
phase rather as monoclinic than orthorhombic. It is clear that any model explain-
ing the electronic conductivity in magnetite also has to explain (or at least to be
compatible with) the Verwey transition and vice versa. It must also account for
the anisotropy of o- below this transition (Chikazumi 1975, Mizushima et al. 1978).
The common feature of recent models of conduction in magnetite is the splitting
of electronic 3d 6 levels of Fe2+(B) in the ordered phase by an energy gap of
~0.1 eV; only states below this gap are populated at 0 K because the number of
Fe 2+ is half the number of the octahedral sites. The electronic charge transport is
then effectuated by carriers either created by thermal activation across the gap or
introduced by impurity atoms or oxygen non-stoichiometry. The separated levels
are usually supposed to form some kind of narrow subbands that overlap above
the Verwey transition. In the simplest case a tight-binding scheme combined with
Coulomb repulsion was used which leads to a Hubbard-type Hamiltonian for
description of the situation (Cullen and Callen 1970, 1973, Fazekas 1972). More
refined theories include also polaronic effects (electron-phonon interaction) and
other short-range energy contributions, included spin correlation and exchange
effects (Haubenreisser 1961, Klinger 1975, Klinger and Samochvalov 1977 and ref.
therein, Buchenau 1975).
266 S. K R U P I C K A A N D P. N O V A K

Due to the polarization effect upon their surroundings the electrons are usually
supposed to be not entirely free to move below the transition and their transfer is
described as a polaron hopping process, perhaps except at the lowest tem-
peratures (<10 K) where polaron coherent tunelling might dominate the conduc-
tion. Above the Verwey transition the long-range order disappears and the energy
gap between both subbands collapses accompanied by a sudden increase of the
carrier number. The complicated temperature course of cr is then believed to
result from a combination of narrow band conduction, short-range correlation of
polarons and scattering processes connected with spatial charge and spin fluctua-
tions. The loss of magnetic order at Tc seems to modify also the electronic
structure (Parker 1975). The recent M6ssbauer study (Lu-San Pan and Evans
1978) indicates that at T > Tc the Fe z+ valency states may appear also in the A
sublattice which may play a role in changing the character of the conduction
process.
The effect of small substitution or oxygen non-stoichiometry is twofold: The
ordered phase becomes imperfect which lowers the temperature of the Verwey
transition and makes it disappear for a certain critical impurity concentration.
Besides, the ratio [Fe2+]/[Fe3+] is changed which introduces carriers into one of
the split subbands (holes into the lower filled band or electrons into the empty
higher one depending on the kind of impurity). In fig. 48 the thermopower versus
temperature is plotted for single crystals with various degrees of oxidation. The

40 "~l

-40g ~" • • I Ill

-oi ... I,IIII


~/,~ l - ~-~,~ • " C

_,of \O.o/ o/
-200 "o
/ '
eo s2o 1so 200 2~o 80 I2o
T(K)
Fig. 48. Absolute thermoelectric power vs temperature; after Kuipers (1978). (a) Experimental data
for various magnetite single crystals. T h e lines are only meant as a guide to the eye. T h e vacancy
concentration decreases from A to E. (b) Calculated values according the model of Kuipers and
Brabers (1976), T denotes the cation vacancy concentration per formula unit.
OXIDE SPINELS 267

t , i ,

,~ x : 8 . 1 0 -3
x x=3xlO -3
v ×=O
+ x' = 10 -4
0 ~: : 1 0 "a
o x : 4 x l O -4

,,?

-2
\;:<:\
0,% *,-o,
-3 ,%\\

-4
a
i i L i I°x
I0 12 14 16
103/T (K -')
60

(lJ V. g -~

2~

0
)
-20

:~
o
Q.
-60
t
£
- I00
co
ox:O
~,~. . x = l O -4
'~ ,~ x = 4 . 1 0 -4
~. ~L • x= 10 -3
F.. PA
x x : 3 ~ 10 -3
j + x = 8~10 -3
- 180

-22o
b
~o
/ 4.

1~o 1~o
I
2bo 240
T(K)
Fig. 49. Electrical properties of Ti-substituted magnetite Fe3-xTixO4 (Kuipers 1978). (a) Conductivity
vs temperature relation. (b) Absolute thermoelectric power vs temperature.
268 S. KRUPICKA AND P. NOV,~K

tendency of the Seebeck coefficient to become positive toward low temperatures


is in agreement with the presence of positive charge carriers (holes) introduced
into filled Fe 2+ subband by increasing the oxygen content and with possible
temperature variations of their mobility. This also demonstrates the sensitivity of
the ordered phase upon the oxygen non-stoichiometry; above the transition, on
the other hand, the influence is poor because both subbands overlap and the
relative change in the carrier number is very small. The impurity ions have similar
effect as the cation vacancies: cations with valencies 1 or 2 diminish the number of
Fe 2÷ and introduce holes; cations with higher valencies act as donors. But in both
cases the effect is again small above the transition temperature. As an example of
the higher valency substitutions the behaviour for Ti substituted magnetite is
given in fig. 49. The lowering of the transition temperature introduced by various
dopants is demonstrated in fig. 50. The critical concentration for the disap-
pearance of the Verwey transition is usually about 0.1 per formula unit.

z~T
o

• AL

Ni
~, Co

Cr
o\!8


o

Oa
M9
Zn
(K) (K) "'"

-5 -5

-10 -I0

-15
-15

\
o.b7 doe 0.03 o 0.01 ?.02 0.'03
X
Fig. 50. Lowering of the transition temperature Tv by various dopants (Miyahara 1972).

Stoichiometric and iron rich ferrites


If the concentration of ions substituted in magnetite increases the resultant
fluctuations of the electrostatic potential make the free motion of electronic
carriers more and more difficult. With certain critical concentrations practically all
carriers, i.e. the Fe 2+ valency states, are localized in local potential minima for
T ~ 0 K. This is called the Anderson localization (Anderson 1958, 1970, Klinger
and Samochvalov 1977). At finite temperatures, however, they may be thermally
O X I D E SPINELS 269

activated to hop to other places (ions) and contribute in this way to the electrical
transport. The existence of hopping in the Fe2+-Fe 3÷ pairs in condition of a
fluctuating local potential was demonstrated in M6ssbauer spectra of Ti sub-
stituted Zn ferrites (Van Diepen and Lotgering 1977).
The o- versus T dependence may usually be fitted by the exponential law

o-= A T -~ e x p ( - q / k T ) /3 = 0, 1 or _3
2~ (43)

in a fairly broad temperature region. For bivalent substitutions the activation


energy q varies from several hundredths of eV for Fe 2+ content to ~0.2 to 0.6 eV
for compositions close to stoichiometric ferrites MFe204 (fig. 51). As was shown
by Miyata (1961) the preexponential factor A is almost fully determined by the
Fe 2+ content and practically does not depend on the kind and concentration of the
other cations present.
At low temperatures an application of eq. (43) to the experimental dependences
would lead to gradually diminishing the effective activation energy which in the
low temperature limit can be approximated by the Mott formula ( 4 2 ) - s e e , e.g.,
Simga and Schneeweiss (1972), Kuipers (1978). Representative room temperature
values of electrical conductivity are given in table 29. According to the small

T A B L E 29
Representative room temperature electrical data of spinel ferrites.

Composition (r (~-1 cm a) q (eV) a (~VK l) Ref.

Fe304 (") 250 nearly -48 1


metallic
Lio.sFe2.504 (a) 3.6 0.17 -500 2
Mno.5Fe2.504 (a) 39.8 0.06 - 118 3
Mnl.0Fe2.oO4 (a) 0.45 0.08 -350 3
Mn12Fe1.804(~) 1.6 × 10-4 0.35 - 1000 3
Ni0.67Fe2.2304(a) 6.65 0.08 - 328 4
Ni0.96Fe2.0404(a) 0.016 0.28 -210 2
CoFea.94Ti0.0304 (a) 0.13 0.14 - 558 5
Co0.99Fez0104 (b) 1.4 x 10-3 0.22 -460 6
Cot.01Fel.9904 (b) 8.0 x 10-8 0.56 +800 6
Cu0.97Fe2.03O4(b) 0.01 0.15 -- 40 7
MgMn204 c°) 2.5 × 10-5 0.36 +97 8

(~)single crystals
(b) polycrystals

1. Kuipers, A.J.M., Thesis, Technical Univ. Eindhoven (1978).


2. Austin, I.G. and D. Elwell, 1970, Contempt. Phys. 11, 455.
3. Simga, Z. et al., 1972, Proc. 11th ICPS Warsaw, p. 1294.
4. Yamada, T., 1975, J. Phys. Soc. Jap. 38, 1378.
5. Yamada, T., 1973, J. Phys. Soc. Jap. 35, 130.
6. Jonker, G.H., 1959, J. Phys. Chem. Solids 9, 165.
7. Rosenherg, M. et al., 1966, Phys. Status Solidi 15, 521.
8. Rosenberg, M. and P. Nicolau, 1964, Phys. Status Solidi 6, 101.
270 S. KRUPI(~KA AND P. NOV/MK

b-

0
"a

x.
x
x

o
xx x q =
;.= ~~ 0,- ["-:"
]
o
o
-,%
~o ~ ~ 0
" h
°\°,,. \o N× h
0% xu \ h\
~'o Xx ,q
--o \ -,, h
"°\ ° ~X ~\ h
°\ \ 4- >~
0 ~ o,., xx
qb~ O ~ % o, & h
¢.~ -0\,%, ° \ h

E 0 x 'q []
X

i i i i i i

"" ~' ,'7 ,'? ~i -4-


i
' ,.~ z:Z~ A

•= ~"
~ t"-

i i i . i i %1 i

~?1 ' ob
Oo/,
o

I,/
o

~E
+~

~N.. ~
~ 0 0

~' o/ !I ~ 1 ~ '-~ ~o ~,

~
o
l---.B op
II o
o.O°o'~
/.Oo
.7V.
/ / d°
,
o II oqo / ,6 P
~; l o / ¢ / /.o~ z
~ I / i o--i o _o
/
Io oo / . P / /
*o Io, - / o l . o / t ,o io
' ~ o, ,o/iA,o~/,~ .,~.
/ o io,/o~,%.,o~ oo/iO _o.O.O
/o 7, ??b%" :o -o'~ "-~" o ---°°°~ -
o ,d i,%o.O / oo~.O- ~
~1 i ,po i i~ i I I
% ~ ~,

+ +o ~o

+ ~
,--4 ~ . ~ ,-
~ZN
OXIDE SPINELS 271

polaron hopping model which is believed to be relevant for the conduction


mechanism of ferrites, the activation energy q is the sum of the energy qn needed
for removing the electron from an Fe 2+ (i.e. the binding energy of the polaron)
and of the mobility activation energy q~ connected with transferring an electron
between Fe 2+ and Fe 3+ ions. q. may be directly deduced from the Seebeck
coefficient versus T -1 dependence; examples of these dependences are given in
fig. 52.
].0 i

~Oo a

b.6 ox~o
E

-o
0.4 \

io ~ o~o-o ........ 3
O. 2 ~-,,~o..~o.-e..o _o. o~ o - o - - o _ o _ c-.- 4

[~7
4 ~ ° '''~" 0~°~0 O,

r &
o.o_o.o-o--o--
L
o--O--°
~ I
100 300 500 700 900
T(K)

4 o CoFe~.9~Tioo30~
.o" x Nio.e~.Fe2.330,~
Ioc 2.3 ) / °'° °° ~' Mn°~Fezs04
/o o MnFe2q
3 o
/° o
/
/e a a x~O X / X X X
X
~ a
2
~ -×I"
~.i / x~ x

z& z~ z~ z~

b
o ' ~ ' ;
1'o ' 1~ ' ~ ' 8 '
lOa/T (K -1)
Fig. 52. Seebeck coefficients vs temperature relation for ferrites with various Fe 2+ content. (a)
Polycrystalline samples (after Klinger and Samochvalov 1977):
(1) Ll0.sFe0.0olFezsO4,
" 2+ 3+ . 2+ 3+04;
(5) N"10.sFe0.2Fe2
(2) (N10.3Zn0.7)o.99Fe0.mFe2"
2+ 3+04 ,. (6) Ni0.6Fe0Z74Fe~+O4;
(3) (N10.3Zn0.7)0.964Fe0.036Fe2"
2+ 3+O4," (7) Zn0.4Fe02+6Fe~+O4;
(4) (N10.3Zn07)0.89Fe0.nFe2'
2+ 3+O4,. (8) Fe~+Fe3+O4.
(b) Single crystals, measurements by Y a m a d a (1973) (©), Y a m a d a (1975) (x), Simga et al. (1972) (A),
(Eli).
272 S. KRUPICKA AND P. NOV~d(

12

tog ~
10

9 O-
O~D,...,,~
8 - D - -

/
a
I I
0.02 o.o~ 0.'o6 o.b8 o.'1

E,

Fe2 +

f,.,,/r~2 +

6o 2+

Ni 2÷

b
Fig. 53. (a) Influence of Mn and Co additions on the resistivity of Mg and Ni ferrites (Van Uitert
1956a, b, 1957): (1) MgFel.aMnxO4; (2) NiFel.gMnxO4z; (3) NiFel.gCoxO4_*. (b) The relevant level
diagram (Elwell et al. 1963).
O X I D E SPINELS 273

When approaching stoichiometric ferrites MFe204 (M bivalent) both resistivity


and activation energy are increased due to the lack of Fe 2+. This is because the
energy levels of M 2+ ions are usually situated below those of Fe z+. From experi-
ment we have the following sequence of levels:

N i 2+ < C o 2+ < M n 2+ ~ F e 2+ ,

which at least at low temperatures prevents the occupation of an Fe 2+ state unless


lower lying levels of M 2÷ are filled up (Lord and Parker 1960, Parker and Smith
1961, Elwell et al. 1963, 1966). This has been exploited in preparation of high
resistivity ferrites using small additions of Co or Mn to, e.g., Mg or Ni ferrites (see
fig. 53; Van Uitert 1956a, b, 1957).

MnxFe3-x04 and CoxFe3-x04 systems


Some transition metal ions, e.g. Mn or Co, may exist in spinels like iron with
various valency states. This enables us to extend the MxFe3-~O4 systems (M = Mn,
Co) into their iron deficient regions. The relevant dependences are given in figs.
54-56. In CoxFe3-xO4 the Seebeck coefficient changes its sign at x = 1 where the
valency exchange Fe2+~--Fe 3+ is replaced by Co2+~-~--Co3+; while the first cor-
responds to the electron (n-type) transport in the second one the Co 3+ holes hop
from one Co 2+ to another. In MnxFe3 ~O4 the situation is more complicated
because together with x the degree of inversion also changes (see section 1.5).
F e 3 0 4 is an inverse spinel but MnFe204 is an almost normal one. It follows that
there are still enough Fe ions in the B sublattice for 1 < x ~< 2 to play an active
role in the conduction mechanism; it is only necessary to create carriers by
exciting some electrons to Fe 2+ levels. This is the reason why the Seebeck
coefficient does not change its sign at x = 1.

I0
8
6
(kT) 4
2
0
-2
-,~
-6
-8
-10
-12
-14
It) 5 0 S lO
°Co %Fog

Fig. 54. Thermopower e~-= e0T for COl-xFe2+xO4; 0.1 > x > - 0 . 1 (Parker 1975). Measurements of
Jonker (1959) taken near 370 K. The full line was calculated by Parker.
274 S. KRUPICKA AND P. N O V ~

8 i ~ i t i

z ~...x.°l o.z
In R i..>o o" oO } qCov)
6 0.6
_o.-~ ~o~'~i
5 __._....-..'" ~1 0,5
/I

e "~" ~'O~'o_ oe
1 o.1

L I I I I
0 1.10 ~05 1.00 0.95 0.90
x
Fig. 55. Room temperature resistivity p and activation energy q for CoxFe3-x04 (Jonker 1959).

i I

/
0.5 " Miyata 1961 o

o x Lof:gering 1964

/
q (ed z~ Sim[a 1972

0.4

z~

/
/o
0.3 t°

0.2

0.1

05 10 15
x in Mn, Fea_~O~
Fig. 56. Activation energy of conductivity for MnxFe3-xO4 (Lotgering 1964, Simga et al. 1972).
OXIDE SPINELS 275

4.1.2. Dielectric constant


There are not many reliable data on dielectric constant E of ferrites and related
spinels measured on the single crystals. The intrinsic E values were usually found
to lie between 8 and 20. The very high dielectric constants often observed at low
frequencies (fig. 57) have been ascribed to the effect of heterogeneity of the
samples, i.e., pores and/or surface layers on grains, causing poor electrical contact
between them (Koops 1951, Heikes and Johnston 1957, Krausse 1969). Sometimes
some electronic polarization effect is supposed to be connected with the conduc-
tion hopping mechanism itself (and with some microscopic inhomogeneities
related to it) which also could contribute to the low frequency dispersion of E
(Rabkin and Novikova 1960, Rezlescu and Rezlescu 1974). For a general outline
see Bosman and Van Daal (1970). At higher frequencies (cm wavelengths) the
measured values may be regarded as insensitive to both of these contributions and
they are usually taken as actual intrinsic dielectric constants corresponding to
normal ionic and electronic polarizations. Some typical E values are given in table
30.
In the frequency region of the lattice vibrations the ionic polarization becomes
slow and damps out (infrared absorption) while the electronic polarization is fast
enough to persist to the region of electronic excitations in near infrared and
visible parts of the spectrum (crystal field and charge transfer transitions). To
estimate the relative contribution of ionic and electronic polarizations, E
measured in the near infrared and at microwave frequencies have to be compared.
For NiFe204 the near infrared value is e ' = 7.0-+ 0.5 at A ~ 5 ~m (Miles et al.
1957). Taking E ~ 10 for the microwave region (see table 30) the contribution of
ionic polarization would be ~3. As far as the temperature dependence is
concerned, E usually increases with increasing temperature together with the
electrical conductivity (Ioffe et al. 1957). Some Cu containing spinels were
reported to show anomalies, however, in both temperature and frequency

i i i
1o5 !

.104 2

103 3

102 -

I0 6
1 1 1 t 1

10 2 103 104 105 10 6 10 z 108


Fr'equeney in els
Fig. 57. Dielectric constant e for various polycrystal]ine ferrites measured as a function of frequency
(Van Uitert 1956b): (1) M n - Z n , (2) Cu-Zn, (3) N i - Z n , (4) NM1350, (5) N1250, (6) NM1250.
276 S. KRUPI(~KA AND P. N O V ~

TABLE 30
Microwave values of complex dielectric constant for some ferrites.

Formula Frequency (GHz) E' E" Ref.

NiFe204 4.5 13.40 3.520 1


quenched
NiFe204 4.5 8.88 0.155 1
slowly cooled
Ni0,sZn0.sFe204 10 15 1 2
Ni0.gZn0.tMn0.02Fe204_+ 9.2 12.17 1.38 × 10-4 3
MnFe204 4.5 9.30 0.475 1

1. Okamura, T., T. Fujimura and M. Date, 1952, Sci. Repts. Res. Inst. Tohoku Univ.
A4, 191.
2. Miles, P.A., W.B. Westphal and A. von Hippel, 1957, Rev. Mod. Phys. 29, 279.
3. Kankowski, E.F. (unpublished)-after Von Aulock, Handbook of Microwave Fer-
rite Materials (Academic Press, New York, 1965).

dependences (Rezlescu and Rezlescu 1974). Note that relaxation maxima were
observed in the temperature and frequency courses of the dielectric loss factor in
the kHz r e g i o n - s e e fig. 58 (Kamiyoshi 1951). The interpretation is not entirely
clear, however. The correlation of the relaxation maxima to the Fe z+ content and
to the relevant electron hopping is demonstrated in fig. 59 (Samochvalov et al.
1967, see also Mizushima and Iida 1967, Mizushima et al. 1978).

Z5 ~ o" '
rj Q) o tab

~o

'

I
- 50 0 50 100
r (°C)
Fig. 58. tan 6 vs temperature at various frequencies f o r a slowly cooled CoFe204 (Kamiyoshi ]95]).

4.1.3. Optical and magnetooptical spectra


The difference between the dielectric constant measured at microwaves and the
one determined from the refractive index in the near infrared is typically ~ 3 to
10. It primarily corresponds to dispersion due to the phonon excitations w h o s e
O X I D E SPINELS 277

40

E', E'"
30
15
I
ooo°°°
C' oo°
oO
/,° /

30 U 2.o
oT .°°e/ o o

o/
¢. 2 6 I oo°1
o °
50 I00 c
2.'
olo
o~'~
T(K) o/
d

2o /o ~o
o-'1
o' /o/~ "re/ /

--
Z,/' ..oy,o
2" \o
I
10 ~
o.o~
__ol i..~o- io--
c,- 13
"o,.,
~ o~
-o-
_"o°-- o- °~~ °o'-°o ~C~c - / c,*
/o / o o x
/ o 3"o

o_o=OL.oJ
0 ~°o9.~8°--#-°6"~cr°, - , , -
200 ZOO 600
r(K)
Fig. 59. Temperature dependence of the complex dielectric constant of magnetite (in the insert) and
ferrites (Ni0.3Zn0.7)>xFe{+Fe204; (1) x = 0.085; (2) x = 0.049; (3) x = 0.01 (Samochvalov et al. 1967).

frequencies are lying in the far infrared. As we are dealing here with electronic
excitations we postpone the discussion of the phonon spectra to the next
paragraph and discuss first the near infrared and visible regions.
The absorptions due to optical phonons are situated mostly below 0.1 eV. At
photon energies above this limit a region of relative transparency occurs in most
oxide spinels that are electrical insulators. This optical "window" is particularly
broad and clear if the crystal contains no paramagnetic cations. The presence of
transition metal ions (3d"), especially in magnetically ordered state, makes pos-
sible various electronic transitions that involve the crystal field split levels of these
ions. These transitions modify the absorption edge, usually defined by the onset of
interband transitions, by extending it towards the lower energies and adding
several more or less distinct peaks. The situation is similar to that one in garnets,
particularly Y3FesO12 (YIG), that is perhaps the most thoroughly investigated
ferrimagnetic oxide as far as the optical properties are concerned (see e.g. Scott
1978). The possible types of transitions in YIG are schematically indicated in fig.
60. The Y3FesO12 garnet contains two Fe 3+ sublattices with octahedral and
tetrahedral coordinations, respectively, and with mutually antiparallel mag-
netizations. Hence the inverse spinel ferrites not containing other transition metal
ions except Fe 3÷ are expected to exhibit optical properties analogous to YIG. The
examples are Li0.sFe2.sO4 and MgFe204; their absorption spectra in the near
infrared are given in fig. 61 together with NiFe204. It seems that the presence of
Ni 2+ brings no qualitative changes in this spectral region. Between 1 and 2 eV
there do not exist enough reliable data due to the poor transparency of the
278 S. KRUPICKA AND P. NOVekK

,p ~p (FO

4s (Fe)

aa (Fe 2,)

"t 4 r, ~

2p (0 2-)
tetrahearat oclaheolral

Fig. 60. Schematic density-of-states plot for Y3FesO12 showing examples of various types of the optical
transitions: (A) crystal field, (B) intersublattice Fe 3+ pair transitions, (C) 2p-+ 3d charge transfer, (D)
2p ~ 4s charge transfer, (E) 3d ~ 4d charge transfer; after Scott (1978).

samples. But the similarity to YIG may again be followed above 2 eV in fig. 62
where the diagonal elements of the dielectric function are plotted for both Li
ferrite and YIG.

Region 0.1 to i e V
The "window" between ~0.15 and 0.4 eV in fig. 61 is clearly indicated though it is
not so deep and broad as in YIG. This may be at least partly due to imperfections
(chemical non-stoichiometry, impurity ions, crystallographic defects). In parti-
cular, the optical absorption seems to be strongly influenced by the presence of
octahedral Fe 2+ forming electronic charge carriers (see section 4.1.1). A good
piece of evidence has been obtained by Sim~a et al. (1979, 1980) who studied the
optical constants of MnxFe3_xO4 single crystals by reflectance and ellipsometric
methods down to 0.5 eV and extended the spectra further to lower energies by
Kramers-Kronig analysis. Figure 63 shows the spectral behaviour of the real and
imaginary parts of the dielectric function. The peak in the vicinity of 0.5 eV which
is very strong in magnetite (x = 0) diminishes with decreasing Fe 2÷ concentration
(increasing x) and finally disappears for x ~> 1.1 where practically no Fe z+ is
expected. The comparison with theoretical predictions points to the small polaron
model which is adequate for describing this effect (Sim~a 1979). Similar low
energy peaks were observed in magnetite also by other a u t h o r s - B u c h e n a u and
Mfiller (1972) and Schlegel et al. (1979); the interpretation in the latter case was
different, however, being based on the analogy with photoemission spectra.
OXIDE SPINELS 279

200 !

(cmt) Log oc

100 2.0 \~. A ii°sF°25


80
60 .__MgFe204
40

20

10
8
6
I
I

I
I

o.'12 o.~s 05 05 o'.~ o.'5 O.L6 0.1?0!8


E (w)
Fig. 61. Near infrared absorption spectra of several spinel ferrites as compared with YIG (after
Zanmarchi and Bongers 1969).

E,;' , ",-,, co' ~o

1 / ,/ .... YSo, o,~ -'--. /~; 2


~Fe2.504 " ' ' ' Q

i i i i
zo a'o 42o 51o 8.o
~o(eV)
Fig. 62. Comparison of diagonal elements of the dielectric functions of Li ferrite and YIG (after
Vigfiovsk!) et al. 1979b).
280 S. KRUPI(2KA AND P. NOVAK

14
12 l
I
I
|
t
I
10
I

8
6
4
2

__~~/~_
I i I i I I ~ I I "rT i i i i [ i i i i
t i

7 ", / ~ . ~ Mno,~ Fe2,q Hn,,Fe.~G


E;
6
5
4
\
3 %

E(eV) E(eV)
i i - - i i

14, -r,
12-; ", Fe~O, Mno.5 Fe 2 5 O~

10
8

i i i i r J i i i i i i i i i i i i

i i i i

i 2 3 1 2 3 4-
E (eV) E (~V)
Fig. 63. Real and imaginary parts of the dielectric function vs energy for Mn=Fe3-xO4spinels (Sim~a et
al. 1979).
O X I D E SPINELS 281

Region 1 to 2 eV
The transitions observed in this spectral range are usually identified as transitions
between 3d levels split by the crystal field. For example, the steep increase of the
absorption in NiFe204 in the vicinity of l e V (fig. 61) was ascribed to the Ni 2+
crystal field transition 3A2~3T2. This peak can be observed in the spectrum of
NiO ()~ = 1.15 Ixm) and manifests itself also in a Faraday rotation dispersion of
NiFe204 (Zanmarchi and Bongers 1969); its energy should be equal to the
crystal-field splitting parameter zl = lODq. This transition, like all crystal field
transitions in octahedrally coordinated ions, is parity-forbidden but may be
allowed due to statical or dynamical violation of the local inversion symmetry.
The parity restriction is not valid for tetrahedral cations (symmetry Td).
On the other hand, the crystal field transitions in many cations including the
important case of Fe 3+ are also spin-forbidden. The fact that in spite of the
selection rules many of these transitions appear in the spectra can be understood
when allowing a simultaneous excitation of a magnon so that AStot = 0 remains
valid (see, e.g., Scott 1978). Another model uses the admixture of higher states via
spin--orbit coupling (Clogston 1959); the spin is no longer a good quantum number
and the spin constraint is automatically removed. But in most cases this fails to
account for the observed oscillator strengths and some other effects (Scott et al.
1974, Dillon 1971).
For assignment of various crystal field transitions the diagrams of crystal field
levels for transition metal ions are being used (Sugano et al. 1970). The possible
Fe 3+ crystal field transitions in YIG that should also occur in inverse spinel ferrites
are listed in table 31. Note that the oscillator strengths are larger for tetrahedral
coordination where the parity constraint is absent.

T A B L E 31
Crystal field transitions in YIG at 77 K (after Scott 1978).

Transition Oscillator
energy (eV) Assignment Site strength

1.372 6Alg(65) --~4Tlg(4G) B 2× 10-5


1.804 6Alg(6S) ---->4T2g(4G) B 2× 10-5
2.000 ~ 2.100 6A1(6S) ~ *ra(4G) A 8x 10-5
2.445 6A 1(6S) ~ *r2(4G) A 1.6 x 10-4
2.568 6A1(65)---)4E, 4Al(4G ) m 3.2 × 10-s
2.652 6Alg(68) ~ 4Eg, 4Alg(4G) B 2x 10-s
2.792 6Alg(6S)-~ 4T2g(4D) B 1x 10 -4
2.994 6A1(65) --~4T2(4D) A 6x 10-s

Region above 2 e V
The spectra in this region are ascribed to charge transfer transitions, the most
probable mechanism being the intersublattice transfer (Blazey 1974, Scott et al.
1975, Wittkoek et al. 1975), e.g.,
282 S. KRUPICKA AND P. NOVSd~

(Fe 3+) + [Fe 3+] + hu -+ (Fe 4+) 5~ [Fe2+]

(see fig. 60). Other possibilities are biexciton transitions, i.e., simultaneous crystal
field transitions in both sublattices with AStor= 0 (Blazey 1974, Krinchik et al.
1977) or charge transfer between 02- and Fe 3+ (fig. 60). But the oscillator
strengths depending on Fe 3+ concentration and the fact that the diamagnetic
dilution in any sublattice influences all absorptions (Krinchik et al. 1979) indicate
the first model involving [Fe3+]-(Fe 3+) pairs is correct. At still higher energies
(above 3-4 eV) also orbital promotion ( 3 d ~ 4s) or interband (2p ~ 4s) transitions
become important.

Magnetooptical effects
Unfortunately there are only scarce experimental data on spinels which can be
directly compared with the above conclusions. One of them is the Li ferrite,
studied by Malakhovsky et al. (1974) and Vigfiovsk~ et al. (1979, 1980a, b). The
other ones are Mn-Fe spinels including magnetite, systematically investigated by
Simga et al. (1979, 1980), and Ni and Co ferrites studied, e.g., by Kahn et al.
(1969), Westwood and Sadler (1971), Krinchik et al. (1977, 1979) and Khrebtov et
al. (1978). On the other hand, the assignment of various transitions can be often
supported and completed on the basis of the magnetooptical data. They include
the Faraday rotation - mainly in the near infrared (up to 1 eV) and both polar and
transversal Kerr rotations above l e V . Actually, many of the references given
above are partly or fully devoted to the magnetooptical studies. Sometimes also
the reflectance circular dichroism has been examined from which both Faraday
and Kerr rotations may be calculated (see, e.g., Ahrenkiel et al. 1974a, 1975). All
magnetooptical effects are intimately related to the off-diagonal elements of the
dielectric tensor function; both diagonal and off-diagonal elements are complex so
that besides the index of refraction and absorption coefficient two additional
independent measurements are necessary for determining the whole dielectric
function (e.g. Kerr rotation and Kerr elipticity).
Examples of the magnetooptical spectra of Li0.sFe2.504 and some other simple
ferrites as compared with those of YIG are given in figs. 64(a) and (b). The
spectral dependences are similar but the sign of rotation in spinel ferrites opposes
that in YIG due to the opposite mutual orientation of sublattice and total
magnetizations in both types of materials. As the magnetooptical effects depend
on both diagonal and off-diagonal elements of the dielectric tensor function their
strengths (magnetooptical activity) generally does not simply correspond to the
oscillator strengths of the underlying optical transition. In order to obtain large
magnetooptic effects the transition has also to be highly circularly polarized. An
example of strong magnetooptical activity are spinels containing Co 2+ in A
positions (Ahrenkiel et al. 1974b). This has been ascribed to the crystal field
transitions 4A2(F)--+ 4T2(F ) and 4A2(F)--+ 4TI(P ) of CO2+(A). In fig. 65 the spectral
dependence of the reflectance circular dichroism for some of such spinels
measured on polycrystalline layers are shown.
OXIDE SPINELS 283

200

Io/° .o.o.t_e.o-m-- o----re.


,,:,~.~" ~gFe, o~

E
,%0
-200 ,,'/
iJ
"o/
-400
c_ / T = 300°K
• o experimenf.a~
-- calculated
~- - 6 0 0

-800
,i
a
- 100~ ~ ,~ ~- ~ '
~,avel.eng~.h X (~m)

i i i r.,~, b i i i i

O.04 i \~
•" i .~ / . , . / - E; L F
; ',,. i "~.J ~.-- :,, y/:.
' "',,o~..~,,'-. ~:/~ ~' y".

, :d ! ~ / i . . . . ..

....":":: -I' : ,. ......


......
'~ r. " " f" •

•-~ ~; ',.0....'!, \ :, : ,,'I ',, ,.


"£ ~'J "../'"..... \ "~" ,' / ' , ~'x ."
-~;z_~_J;( ;.~- \ :'.-.Z.g '- .>q
C; Y / G ~ - ..~ ,' / .... -

-0.04 ' " ' '


iI Ie

b "/
~- ,. ~
2 4
phol;on energy (eV)

Fig. 64. Comparison of magnetooptical behaviour of Fe spinels and YIG. (a) Faraday rotation in near
infrared (Zanmarchi and Bonders 1969). (b) Spectral dependence of the complex off-diagonal element
for YIG and Li ferrite (Vi~fiovsk~) et al. 1979b).
284 S. KRUPI(~KA AND P. NOV/~d,~

~ ~ T - - T - - T - - r - - 20

4 ~ Rco(o/o)
ReD (%) _ _ Co Rh, 5 Feo 50,; 12
j ' .
2

,, ,,
0 0

l -4

-e ,,,, \
ll~l - 12

~ ~ ~ ~ -20
0.6 0.8 [0 1.2 1.4 ¢.6 1.8

Fig. 65. Reflectance circular dichroism of CoRhl.sFe0.sO4 and CoCr204 at 80 K (Ahrenkie] and Coburn
1975).

4.2. Mechanical and thermal properties

4.2.1. Infrared spectra


The first important paper on the infrared spectra of spinels was published by
Waldron (1955). To explain the experimental data obtained for seven ferrite
spinels Waldron refers to a rhombohedral unit cell containing only 14 ions (for a
normal spinel MFe204 it consists of two MO4 tetrahedra and one Fen tetrahedron).
Four modes were found to be infrared active. Two of them, having the higher
frequency, were supposed to arise from the motion of the oxygen ions, the
remaining two were assigned to the motions of the cations only. Group theory,
taking the full cubic crystallographic unit cell into account, was applied to the
vibrational problem by White and De Angelis (1967) and by Lutz (1969).
Waldron's conclusion that only four modes are infrared active in an ideal normal
spinel was proved to be right, the origin of modes was, however, found to be
more complex. Very thorough experimental investigation of infrared spectra of
normal spinels, performed by Preudhomme and Tarte (1971a, b, 1972) confirmed
the complexity of the problem. Typical infrared patterns for normal 2-3 spinels
are shown in fig. 66, the observed frequencies are summarized in table 32. It is to
be mentioned that Grimes (1972b) suggested an entirely different explanation of
infrared spectra of spinels based on the two-phonon processes.
The calculation of the force constants of thiospinels was performed by Brtiesch
and D'Ambrogio (1972), Lutz and Haueuseler (1975) and Lauwers and Herman
(1980). The last authors made corresponding analysis also for MgA1204 and
ZnGa204 spinels.
When the symmetry of the spinel structure is lower than cubic or/and sup-
plementary ordering of cations exists, more infrared bands may appear (White
and De Angelis 1967). A splitting of one infrared band is observed in some spinels
containing octahedraUy coordinated Jahn-Teller ions Mn 3+ and Cu 2+ (table 33). It is
to be noted that the bands may be split also due to the presence of two different kinds
OXIDE SP1NELS 285

i i i i i

80

60

o~ 40

20

4° _ i i i ~ i __

#_
8O

60

40

20

80

60

40
Zn Fe2 04
20

800 ' 600 4bo Cm-1


' 2bo

Fig. 66. Typical infrared patterns for three normal 2-3 spinels (Preudhomme and Tarte 1971b).

of ions in the same sublattice. Such a splitting exists, e.g., in the system ZnAlxCr2-xO4
( P r e u d h o m m e and Tarte 1971b).

4 . 2 . 2 . JUlastic c o n s t a n t s
There are only few spinels for which all the elastic constants are known.
Corresponding values at room temperature together with the anisotropy factor
A = 2 c 4 4 / ( c l l - cl2) are summarized in table 34 (for an isotropic material A = 1). It
is seen that NiCr204 and to some extent TiFe204 differ markedly from other spinel
systems as far as the elastic constants are concerned. This anomaly is connected
with the cooperative Jahn-Teller effect (section 1 . 6 ) - f o r NiCr204 the cor-
responding critical t e m p e r a t u r e TjT is close to room temperature. In the vicinity of
Trr the elastic response of the system is sensitive to the t e m p e r a t u r e - t h e crystal
softens (at least to some extent) as TIT is approached. For the system
NixZnl_xCr204 this softening is demonstrated in fig. 67. A similar situation exists
also in TiFezO4 (see section 3.1.3) and FeCrzO4, the corresponding temperatures
TjT are low, h o w e v e r , compared to NiCr204. For spinels which do not contain
Jahn-Teller ~ions the elastic constants depend only weakly on the t e m p e r a t u r e
(e.g. Kapitonov and Smokotin 1976).
286 S. KRUPIOKA AND P. NOV~6d~2

TABLE 32
Infrared absorption bands of several spinels.

Absorption bands (cm-1)


System /21 v2 /23 iv4 Ref.

Fe304 570 390 268 178 1


NiFe204 593 404 330 196 1
CoFe204 575 374 320 181 2
MnFe204 545 390 335 3
ZnFe204 552 425 336 166 4
CdFe204 548 412 319 4
CoCr204 630 530 402 197 4
MgAI204 688 522 580 309 4
Fe2GeO4 688 402 319 178 5
NizGeO4 690 453 335 199 5

1. Grimes, N.W. and A.J. Collet, 1971, Nature


(Phys. Sci.) 230, 158.
2. Waldron, R.D., 1955, Phys. Rev. 99, 1727.
Mitsuishi, A. et al., 1958, J. Phys. Soc. Jap. 13,
1236.
3. Brabers, V.A.M. and J. Klerk, 1974, Solid State
Commun. 14, 613.
4. Preudhomme, J. and P. Tarte, 1971, Spectrochim.
Acta 27A, 1817.
5. Preudhomme, J. and P. Tarte, 1972, Spectrochim.
Acta 28A, 69.

TABLE 33
Splitting of the /24 band in three spinels exhibiting the cooperative
Jahn-Teller effect (after Siratori 1967).

Compound /2~ , /22 t /24 $ c/a

ZnMn204 265 cm-1 167 cm -1 232 cm -1 1.14


Mn304 247 165 220 1.16
CuCr204 135 194 155 0.91

* Stronger line of the split v4.


+ Weaker line of the split /24'
$ Weighted mean of the two lines.

I n t a b l e 35 t h e d a t a o n t h e e l a s t i c m o d u l i a n d t h e c o m p r e s s i b i l i t y m e a s u r e d o n
polycrystalline ferrites at r o o m t e m p e r a t u r e are given.
W e n o t e t h a t in a n a l o g y t o t h e m a g n e t i c r e l a x a t i o n , a n e l a s t i c r e l a x a t i o n w a s
o b s e r v e d in s o m e s p i n e l f e r r i t e s ( G i b b o n s 1957, I i d a 1967) c o n n e c t e d w i t h
d i f f u s i o n a n d r e a r r a n g e m e n t p r o c e s s e s in t h e l a t t i c e .
OXIDE SP1NELS 287

TABLE 34
Elastic constants of several spinels (in units of 10-11 dyn/cm) at
room temperature. A is the anisotropy factor (see text).

System cx c~2 c44 A Ref.

MgA1204 27.9 15.3 15.3 2.43 1


Fe304 27.5 10.4 9.55 1.12 2
NiFe204 21.99 10.94 8.12 1.47 3
Lio.sFez504 24.07 13.41 9.29 1.74 4
ZnCr204 25.57 14.23 8.46 1.49 5
FeCr204 32.2 14.4 11.7 1.31 6
NiCr204 17.5 17.1 5.86 24.1 5
TiFe204 cu cm = 2.65 3.96 2.99 7

1. Lewis, M.F., 1966, J. Acoust. Soc. Am. 40, 728.


2. Doraiswami, M.S., 1947, Proc. Indian Acad. Sci. A25, 413.
3. Gibbons, D.F., 1957, J. Appl. Phys. 28, 810.
4. Kapitonov, A . M . and E.M. Smokotin, 1976, Phys. Status
Solidi a34, K47.
5. Kino, Y. et al., 1972, J. Phys. Soc. Jap. 33, 687.
6. Hearmon, R.F.S., 1956, Adv. Phys. 5, 323.
7. Ishikawa, Y. and Y. Syono, 1971, J. Phys. Soc. Jap. 31, 461.

vt3 f x=O
(krns")

2 o.~"
? x : 0.37
TN
oo• °

J
f
• e •
x = 0.73 ."

F
r.
(X)
Fig. 67. Soft mode sound velocity Vt=[(Cll--Cl2)/2p] lj2 as a function of temperature in a
NixZnl-xCr204 system. Ta, TN are the critical temperatures for cooperative Jahn-Teller transition and
the N6el transition respectively (Kino et al. 1972).

4.2.3. H e a t capacity
In f e r r i m a g n e t i c s p i n e l s t h e specific h e a t at l o w t e m p e r a t u r e s ( T < Tc, OD) is
d o m i n a t e d by t h e m a g n e t i c c o n t r i b u t i o n . T h e t e m p e r a t u r e d e p e n d e n c e of Cp is
t h e n w e l l d e s c r i b e d by t h e spin w a v e t h e o r y , w h i c h p r e d i c t s Cp ~ T a/2 ( s e c t i o n
2.4). A t h i g h e r t e m p e r a t u r e s t h e l a t t i c e c o n t r i b u t i o n ( p r o p o r t i o n a l t o T s f o r
288 S. KRUPI(~KA A N D P. NOV_,~d(

T A B L E 35
Elastic moduli of several polycrystalline ferrites; after Seshagiri Rao et al.
(1971).

X-ray Elastic moduli


density (1011 dyn/cm 2) /3 × 1013 Poisson
System (g/cm 3) E n k (cm2/dyn) ratio

MgFe204 4.52 19.73 7.34 21.17 4.72 0.34


CoFe204 5.29 17.34 6.54 16.62 ,6.02 0.33
NiFe204 5.38 15.59 5.89, 14.69 6.81 0.32
ZnFe204 5.33 18.64 7.27 14.27 7.01 0.28

T < OD) prevails. An example of the temperature dependence of Cp is shown in


fig. 68. In table 36 the values of C. at room temperature are summarized together
with the values of Debye temperatures OD deduced from the low temperature
measurements. Venero and Westrum (1975) noted that at elevated temperatures
the lattice part of Cp for spinels may be well described by Kopp's rule based on
the component oxides. For example for normal 2-3 spinels:

Cp(AB204) = Cp(AO) + Cp(B203).

T(K)
100 200 300 400 500 600

40 mm.m.m"m
"ww~m" mrm m m'm''m'm'am~

.-. i....-- ............


-~ 30 ,-" _~°~ / 03

"-'20 • o"° I /" dO, "~

,o //
: / /
./ o/°/-Io,I
t ~o k

0 L-'-'~° ~ , ,~ ~ , , , I
0 10 20 3'0 40
T(K)
Fig. 68. Heat capacity vs temperature for Li0.sFe2.504(O) and Lio.sAlzsO4 (O) (Venero and Westrum
1975).

4.2.4. Thermal conductivity


Thermal conductivity is a composite e f f e c t - i n magnetic spinels besides phonons,
both electrons and magnons may participate in the transfer of heat. It was shown,
OXIDE SPINELS 289

TABLE 36
Thermal properties of several spinels.

Cp K ff x 10 6
System (cal/K moo Oo (K) (cal/s cm K) (K-1) Ref.

Fe304 36.18 660 0.015 1, 5, 6


NiFe204 34.81 625(7)* 0.009 8 3, 4, 7,
12
CoFe204 36.53 584 (10)* 0.015 3, 4, 9
MnFe204 0.017 12 8, 12
CuFe204 0.015 9
ZnFe204 33.22 0.015 7.5 1, 6, 11
MgFe204 762 (15)* 4
Li0.sFe2.504 34.43 512(5) 0.015 12 2, 4, 9,
14
MgAI204 27.79 0.036 4.61t 3, 10, 13
5.905

* Nonstoichiometric samples.
~ Natural spinel.
Synthetic spinel.

1. Bartel, J.J. and E.F. Westrum, Jr., 1975, J. Chem. Thermodyn. 7, 706.
2. Venero, A.F. and E.F. Westrum, Jr., 1975, J. Chem. Thermodyn. 7, 693.
3. King, E.G., 1956, J. Phys. Chem. 60, 410; 1955, Ibid 59, 218.
4. Kouvel, J.S., 1956, Phys. Rev. 102, 1489.
5. Polack, S.R. and K.R. Atkins, 1962, Phys. Rev. 125, 1248.
6. Smit, J. and H.P. Wijn, 1959, in: Ferrites (Wiley, New York) p. 225.
7. Kamilov, I.K., 1963, Sov. Phys-Solid State 4, 1693.
8. Suemune, Y., 1966, Jap. J. Appl. Phys. 5, 455.
9. Smit, J. and H.P. Wijn, 1954, Physical Properties of Ferrites, in: Advances in
Electronic and Electron Physics (Acad. Press, New York) vol. 6, 83.
10. Slack, G.A., 1964, Phys. Rev. 134A, 1268.
11. Weil, L., 1950, Compt. Rend. 231, 122.
12. Bekker, Y.M., 1967, Izv. Akad. Nauk SSSR Neorg. Mater. 3, 196.
13. Singh, H.P. et al., 1975, Acta Crystallogr. A3I, 820.
14. Brunel, M. and F. de Bergevin, 1964, Compt. Rend. 258, 5628.

however, that in spinels the role of m a g n o n s is n e g a t i v e ( D o u t h e t a n d F r i e d b e r g


1 9 6 1 ) - they cause a scattering of p h o n o n s thus r e d u c i n g the t h e r m a l conductivity.
T h e c o n t r i b u t i o n of electrons to the heat transfer is b e l i e v e d to be small in spinels
(Slack 1962). T h e c o n d u c t i v i t y is very sensitive to the i m p u r i t i e s particularly to the
t r a n s i t i o n m e t a l ions with an orbitally d e g e n e r a t e g r o u n d level (fig. 69). I n table 36
the t h e r m a l c o n d u c t i v i t y at r o o m t e m p e r a t u r e for several spinels is given.

4.2.5. T h e r m a l expansion
F o r most spinels the t e m p e r a t u r e d e p e n d e n c e of the lattice p a r a m e t e r a m a y b e
well a p p r o x i m a t e d by

a = ao+ b o T + b~T 2.
290
S. KRUPIOKA AND p. NOVzid<

"d
,'I / J ,o'°""o,. \
°\ \
o ~o.2°o '~

/ • \

o
-6
E / 24×I0-

'.0

o 30 ~oo 300 ¢ooo


T(K)
Fig. 69. Temperature dependence of the thermal conductivity for Mg~_xFe~A/204system (Slack 1964).
Substitution of Fe 2+ ions into the A sublattice (orbital/y degenerate 2E ground state with a weak
Jahn-Teller coupling) leads to pronounced reduction of the conductivity.

20
~'103 x =-

15 0.9 1.0
0.8
0.6 0.Tj,.Y~~

400 500

T (K)
Fig. 70. Linear expansion coefficient ee vs temperature for Ni, Znl-xFe204 ferrites (after Henriet-
Iserentant and Robbrecht 1972). The extrema of a appear at To
OXIDE SP1NELS 291

However, the onset of ithe magnetic order causes an anomaly in the temperature
dependence of a connected with the volume magnetostriction-this is demon-
strated for the example of the\NixZnl_xFe204 system in fig. 70. Values of the
linear expansion coefficient a

1 Oa
Og --
a OT- bo/a + (2bJa)T+...

are given in table 36.

Appendix: Intrinsic magnetic properties

In this part some additional data mainly concerning the. intrinsic magnetic
properties of the most important ferrimagnetic spinels are given. For more
complete or detailed information the reader should consult the following lit-
erature:
(a) Books and tables:
Landolt-B6rnstein tables, New Series, ed., K.H. Hellwege and A.M. Hellwege
(Springer Verlag 1970) vol. 4, part b.
J. Smit and H.P.J. Wijn: Ferrites (Wiley, New York, 1959).
S. Krupi6ka: Physik der Ferrite und der verwandten magnetischen Oxide
(Academia Praha-Vieweg, Braunschweig, 1973).
Handbook of Microwave Ferrite Materials, ed., W.H. von Aulock (Academic
Press, New York and London, 1965).
A. Oleg et al.: Tables of Magnetic Structures determined by Neutron Diffraction
(Inst. of Nuclear Techniques, Krakow, 1970).
(b) Papers:
E.W. Gorter: Philips Res. Rep. 9 (1954), 295, 321, 403 (crystal chemistry and
ferrimagnetism).

120 'i ~ ~ '

100

8o co/ ~ ~

-273-200-100 0 ZOO 200 300 400 500 600 700


T(°C)
Fig. 71, Temperature dependence of magnetization of several ferrites (Smit and Wijn 1959), the
measurements were made by Pauthenet (1950), (1952).
292 S. KRUPI~KA AND P. NOVI~K

26L'
i o o o o'o : OOOOooo
......... ~ 4I
~- °OOoo a 1

co

~, .%b Kb o O
,~~- 81 ooo o°°

= ~,, Kbb
2~oo8888o'~ ................
OI eKCIG °°°o °O Ooo

- Ku°°o
f °°°°
20 40 60 80 I00 120
T(K)
Fig. 72. Temperature dependence of the anisotropy constants of magnetite below the Verwey
transitions (Chikazumi 1975). The free energy is expressed as: F = Fo + Kaoe~ + Kba 2 --
guog211 q- gaaOd4 + KbbOL4 + KabOg2aOL2 where aa, oq,, oem are the direction cosines of the magnetization with
respect to the monoclinic a, b axes and [111] respectively.

15 x~x
.KI .iO 4

./;./ °-,~,.k."~-~,a04
4 _1~i !
I°!2,,s -155~' :I."
-185 °C I
F%04 0/25 0,'50 0/75 NiOFe203
Fig. 73. NixFe3-xO4 system. First anisotropy constant K1 plotted for several temperatures as a function
of Ni-content x (Elbinger 1962).
OXIDE SPINELS 293

n i n -

~o

%
.~ -40 O. ~ o ~ o ~ o-- o - - - o ~ o...___.. °
O.z~ ° " - " e ~ o--o ,N--O O~ J"'~ 0 ~ 8

0 o ~ £ O ~ O'O-- O - - O - - O ~ O ~ O_O
-2o
~ O~ I ' ~ O ...........~ O ......~ O ~ 0

E
0

3bo
T(K)
i i

0
80 /./,.-"0"-"" " "
x

/
60 / 0.2

n~
,o /" . / ..... -::
E
,,.z%.-........ o;
0 ,0
,~'0 0.8
o--o--O--O-- .o--o , o--o--o
-20
~bo 1~o ' ebo e'5o abo
r(K)
Fig. 74. NixFe3-xO4 system. Magnetostrictions A]00 and An1 plotted against temperature for several
Ni-concentrations x (Brabers et al. 1980).

0.5 , , , , , , , , ,

-0.5
~ / / / i " :~~ ' ~ ' , o ~ o
~ o~ ~ .
.x

~ ~///// \\i '°°


? _,,o \llx/o~
~--.o ~,/ \~o/ o.t11"
-e.s o.'2 o'.~ o:e o:e /o I.? I.? 7.'6 ~.'o
X
Fig. 75. MnxFe3-,O4 system. First anisotropy constant Kx plotted for several temperatures as a
function of Mn-content x (Penoyer and Shafer 1959).
294 S. K R U P I C K A A N D P. N O V A K

tO0

80 °'~°~°'°°'o
120
x

E 60 /°/° ~'%~Du + 1.0 o¢~''\o


-A ~oo" I0 ~
&
100 .~,~ I.F °I
..< 40
•...a. o ~ ,
1.9
/i •

A!oo
\
oo
20 o oo_o.o~o,o
80
o,
I
0

-20
- - % - co-o- 0-°--% "o ~ o _ - 60
/:{ o tt ~
"\ \.
x

.o~ 2o ~. °. . . , ~ ' ' ' - ~- ' ~ o ' ~ . ,, e , i ,, i %- \ \


,.........,..-~ -I..,.+,~, x ~ - X . x . I '~a •

-40 ~o / x x
o O. tO e 0.85 40 ..~l ,t t +"4- X~x~ 'A'~'a
sI I ~4- ×,~.
0.40 * 0.95
-60 "~ 0 . ¥ 5 • 1.05 o.o.o-

a 20
b L i i i

-150 -lO0
02 0;4 de 018
-5'0 -50
T (°C) r/rc
Fig. 76. MnxFe3 xO4 system. Temperature dependence of magnetostrictions ,~100, Zlklll- (a) x ~< 1.05
(Miyata and Funatogawa 1962); (b) 1.9~> x i> 1 (Brabers et al. 1977); Am is small and negative
( - 5 x 10-6~<Am ~<0) for x > 1.

G. Blasse: Philips Res. Rep. (1964), Suppl. no. 3 (crystal chemistry and fer-
rimagnetism).
A. Broese van Groenou, P.F. Bongers and A.L. Stuyts: Mater. Sci. Eng. 3
(1968/69), 317 (review paper).

/4 I

\ CO,:Fe3_xO4

cb
~3
f...

2
\ % x=O.04
\÷ o
\+
~x=oofo~
"%+ +-
0%0 °~Oo
4"°~°'~"oo °~°e°°'°°
o. . . . . . . . . . i~,OCLo_o..____%_o
x=O
i

%0 200 5o0
T(K)
Fig. 77. Temperature dependence of K1 for Co-substituted magnetite (Bickford et al. 1957).
OXIDE SPINELS 295

100
K,.lO-~rg/cm3 I
50!
':N'°'o-%\ R@, \4,

0 i o ,A t , i i
50 "o\o%
~ 150 200 250 300 r/~/~
. [~

" "o. Q~ l e t ~'-~ i X = 02


-50 o..%~ %
~... ~ /o
...o.o°
~/~o.~
_ o ~ . x:o.8
b~=o.o

-I00 ~" ~ ' > ' ~ - o - o _ ~ ; . . . . . ~2~.~o~_o-/o ~= 0.2


.Q,,o-O ~ z ~ / ~ t k . ~o o..o- ~-o.1
~~a'~:-.~.~'°--~o ~" \'o~--o-°"°'...K~
~,~ I r ~ ' ~ - ¢ /(= 005
~ o ~ o *"~-a~,~ "

-150

Fig. 78. Temperature dependence of K1 for MgxFe3 xO4 system (Brabers et al. 1980).

10

4t0.1 0.15 0.2

* 4.2K
• 77K

-IO A ~ ~ ~ ~
A 300K
~-

,o -20 " ; i
×

-30
L t

Fig. 79. Dependence of magnetostrictions Am0, /~111 for MgFe204 on the degree of inversion 8 (after
Arai 1973).
296 S. K R U P I C K A A N D P. N O V / k K

• t"- q'3
r',- oo t"q t"q oq d
),(
~5
oq.
z
q3
t"q ,,.~ oo'~, O
e'~
I I

o'3 ~,~
o
~D
I I I I I I
I

er~ u'3 t",,I ,-.t


/'q eel ~-.a ee3 ~ oq
I I I eq
I I I

',,O P,- q3 P'- ~,o o0 _=


;>
X I I
e,,
.'=
E
O

? ,-... I

¢e3
x~ t"'- t"q
~rq.
~5
I "O

.,o
e¢3
e-.

02
7 I I I I I r~

×~ o.
(,.q t"q
c5 I I I I I I ] I I

t",- G', ,...-~ , ~ ¢¢3 ee) 02


tt~ t'e3 e¢3

o
'~" 0 ~ (",,I
e¢3 0'3
O

02
P-,- o o ~ ) tt3
q3
tr3 ~1~ 1"-,- 17-,- tt3 p-- oo G'~
©
E

q eq
o eq
¢)
e~ 2
.¢)
r)
OXIDE SPINELS 297

o ~ u ~

b o

0 r~
0

©
u:

x = -= d 'E .=
oo r~ © :0

0
"O ~ '.~ ,

M
©
~.,~

~~ ~ ~~
~:~
~'~ ~ ~=~

N &
eg~e ~e~g

a ~ t--- ~ ...-k

e.
~ ' - ~
a ~mN

,~ < .~
..~ . ~ ~ ~,o 0 ©
.. = .~ .~ .~

~.~ ~-~ ~.~


-- ,._;
a) ; ~ ' ~

~oo~, ~.~
~ Nco~ N ~g ,. ..o r.,4 ~
- ~ •~ .r. ~ >: •

2~
~ i ~ ~ ~~ ~ '~
298 S. KRUPI(~KA AND P. NOVAK

TABLE 38
Hyperfine magnetic fields in kOe at an A-site Fe 57 nucleus, a
B-site iron nucleus with six nearest neighbor A-site iron ions, and
the average field at the B-site iron nuclei in various ferrimagnetic
spinels (van der Woude and Savatzky 1971).

Ferrite Hhpf(A) Hhpf(B,6nn Fe) Hhpf(B)

MgFe204(sc) -500 -537 -530


MgFe204(q) - 509 - 540 - 525
MnFe204 - 512 - 550 - 520
NiFe204 - 515 - 555 - 555
CoFezO4(sc) - 511 - 545 - 541
CoFe204(q) - 511 550 - 536
Li0.sFez.sO4 -518 -545 -545
ZnFe204 - - 557 (a) - 485

(a) Obtained by adding supertransferred hyperfine field to Hhpf,


sc and q indicate slowly cooled and quenched samples respec-
tively.

Acknowledgement

The authors acknowledge valuable discussions and the help of Dr. Simga when
p r e p a r i n g t h i s c h a p t e r . G r a t i t u d e is a l s o e x p r e s s e d t o D r . Z f i v 6 t a f o r his c r i t i c a l
r e a d i n g of t h e m a n u s c r i p t .

References

Adler, D., 1968, Insulating and Metallic States Akino, T. and K. Motizuki, 1971, J. Phys. Soc.
in Transition Metal Oxides, in: Solid State Jap. 31, 691.
Physics, eds., Seitz, F., D. Turnbull and H. Alvarado, S.F., W. Eib, F. Meier, D.T. Pierce,
Ehrenreich (Academic Press, New York, K. Sattler, H.C. Siegmann and J.P. Remeika,
London) vol. 21, pp. 1-113. 1975, Phys. Rev. Lett. 34, 319.
Ahrenkiel, R.K. and T.J. Coburn, 1975, IEEE Alvarado, S.F., M. Erbudak and P. Munz, 1977,
Trans. Magn., MAG-11, 1103. Physica 86--88B + C, 1188.
Ahrenkiel, R.K., T.J. Coburn and E. Carnall, Anderson, P.W., 1956, Phys. Rev. 102, 1008.
Jr., 1974a, IEEE Trans. Magn. MAG-10, 2. Anderson, P.W., 1958, Phys. Rev. 109, 1492.
Ahrenkiel, R.K., T.J. Coburn, D. Pearlman, E. Anderson, P.W., 1959, Phys. Rev. 115, 2.
Carnall, Jr., T.W. Martin and S.L. Lyu, Anderson, P.W., 1963, Exchange in Insulators,
1974b, AIP Conf. Proc. 24, 186. in: Magnetism, eds., Rado, G.T. and H. Suhl
Ahrenkiel, R.K., S.L. Lyu and T.J. Coburn, (Academic Press, New York, London) vol. 1,
1975, J. Appl. Phys. 46, 894. pp. 25-86.
Aiyama, Y., 1966, J. Phys. Soc. Jap. 31, 1684. Anderson, P.W., 1970, Comments Solid State
Akino, T., 1974, J. Phys. Soc. Jap. 36, 84. Phys. 11, 193.
OXIDE SPINELS 299

Arai, K.I., 1.973, Rep. Res. Inst. Electr. Com- Briining, S. and H.Ch. Semrnelhack, 1979,
mun. Tohoku Univ. 25, 79. Conference on Physics of Magnetic Materials
Arai, K.I. and N. Tsuya, 1973, J. Phys. Chem. COMECON, Dresden, unpublished.
Solids 34, 431. Buchenau, M., 1975, Phys. Status Solidi b70,
Arai, K.I. and N. Tsuya, 1974, Phys. Status 181.
Solidi b66, 547. Buchenau, U. and I. Mfiller, 1972, Solid State
Arai, K.I. and N. Tsuya, 1975, J. Phys. Chem. Commun. 11, 1291.
Solids 36, 463. Buckwald, R.A., A.A. Hirsch, D. Cabib and E.
Aubert, G., 1968, J. Appl. Phys. 39, 504. CaUen, 1975, Phys. Rev. Lett. 35, 878.
Baltzer, P.K., 1962, J. Phys. Soc. Jap. 17, Suppl. Callen, E., 1968, J. Appl. Phys. 39, 519.
B-l, 192. Carr Jr., W.J., 1966, Secondary Effects in Fer-
Barth, T.F.W. and E. Posnjak, 1932, Z. Kris- romagnetism, in: Handbuch der Physik, ed.,
tallogr. 82, 325. Fliigge, S. (Springer Verlag, Heidelberg,
Baszyfiski, J. and Z. Frait, 1976, Phys. Status Berlin, New York) vol. 18/2.
Solidi b73, K85. Chikazumi, S., 1975, Technical Report ISSP
Bernstein, P. and T. Merceron, 1977, J. Phys. ser. A, no. 737.
(France) 38, C1-211. Clogston, A.M., 1955, Bell Syst. Tech. J. 34,
Bertaut, F., 1951, J. Phys. Rad. 12, 252. 739.
Bickford, L.R., J.M. Brownlow and R.F. Clogston, A.M., 1959, J. Phys. Rad. 20, 151.
Penoyer, 1957, Proc. Inst. Electr. Eng. 104B, Cullen, J.R. and E. Callen, 1970, J. Appl. Phys.
238. 41, 879.
Birss, R.R., 1964, Symmetry and Magnetism Cullen, J.R. and E. Callen, 1973, Phys, Rev. B7,
(North-Holland, Amsterdam) ch. 5, §3. 397.
Blasse, G., 1964, Philips Res. Rep. Suppl. no. 3. De Boer, F., J.H. van Santen and E.J.W. Ver-
Blasse, G., 1965, Philips Res. Rep. 20, 528. wey, 1950, J. Chem. Phys. 18, 1032.
Blasse, G. and E.W. Goner, 1962, J. Phys. Soc. Delorme, C., 1958, Bull. Soc. Fr. Mineral.
Jap. 17, Suppl. B-l, 176. Crystallogr. 81, 79. /
Blazey, K.W., 1974, J. Appl. Phys. 45, 2273. Dillon Jr., J.F., 1971, Magneto-Optical Proper-
Boshaan, A.J. and H.J. van Daal, 1970, Adv. ties of Magnetic Crystals, in: Magnetic Pro-
Phys. 19, 1. perties of Materials, ed., Smit, J. (McGraw-
Bouchard, R.J., 1967, Mater. Res. Bull. 2, 459. Hill, New York) p. 149-204.
Brabers, V.A.M., 1971, J. Phys. Chem. Solids Drabble, J.R., T.D. White and R.M. Hooper,
32, 2181. 1971, Solid State Commun. 9, 275.
Brabers, V.A.M., J. Klerk and Z. Simga, 1977, Driessens, F.C.M., 1968, Ber. Bunsenges. Phys.
Physica 86--88B + C, 1461. Chem. 72, 1123.
Brabers, V.A.M., T. Merceron, M. Porte and Dunitz, J.D. and L.E. Orgel, 1957, J. Phys.
R. Krishnan, 1980a, J. Mag. Mag. Mater. Chem. Solids 3, 20, 318.
15-18, 545. Eibschiitz, M., V. Ganiel and S. Strikman, 1966,
Brabers, V.A.M., J.C.J.M. Terhell and J.H. Phys. Rev. 151, 245.
Hendriks, 1980b, J. Mag. Mag. Mater. Elbinger, G., 1962, Z. Angew. Phys. 4, 274.
15-18, 599. Elwell, D., R. Parker and A. Sharkey, 1963, J.
Bradburn, T.E. and G.R. Rigby, 1953, Trans. Phys. Chem. Solids 24, 1325.
Br. Ceram. Soc. 52, 417. Elwell, D., B.A. Griffiths and R. Parker, 1966,
Bragiflski, A., 1965, Phys. Status Solidi 11, 603. Br. J. Appl. Phys. 17, 587.
Bragiflski, A. and T. Merceron, 1962, J. Phys. Englman, R., 1972, The Jahn-Teller Effect in
Soc. Jap. 17, 1611. Molecules and Crystals (Wiley, New York).
Broese van Groenou, A., 1967, J. Phys. Chem. Evans, B.J., 1975, AIP Conf. Proc. 24, 73.
Solids 28, 325. Faller, J.G. and C.E. Birchenall, 1970, J. Appl.
Broese van Groenou, A. and R.F. Pearson, Crystallogr. 3, 496.
1967, J. Phys. Chem. Solids 28, 1027. Fazekas, P., 1972, Solid State Commun. 10, 175.
Broese van Groenou, A., P.F. Bongers and Folen, V.J., 1960, J. Appl. Phys. 31, 166S.
.A.L. Stuyts, 1968, Mater. Sci. Eng. 3, 317. Fujii, Y., G. Shirane and Y. Yamada, 1975,
BriJesch, P. and F. D'Ambrogio, 1972, Phys. Phys. Rev. Bll, 2036.
Status Solidi 50, 513.
300 S. KRUPI~2KA AND P. NOV.~d(

Gehring, G.A. and K.A. Gehring, 1975, Rep. Henning, J.C.M. and H. van den Boom, 1977,
Progr. Phys. 38/1, 5. Physica B + C86--88, 1027.
Gerber, R., 1968, Czech. J. Phys. B18, 1204. Henriet-Iserentant, Ch. and G. Robbrecht,
Gerber, R. and G. Elbinger, 1964, Phys. Status 1972, C.R. Hebd. Seances Acad. Sci. B275,
Solidi 4, 103. 323.
Gerber, R. and G. Elbinger, 1970, J. Phys. C3, Hisatake, K. and K. Ohta, 1977, J. Phys.
1363. (France) 38, C1-219.
Gerber, R., L. Michalowsky, K. Motzke and E. Holba, P., M. Nevfiva and E. Pollert, 1975,
Steinbeiss, 1966, Phys. Status Solidi 16, 793. Mater. Res. Bull. 10, 853.
Gibart, P. and G. Suran, 1975, Internat. Coll. Holtwijk, T., W. Lems, A.G.H. Verhulst and
Mag. Thin Films, Regensburg. U. Enz, 1970, IEEE Trans. Mag. MAG-6,
Gibbons, D.F., 1957, J. Appl. Phys. 28, 810. 853.
Giesecke, W., 1959, Z. Angew. Phys. 11, 91. Hwang, L., A.H. Heuer and T.F. Mitchell,
Gilleo, M,A., 1958, Phys. Rev. 109, 777. 1973, Philos Mag. 28, 249.
Gilleo, M.A., 1960, J. Phys. Chem. Solids 13, Iida, S., 1960, J. Appl. Phys. 31,251S.
33. Iida, S., 1967, J. Phys. Soc. Jap. 22, 1233.
Glasser, M.L. and F.J. Milford, 1963, Phys. Iida, S. and T. Inoue, 1962, J. Phys. Soc. Jap.
Rev. 130, 1783. 17, Suppl. B-l, 281.
Glaz, W., H. Szydlowski and S. Pachocka, 1980, Iida, S. and H. Miwa, 1966, J. Phys. Soc. Jap.
Acta phys. polonica A58, 263. 21, 2505.
Goodenough, J.B., 1958, J. Phys. Chem. Solids Iida, S., K. Mizushima, N. Yamada and T.
6, 287. Iizuka, 1968, J. Appl. Phys. 39, 818.
Goodenough, J.B., 1963, Magnetism and Crys- Iida, S., K. Mizushima, M. Mizoguchi, J. Mada,
tal Structure in Nonmetals, in: Magnetism, S. Umemura, K. Nakao and J. Yoshida, 1976,
vol. 3, eds., Rado, G.T. and H. Suhl AIP Conf. Proc. 29, 388.
(Academic Press, New York and London), Iida, S., K. Mizushima, M. Mizoguchi, J. Mada,
pp. 1~63. S. Umemura, K. Nakao and J. Yoshida, 1977,
Goodenough, J.B., 1963, Magnetism and J. Phys. (France) 38, C1-73.
Chemical Bond (Wiley, New York). Iida, S., K. Mizushima, M. Mizoguchi, S.
Gorter, E.W., 1954, Philips Res. Rep. 9, 295. Umemura and J. Yoshida, 1978, J. Appl.
Gorter, E.W., 1955, Proc. IRE 43, 1945. Phys. 49, 1455.
Greenough, R.D. and E.W. Lee, 1970, J. Phys. Iizuka, S. and S. Iida, 1966, J. Phys. Soc. Jap.
D3, 1595. 21, 222.
Grimes, N.W., 1972a, Philos. Mag. 26, 1217. Ioffe, V.A., G.I. Khvostenko and Z.N. Zonn,
Grimes, N.W., 1972b, Spectr. Acta 28A, 2217. 1957, Soviet. Phys.-J. Tech. Phys. 27, 1985.
Grimes, N.W., 1974, Proc. Roy. Soc. A338, 223. Ishikawa, Y. and Y. Syono, 1971a, Phys. Rev.
Grimes, N.W., T.J. O'Connor and P. Thomp- Lett. 26, 1335.
son, 1978, J. Phys. C l l , L505. Ishikawa, Y. and Y. Syono, 1971b, J. Phys. Soc.
Groupe de diffusion in61astique des neutrons, Jap. 31, 461.
1971, J. Phys. (France) 32, C1-118. Ishikawa, Y., S. Sato and Y. Syono, 1971, J.
Gutowski, M., 1978, Phys. Rev. B18, 5984. Phys. Soc. Jap. 31, 452.
Hargrove, R.S. and W. Kiindig, 1970, Solid Ivanov, B.D., B.A. Kalinikos, O.A. Rybinskij
State Commun. 8, 303. and D.N. Czartorizskij, 1972, Fiz. Tverd.
Hastings, J.M. and L.M. Corliss, 1953, Rev. Tela 14, 653.
Mod. Phys. 25, 114. Ivanova, A.V., V.N. Seleznev and A.I. Drokin,
/-Iaubenreisser, W., 1961, Phys. Status Solidi 1, 1979, Zhurnal tekhn, fiziki 49, 2493.
6t9. Jacob, K.T. and J. Walderraman, 1977, J. Solid
Heeger, A.J. and T. Houston, 1964, J. Appl. State Chem. 22, 291.
Phys. 35, 836. Jacobs, I.S., 1959, J. Phys. Chem. Solids 11, 1.
Heikes, R.R. and W.B. Johnston, 1957, J. Jacobs, I.S., 1960, J. Phys. Chem. Solids 15, 54.
Chem. Phys. 26, 582. Jirfik, Z. and S. Vratislav, 1974, Czech. J. Phys.
Henning, J.C.M., 1980, Phys. Rev. B21, 4983. B24, 642.
Henning, J.C.M., J.H. den Boef and G.G.P. Jonker, G.H., 1959, J. Phys. Chem. Solids 9,
van Gorkom, 1973, Phys. Rev. B7, 1825. 165.
OXIDE SPINELS 301

Kahn, F.J., P.S. Pershan and J.P. Remeika, Krupi~ka, S., 1962, J. Phys. Soc. Jap. 17, Suppl.
1969, Phys. Rev. 186, 891. B-l, 304.
Kamiyoshi, K., 1951, Sci. Rep. Res. Inst. Krupi6ka, S. and P. Novfik, 1964, Phys. Status
Tohoku Univ. A3, 716. Solidi 4, Kl17.
Kanamori, J., 1959, J. Phys. Chem. Solids 10, Krupi~ka, S. and F. Vilfm, 1957, Czech. J. Phys.
67. 7, 723.
Kaplan, J. and C. Kittel, 1953, J. Chem. Phys. Krupi~ka, S. and F. Vilfm, 1961, Czech. J. Phys.
21, 760. Bll, 10.
Kaplan, T.A., 1958, Phys. Rev. 109, 782. Krupi6ka, S. and K. Zfiv6ta, 1968, J. Appl.
Kaplan, T.A., 1960, Phys. Rev. 116, 888. Phys. 39, 930.
Kaplan, T.A., K. Dwight, D.H. Lyons and N. Krupi6ka, S., L. Cervinka, P. Novfik and K.
Menyuk, 1961, J. Appl. Phys. 32, 13S. Zfiv6ta, 1964, Proc. Int. Conf. Magnetism,
Kato, Y. and T. Takei, 1933, J. Inst. Electr. Nottingham, p. 650.
Eng. Jap. 53, 408. Krupi~ka, S., Z. Sim~a and Z. Smetana, 1968,
Keffer, F., 1966, Spin Waves, in: Handbuch der Czech. J. Phys. B18, 1016.
Physik, vol. 18/2, ed., Fliigge, S. (Springer Krupi~ka, S., Z. Jir~k, P. Nov~k, V. Roskovec
Verlag, Berlin, Heidelberg, New York) pp. and F. Zounovfi, 1977, Physica 86-88B+ C,
1-273. 1459.
Khrebtov, A.P., A.A. Askochensky and J.M. Krupi~ka, S., Z. Jirfik, P. Novfik, F. Zounovfi
Speranskaya, 1978, Izv. Akad. Nauk SSSR and V. Roskovec, 1980, Acta Physica Slovaca
Ser. Fiz. 42, 1652. 30, 251.
Kienlin, A.v., 1957, Z. Angew. Phys. 9, 245. Kubo, T., A. Hirai and H. Abe, 1969, J. Phys.
Kino, Y., B. Lfithi and M.E. Mullen, 1972, J. Soc. Jap. 26, 1094.
Phys. Soc. Jap. 33, 687. Kuipers, A.J.M., 1978, Thesis, Technical Uni-
Klerk, J., V.A.M. Brabers and A.J.M. Kuipers, versity Eindhoven.
1977, J. Phys. (France) 38, C1-187. Kuipers, AJ.M. and V.A.M. Brabers, 1976,
Klinger, M.I., 1975, J. Phys. C8, 3595. Phys. Rev. B14, 1401.
Klinger, M.I. and A.A. Samochvalov, 1977, Lauwers, H.A. and M.A. Herman, 1980, J.
Phys. Status Solidi, h79, 9. Phys. Chem. Solids 41, 223.
Knowles, J.E., 1964, Proc. Int. Conf. Mag- Liolioussis, K.T. and A.J. Pointon, 1977, J.
netism, Nottingham, p. 619. Phys. (France) 38, C1-191.
Knowles, J.E., 1974, Philips Res. Rep. 29, 93. Lord, H. and R. Parker, 1960, Nature 188, 929.
Knowles, J.E. and J. Rankin, 1971, J. Phys. Lotgering, F.K., 1964, J. Phys. Chem. Solids 25,
(France) 32, C1-845. 95.
K6hler, D., 1959, Z. Angew. Phys. 11, 103. Lu-San Pan and B.J. Evans, 1978, J. Appl.
K6nig, U., Y. Gross and G. Chol, 1969, Phys. Phys. 49, 1458.
Status Solidi 33, 811. Lutz, H.D., 1969, Z. Naturforsch. 24a, 1417.
Koops, C.G., 1951, Phys. Rev. 83, 121. Lutz, H.D. and H. Haeuseler, 1975, Bet. Bun-
Kopp, W., 1919, Dissertation (Zurich). senges. Phys. Chem. 79, 604.
Kouvel, J.S., 1956, Phys. Rev. 102, 1489. Malakhovsky, A.V., I.S. Edelman, V.P.
Kowalewski, L., 1962, Acta Phys. Pol. 21, 121. Gavrilin, G.I. Barinov, 1974, Sov. Phys. Solid
Kozhuhar, A.Y., G.A. Tsinadze, V.A. Shapo- State, 16, 266.
valov and V.N. Seleznev, 1973, Phys. Lett. Marais, A. and T. Merceron, 1959, C.R. Hebd.
42A, 377. Seanees Acad. Sci. 248, 2976.
Krause, J., 1969, J. Angew. Phys. 27, 251. Marais, A. and T. Merceron, 1965, C.R. Hebd.
Kriessman, C.J. and S.E. Harrison, 1956, Phys. Seances Acad. Sci. 261, 2188.
Rev. 103, 857. Marais, A. and T. Merceron, 1967, Phys. Status
Krinchik, G.S., A.P. Khrebtov, A.A. Asko- Solidi, 24, 635.
chensky, E.M. Speranskaya and S.A. Bely- Marais, A. and T. Merceron, 1974, Phys. Status
aev, 1977, Zh. Eksp. Teor. Fiz. 72, 699. Solidi, a22, K209.
Krinchik, G.S., K.M. Muchimov, Sh.M. Marais, A., T. Merceron and M. Porte, 1969,
Sharipov, A.P. Khrebtov and M. Speran- C.R. Hebd. Seances Acad. Sei. B268, 730.
skaya, 1979, Zh. Eksp. Teor. Fiz. 76, 2126. Marais, A., T. Merceron, G. Maxim and M.
Krupi6ka, S., 1960, Czech. J. Phys. B10, 782. Porte, 1972, Phys. Status Solidi, 52, 631.
302 S. KRUPICKA AND P. NOV~iJ(

Maxim, G., 1969, Phys. Status Solidi, 35, 211. Palmer, W., 1960, Phys. Rev. 120, 342.
McClure, D.S., 1957, J. Phys. Chem. Solids, 3, Parker, R., 1975, Electrical Transport Proper-
311. ties, in: Magnetic Oxides, ed., Craik, D.J.
McMurdie, H.P., B.M. Sullivan and F.A. (Wiley, New York) pp. 421-482.
Mauer, 1950, J. Res. Nat. Bur. Stand. 45, 35. Parker, R. and M.S. Smith, 1961, J. Phys.
Merceron, T., 1965, Ann. Phys. 10, 121. Chem. Solids, 21, 76.
Michalk, C., 1968, Phys. Status Solidi, 27, K51. Parker, R. and C.J. Tinsley, 1976, Phys. Status
Michalowsky, L., 1965, Phys. Status Solidi, 8, Solidi, a33, 189.
543. Pauthenet, R., 1950, C.R. Hebd. Seances Acad.
Miles, P.A., W.B. Westphal and A. yon Hippel, Sci. 230, 1842.
1957, Rev. Mod. Phys. 29, 279. Pauthenet, R., 1952, Ann. Phys. (France) 7,
Miller, A., 1959, J. Appl. Phys. 30, 24S. 7i0.
Miyahara, Y., 1972, J. Phys. Soc. Jap. 32, 629. Pauthenet, R. and L. Bochirol, 1951, J. Phys.
Miyata, N., 1961, J. Phys. Soc. Jap. 16, 206. Rad. 12, 249.
Miyata, N. and Z. Funatogawa, 1962, J. Phys. Patton, C.E., 1975, Microwave Resonance and
Soc. Jap. 17, Suppl. B-l, 279. Relaxation, in: Magnetic Oxides, ed. Craik,
Mizoguchi, M., 1978a, J. Phys. Soc. Jap. 44, D.J. (Wiley, New York) pp. 575-649.
150~1. Penoyer, R.F. and L.R. Bickford, Jr., 1957,
Mizoguchi, M., 1978b, J. Phys. Soc. Jap. 44, Phys. Rev. 108, 271.
1512. Penoyer, R.F. and M.W. Shafer, 1959, J. Appl.
Mizushima, M., 1963, J. Phys. Soc. Jap. 18, Phys. 30, 315S.
1441. Perthel, R., 1962, J. Phys. Soc. Jap. 17, Suppl.
Mizushima, K. and S. Iida, 1967, J. Phys. Soc. B-I, 288.
Jap. 22, 1300. Philips, B., S. Somiya and A. Muan, 1961, J.
Mizushima, K., K. Nakao, S. Tanaka and S. Am. Ceram, Soc. 16, 167.
Iida, 1978, J. Phys. Soc. Jap. 44, 1831. Piekoszewski, J., J. Suwalski and L. Dabrowski,
Mott, N.F., 1969, Philos. Mag. 19, 835. 1977, Aeta Phys. Pol. A51, 179.
Motzke, K., 1962, Phys. Status Solidi, 2, K52, Platz, W. and J. Heber, 1976, Z. Phys. B24, 333.
K307. Plumier, R., 1969, Theses, Paris.
Motzke, K., 1964, Phys. Status Solidi, 4, K13. Pointon, A.J. and G.A. Wetton, 1973, AIP
Mozzi, R.L. and A.E. Paladino, 1963, J. Chem. Conf. Proc. 10, 1573.
Phys. 39, 435. Pollert, E. and Z. Jirfik, 1976, Czech. J. Phys.
Navrotsky, A., 1973a, J. Solid State Chem. 6, B26, 481.
21. Preudhomme, J. and P. Tarte, 1971a, Spec-
Navrotsky, A., 1973b, Earth Planet. Sci. Lett. trochim Acta 27A, 961.
19, 471. Preudhomme, J. and P. Tarte, 1971b, Spec-
Navrotsky, A., 1974, J. Solid State Chem. 11, trochim Acta 27A, 1817.
10. Preudhomme, J. and P. Tarte, 1972, Spec-
Navrotsky, A., 1975, J. Solid State Chem. 12, trochim. Acta 28A, 69.
12. Prince, E., 1964, J. Phys. Rad. 25, 503.
Navrotsky, A. and L. Hughes, Jr., 1976, J. Solid Prince, E., 1965, J. Appl. Phys. 36, 161.
State Chem. 16, 185. Rabkin, L.I. and Z.I. Novikova, 1960, Ferrites
Navrotsky, A. and O.J. Kleppa, 1967, J. Inorg. (Minsk) p. 146.
Nucl. Chem. 29, 2701. Rado, G.T. and J.M. Ferrari, 1975, Phys. Rev.
Navrotsky, A. and O.J. Kleppa, 1968, J. Inorg. B12, 5166.
Nucl. Chem. 30, 479. Rado, G.T. and J.M. Ferrari, 1977, Phys. Rev.
N6el, L., 1948, Ann. Phys. (France) 3, 137. B15, 290.
N~el, L., 1954, J. Phys. Rad. 15, 225. Reinen, D., 1968, Z. Anorg. Allg. Chem. 356,
Novfik, P., 1966, Czech. J. Phys. B16, 723. 182.
Novfik, P., 1972, Czech. J. Phys. B22, 1134. Rezlescu, N. and E. Rezlescu, 1974, Phys. Sta-
Ohnishi, H., T. Teranishi and S. Miyahara, tus Solidi a23, 575.
1959, J. Phys. Soc. Jap. 14, 106. Reznickiy, L.A., 1977, Izv. Akad. Nauk SSSR,
Okada, T. and T. Akashi, 1965, J. Phys. Soc. Neorg. Mat. 13, 1669.
Jap. 20, 639. Rogers, D.B., 1967, Cs. 6as. fyz. A17, 398.
OXIDE SPINELS 303

Rogers, D.B., R.J. Arnott, A. Wold and J.B. Simga, Z., P. Sirok~, F. Lukeg and E. Schmidt,
Goodenough, 1963, J. Phys. Chem. Solids 24, 1979, Phys. Status Solidi, b96, 137.
347. Simga, Z., P. Sirok~, J. Kolfi~ek and V.A.M.
Rogers, D.B., J.L. Gilson and T.E. Gier, 1967, Brabers, 1980, J. Mag. Mag. Mater. 15-18,
Solid State Commun. 5, 143. 775.
Rosencwaig, A., 1970, Can. J. Phys. 48, 2857. Simgovfi, J. and Z. Simga, 1974, Czech. J. Phys.
Roth, W.L., 1964, J. Phys. Rad. 25, 507. B24, 449.
Rouse, K.D., M.V. Thomas and B.T.M. Willis, Simgovfi, J., M. Marygko and K. Suk, 1976,
1976, J. Phys. C9, L231. Phys. Status Solidi, a38, K163.
Samochvalov, A.A., S.A. Ismailov and A.A. Siratori, K., 1967, J. Phys. Soc. Jap. 23, 948.
Obuchov, 1967, Fiz. Tverd. Tela, 9, 884. Siratori, K., E. Kito, G. Kaji, A. Tasaki, S.
Sawatzki, G.A., F. van der Woude and A.A. Kimura, I. Shindo and K. Kohn, 1979, J.
Morish, 1969, Phys. Rev. 187, 747. Phys. Soc. Jap. 47, 1779.
Scheerlinck, D., W. Wegener, S. Hautecler and Sixtus, K.J., 1960, Solid State Physics in Elec-
V.A.M. Brabers, 1974, Solid State Commun. tronics and Telecommunication, vol. 3
15, 1529. (Academic Press, New York) p. 91.
Schlegel, A., S.F. Alvarado and P. Wachter, Slack, G.A., 1964, Phys. Rev. A134, 1268.
1979, J. Phys. C12, 1157. Slonczewski, J.C., 1958, Phys. Rev. 110, 1341.
Schl6mann, E., 1957, J. Phys. Chem. Solids, 2, Slonczewski, J.C., 1960, J. Phys. Chem. Solids,
214. 15, 335.
Schl6mann, E., 1960, Solid State Physics in Smit, J., 1968, Solid State Commun. 6, 745.
Electronics and Telecommunication vol. 3 Stair, J. and H.P.J. Wijn, 1959, Ferrites (Wiley,
(Academic Press, New York) p. 322. New York).
Scott, G.B., 1978, The Optical Absorption and Smit, J., F.K. Lotgering and R.P. van Stapele,
Magneto-Optic Spectra of Y3FesO12, in: Proc. 1962, J. Phys. Soc. Jap. 17B-l, 268.
International School on Physics "Enrico Smith, D.O., 1952, Prog. Rep. Lab. Insul. Res.
Fermi," ed., Paoletti, A. (North-Holland, MIT, 11, 53.
Amsterdam) pp. 445-466. Sobotta, E.A. and J. Voitlfinder, 1963, Z. Phys.
Scott, G.B., D.E. Lacklison and J.L. Page, Chem., Frankfurt, 39, 54.
1974, Phys. Rev. B10, 971. Sparks, M., 1964, Ferromagnetic Relaxation
Scott, G.B., D.E. Lacklison, H.I. Ralph and Theory (McGraw-Hill, New York).
J.L. Page, 1975, Phys. Rev. B12, 2562. Sugano, S., Y. Tanabe and M. Kamimura, 1970,
Seleznev, V.N., I.K. Pukhov, A.I. Drokin and Multiplets of Transition Metal Ions in Crys-
V.A. Shapovalov, 1970, Fiz. Tver. Tela 12, tals (Academic Press, New York).
885. Svirina, E.P., 1970, Izv. Akad. Nauk USSR,
Seshagiri Rao, T., B. Revathi and M. Pur- Ser. Fiz. 34, 1162.
nanandam, 1971, Indian J. Pure Appl. Phys. Szymczak, H., R. Wadas, W. Wardzynski and
9, 797. W. Zbieranowski, 1973, Proc. Int. Conf. Mag.
Serres, A., 1932, Ann. Phys. (France) 10, 32. ICM-73 (Moscow) p. 425.
Shannon, R.D., 1976, Acta Crystallogr. A32, Tachiki, M., 1960, Progr. Theor. Phys. 23, 1055.
751. Teale, R.W., 1967, Proc. Phys. Soc. 92, 411.
Shindo, I. and K. Kohn, 1979, J. Phys. Soc. Jap. Teale, R.W. and D.W. Temple, 1967, Phys.
47, 1779. Rev. Lett. 19, 904.
Shirane, G., S. Chikazumi, J. Akimitsu, K. Teh, H.C., M.F. Collins and A.H. Mook, 1974,
Chiba, M. Matsui and Y. Fujii, 1975, J. Phys. Can. J. Phys. 52, 396.
Soc. Jap. 39, 949. Thompson, P. and N.W. Grimes, 1972a, J.
Siemons, W.J., 1970, IBM J. Res. Dev. 14, Appl. Crystallogr. 10, 369.
245. Thompson, P. and N.W. Grimes, 1972b, Philos.
Simga, Z., 1979, Phys. Status Solidi b96, 581. Mag. 36, 501.
Simga, Z. and O. Schneeweiss, 1972, Czech. J. Torrie, B.H., 1967, Solid State Commun. 5, 715.
Phys. B22, 1331. Tretyakov, Yu., 1967, Termodinamika Ferritov
Simga, Z., J. ~imgovfi and V.A.M. Brabers, (Chimija, Leningrad) p. 41.
1972, Proc. llth ICPS, vol. 2 (Polish Sci. Tyablikov, S.V., 1956, Fiz. Met. Metalloved. 3,
Publishers, Warsaw) p. 1294. 3.
304 S. KRUPI~KA AND P. NOV_.~K

Umemura, S. and S. Iida, 1979, J. Phys. Soc. Westwood, W.D. and A.G. Sadler, 1971, Can. J.
Jap. 47, 458. Phys. 49, 1103.
Van der Woude, F. and G.A. Sawatzky, 1971, White, W.B. and B.A. De Angelis, 1967, Spec-
Phys. Rev. B4, 3159. trochim. Acta, 23A, 985.
Van Diepen, A.M. and F.K. Lotgering, 1977, Willshee, J.C. and J. White, 1967, Trans. Br.
Physica 86.-88B + C, 981. Ceram. Soc. 66, 548, 551.
Van Uitert, L.G., 1956a, J. Chem. Phys. 24, Wittekoek, S., T.J.A. Popma, J.M. Robertson
306. and P.F. Bongers, 1975, Phys. Rev. B12,
Van Uitert, L.G., 1956b, Proc. IRE 44, 1294. 2777.
Van Uitert, L.G., 1957, J. Appl. Phys. 28, 320. Wolf, W.P., 1957, Phys. Rev. 108, 1152.
Venero, A.F. and E.F. Westrum, 1975, J. Yafet, Y. and C. Kittel, 1952, Phys. Rev. 87,
Chem. Thermodynamics 7, 693. 290.
Verwey, E.J.W., 1 9 5 1 , Semiconducting Yakovlev, Y.M., E.V. Rubalskaya, L.G. Godes,
Materials (Butterworth Sci. Publications, B.L. Lapovok and T.N. Bushueva, 1971, Fiz.
London) p. 151. Tver. Tela 13, 1151.
Verwey, E.J.W. and P.W. Haaijman, 1941, Yamada, T., 1973, J. Phys. Soc. Jap. 35, 130.
Physica 8, 979. Yamada, T., 1975, J. Phys. Soc. Jap. 38, 1378.
Vi~fiovsk~, S., R. Krishnan, V. Prosser, N.P. Yamada, N. and S. Iida, 1968, J. Phys. Soc. Jap.
Thuy and I. St[eda, 1979a, Appl. Phys. 18, 24, 952.
243. Yanase, A., 1962, J. Phys. Soc. Jap., 17, 1005.
Vi~fiovsk3), S., N.P. Thuy, J. St6p~inek, R. Yasuoka, H., A. Hirai, M. Matsumura and T.
Krishnan and V. Prosser, 1979b, J. Appl. Hashi, 1962, J. Phys. Soc. Jap. 17, 1071.
Phys. 50, 7466. Yosida, K. and M. Tachiki, 1957, Prog. Theor.
Vi~fiovsk~), S., R. Krishnan, N.P, Thuy, J. St6p- Phys. 17, 331.
~inek, V. Pa[izek and V. Prosser, 1980, J. Yshikawa, Y., S. Sato and Y. Syono, 1971, J.
Mag. Mag. Mater. 15--18, 831. Phys. Soc. Jap. 31, 453.
Vratislav, S., J. Zaj/6ek, Z. Jir~ik and A.F. Zanmarchi, G. and P.F. Bongers, 1969, J. Appl.
Andresen, 1977, J. Mag. Mag. Mater. 5, 41. Phys. 40, 1230.
Wagner, R., 1961, Ann. Phys. 7, 302. Z~iv6ta, K. and P. Nov~k, 1971, J. Phys.
Waldron, R.D., 1955, Phys. Rev. 99, 1727. (France) 32, C1-64.
Wanic, A., 1972, Int. J. Mag. 3, 349. Zftv6ta, K., E.I. Trinkler and F. Zounov~, 1966,
Watanabe, H. and B.N. Brockhouse, 1962, Phys. Status Solidi 14, K9.
Phys. Lett. 1, 189. Z~iv6ta, K., F. Zounov~i and E.I. Trinkler, 1968,
Watanabe, Y., K. Urade and S. Saito, 1978, Czech. J. Phys. B18, 1314.
Phys. Status Solidi, b90, 697. Zhilyakov S.M. and E.P. Naiden, 1977, Izv.
Wegener, W., D. Scheerlinck, E. Legrand, S. V.U.Z. Fiz no. 1, 56.
Hautecler and V.A.M. Brabers, 1974, Solid
State Commun. 15, 345.
chapter 5

FUNDAMENTAL PROPERTIES OF
HEXAGONAL FERRITES WITH
MAGNETOPLUMBITE
STRUCTURE

H. KOJIMA
Research Institute for Scientific Measurements
Tohoku University, 2-1-1 Katahira, Sendai
JAPAN

Ferromagnetic Materials, Vol. 3


Edited by E.P. Wohlfarth
© North-Holland Publishing Company, 1982

305
CONTENTS

1. G e r e r a l . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
1.1. C h e m i c a l c o m p o s i t i o n . . . . . . . . . . . . . . . . . . . . . 307
1.2. P h a s e d i a g r a m . . . . . . . . . . . . . . . . . . . . . . . 308
1.2.1. B a O - F e 2 0 3 s y s t e m s . . . . . . . . . . . . . . . . . . . 308
1.2.2. S r O - F e 2 0 3 s y s t e m s . . . . . . . . . . . . . . . . . . . 312
1.2.3. P b O - F e 2 0 3 s } s t c m s . . . . . . . . . . . . . . . . . . . 313
1.3. P r e p m a t i o n . . . . . . . . . . . . . . . . . . . . . . . . 315
2. M c o m p o u n d . . . . . . . . . . . . . . . . . . . . . . . . . 317
2.1. BaFe1:.O19, SrFe12019 ~ n d Pb Fe1=O19 . . . . . . .. . . . . . . . . . 318
2.1.1. Crystal s t r u c t u r e . . . . . . . . . . .. . . . . . . . . . 318
2.1.2. M a g n e t i c ~,t~ u c t u r e . . . . . . . . . . . . . . . . . . . 323
2.1.3. S a t m atic n me gnet!2 a t i c n . . . . . . . . . . . . . . . . . 325
2.1.4. M a g n e t o c r ~ stalline a n i s o t r c py . . . . . . . . . . . . . . . 329
2.1.5. C o e r c i v e f o l c e . . . . . . . . . . . . . . . . . . . . . 334
2.1.6. F a r a m ~ g n e l i c proloert~e s . . . . . . . . . . . . . . . . . 342
2.1.7. Magnetic aftereffect . . . . . . . . . . . . . . . . . . . 344
2.1.8. FMR . . . . . . . . . . . . . . . . . . . . . . . . 345
2.1.9. NMR . . . . . . . . . . . . . . . . . . . . . . . . 347
2.1.10. M 6 s s b a u e r effe~ t . . . . . . . . . . . . . . . . . . . 351
2.1.11. Domain observation . . . . . . . . . . . . . . . . . . 354
2.1.12. Optical pioperties . . . . . . . . . . . . . . . . . . . 358
2.1.13. M a g n e l ostriction . . . . . . . . . . . . . . . . . . . 360
2.1.14. M e c h a n i c a l proFertie s . . . . . . . . . . . . . . . . . . 361
2.1.15. H e a t c a p a c i t y . . . . . . . . . . . . . . . . . . . . 361
2.1.16. q h m m a l e x p a n s i c n . . . . . . . . . . . . . . . . . . . 362
2.1.17. E l e c t r i c a n d dielectric IZrOl:erties . . . . . . . . . . . . . . 364
2.2. Substitute d M c o m p o u n d . . . . . . . . . . . . . . . . . . . 367
2.2.1. B a O - S r O - F b O - F e 2 0 3 s y s t e m s . . . . . . . . . . . . . . . 367
2.2.2. G t h e r substitutions c f Ba:+ ions . . . . . . . . . . . . . . . 368
2.2.3. Substitt t i c n c f Fe3+ l'ons . . . . . . . . . . . . . . . . . 368
Sub~,titution with A l>, Ga3+ a n d C r3+ icn . . . . . . . . . . . 368
Sub,~ tilution with Sc3+ a n d In 3+ . . . . . . . . . . . . . . 375
Substitution of Fe3+ with p a i r e d ions; C o - T i s y s t e m . . . . . . . 378
C t h e r c c m b i n a t i o n s with I I - I V pairs . . . . . . . . . . . . 379
2.2.4. Effect of substitutions on t h e t e m p e r a t u r e d e p e n d e n c e of m a g n e t i z a t i o n 38O
2.2.5. Substitution with a n i c n s . . . . . . . . . . . . . . . . . 384
References . . . . . . . . . . . . . . . . . . . . . . . . . . . 387

306
1. General

A group of ferrimagnetic oxides with hexagonal structures will be described in this


section'S. Most of these compounds have been developed over the past two
decades and it can be said that the first fundamental step of the investigations on the
properties of hexagonal ferrites, now seems to be nearly completed. Smit and Wijn
(1959) and Wijn (1970) collected comprehensive data on the compounds, and
moreover, many authors - such as Schieber (1967), Tebble and Craik (1969), Galasso
(1970), Standley (1972), and others-spent a few chapters to treat the magnetic
properties or crystal structures of the hexagonal iron oxides in their books.

1.1. Chemical composition

~fhe chemical compositions of the hexagonal compounds are shown in fig. 1" as
part of a ternary phase diagram and in table 1 for the BaO-MeO-Fe203 system.
Here, Me represents a divalent ion among the first transition elements, Zn, Mg, or
a combination of ions whose valency is two. S denotes a cubic spinel MeO.Fe203
and will be explained in detail in ch. 3 of this handbook by Krupi6ka, in vol. 2, ch.
3 by Slick (1979) and in vol. 2, ch. 4 by Nicolas (1979). However, we refer to it
here as to one of the major constituent blocks of the hexagonal compounds. M
compounds, which have the chemical formulae of BaO.6Fe203 (BaM),
SrO.6Fe203 (SrM) and PbO.6Fe203 (PbM) etc., were developed in the initial stage
by Went et al. (1952) and Fahlenbrach and Heister (1953) as a typical hexagonal
ferrimagnetic oxide for permanent magnet materials. They are isomorphous with
the mineral magnetoplumbite, the chemical composition of which is ap-
proximately PbzFe15Mny(A1Ti)O38.
Other interesting hexagonal compounds are potential industrial magnetic
materials in the systems: 2MeO-BaO.8Fe203 (Ba-Me-W), 2MeO-2BaO.6Fe203
(Ba-Me-Y), 2MeO.3BaO.12Fe203 (Ba-Me-Z), 2MeO.2BaO.14Fe203 (Ba-Me-
X), and 2MeO.4BaO.18Fe203 (Ba-Me-U); they have more complicated crystal
structures than BaM, and were originally investigated by Jonker et al. (1954) for

? CGS unit are used in this section. SI units are shown in brackets.
* More detailed diagrams of B a O - Z n O - F e 2 0 3 and SrO-ZnO--Fe203 systems were recently reported
by Slokar and Lucchini (1978a, b).

307
308 H. KOJIMA

/
°,°%\/\
\~ / \\\/ \ %

E}00
Fig. 1. BaO-MeO-Fe203 system, showing the relationships of chemical compositions among fer-
rimagnetic hexagonal compounds (see table 1).

use at very high frequencies. They will be described in ch. 6 by Sugimoto, and in
vol. 2, ch. 3 (Slick 1979) and in vol. 2, ch. 4 of this handbook (Nicolas 1979). While
BaO.Fe203 (B) was reported as an antiferromagnetic hexagonal compound by
Okazaki et al. (1961a, b), here, it is rather important as an intermediate phase,
when BaM is prepared from Fe203 and BaO (Beretka and Ridge 1958, St~iblein
and May 1969, and Wullkopf 1975). In the BaO-Fe203 system, BaO-2Fe203 (T) does
not exist as a stable phase at room temperature (Okazaki et al. 1961a) but it is an
essential constituent block of Y, Z and U compounds. The constitutions and phase
relationships of each compound will be evident from fig. 1 and table 1.

1.2. Phase diagram

1.2.1. BaO-Fe203 systems


Numerous compounds have been found in the BaO-Fe203 system, for example,
7BaO.2Fe203 (Batti 1960), 5BaO-Fe203 (Bye and Howard 1971), 3BaO.Fe203
(Okazaki et al. 1961a, Bye and Howard 1971), 2BaO.Fe203 (Erchark et al. 1946),
BaO.Fe203 (Okazaki et al. 1955), 2BaO.3Fe203 (Okazaki et al. 1961a, Appendino
and Montorsi 1973), 5BaO-7Fe203 (Appendino and Montorsi 1973), BaO-2Fe203
(Okamoto et al. 1975), BaO-6Fe203 (Adelsk61d 1938, Went et al. 1952),
3BaO-4FeO-14Fe203 (Brady 1973), and in an oxidizing atmosphere BaFeO3-x
(Mori, 1970), and further in a reducing atmosphere, BaO-FeO.3Fe203 (Braun 1957),
BaO.FeO.7Fe203 (Braun 1957) and BaO.2FeO.8Fe203 (Wijn 1952, Neumann and
Wijn 1968).
Figure 2 is the phase diagram determined by Goto and Takada (1960) under the
oxygen partial pressure of Po2 = ~ arm [0.02 MPa] in the solid phase region and
FUNDAMENTAL PROPERTIES OF H E X A G O N A L FERRITES 309

t~

tt3 ,,D tt~

¢¢3 ~o ~=~.
¢¢5 tt~ ~ = ~
e~

0
I I
tt3
d

e--
m
o ©
m d66666

< o66 ~66666


© © © © ©

Nmm
0

0
c-
o
0

e-

i e-

e"

*
310 H. K O J I M A

Fez03 / BoO
0.5 1,0 1.5 2.0 4.0 6.0
I I I I

1600
"',~ 1565.
L
/ /
1400
o0 1370
370
1330

O912oo
El.
E (I) I (2) {3)
O9

I000

800 I i [
~' 40 60 8o ~ Ioo
2BoO'Fe203 BoO'Fe203 BoO-6Fe203 Fee03
Mole %
Fig. 2. Phase diagram of a BaO-Fe203 system (Goto and Takada 1960). Poz = { atm [0.02 MPa] in the
solid phase region and Po2= 1 atm [0.1 MPa] in the liquid phase region. (1) 2BaO.Fe203 + BaO.Fe203;
(2) BaO-Fe203 + BaO.6Fe203; (3) BaO-6Fe203 + Fe203.

Po2 = 1 atm [0.1 MPa] above the eutectic temperature. Batti (1960) and Van H o o k
(1964) obtained somewhat different phase relations for the same system at
Po2 = ½atm [0.02 MPa] and Po2 = 1 arm [0.1 MPa], respectively, which are given in
fig. 3 and fig. 4. The main differences among these diagrams around the corn-

/
/
/
1500 /
/ 1474

L
1420

p 1400
O9 1380 / ~ (3)
E
I-
. . . . ,2,
1300

BaO 20):2 2!, 40 ,!, 60 8o,!6 Fe203


Weight %
Fig. 3. Phase diagram of a BaO-Fe203 system (Batti 1960). Po2 = ½atm [0.02 MPa]. (1) 2BaO.Fe203 +
BaO.Fe203, (2) BaO.Fe203 + BaO-6Fe203, (3) BaO.6Fe203 + Fe203.
FUNDAMENTAL PROPERTIES OF H E X A G O N A L FERRITES 311

1600
.--%U
Liq
,9,
1500
i5 (6)1(7) 1495
//" "'*,%
/ \
1400 % (5) !
/(I) (I) x ¢ C2)
Q..
i /
E 1330 v
1300 1315
13) (4)

BoO 4 0 6O 80 Fe203
Mole %
Fig. 4. Phase diagram of BaO-Fe203 (Van Hook 1964). Po2 = 1 atm [0.1 MPa]: (1) BaO-Fe203+ liq.,
(2) BaO-6Fe203 + liq., (3) BaO-6Fe203 + BaO-Fe203, (4) BaO-6Fe203 + Fe203, (5) BaO-6Fe203 +
Fe304, (6) 2FeO-2BaO-14Fe203 + liq., (7) 2FeO-2BaO-14Fe203 + Fe304, (8) 2FeO.BaO-8Fe2Oa + liq.,
(9) 2FeO-BaO.8Fe203 + Fe304, (10) Fe304 + liq.

Temperature (°C)
1500 1400 1500
102 ,~]) [ II I !

I0
13_

/
0
X

E
io-I
('4

io-Z

i, -~BaW
(3]
Io-3
0.54
I 1.58 0.62
il 0.66
I/T (10 -3 K-~ )
Fig. 5. P - T diagram of Fe304, BaO-6Fe203 (Van Hook 1964), 2FeO-BaO-8Fe~O3 (Neumann and
Wijn 1968) and 2FeO.SrO.8Fe203 (Goto et al. 1974).
312 H. K O J I M A

position of BaO.6Fe203 are the solubility limit into the B phase and the existence
of a phase transition at high temperatures. Though there is still some ambiguity,
the solid solution range is considered to be practically narrower than
Fe203/BaO = 5.0-6.0 as shown in fig. 2. According to Batti (1960) and Van Hook
(1964) it is within 5.%6.0, and 5.8--6.0 according to Stfiblein and May (1969).
Moreover, almost no solubility range is claimed by Reed and Fulrath (1973).
As for the liquidus line, it seems to be most probable that BaM dissociates into
a liquid phase and 2FeO-2BaO.14Fe203 (Ba-Fe-X) as shown in fig. 4, even when
it is heated in air (/°o2 = ½arm [0.02 MPa]). In fact, an incongruent melting will
occur in the composition. Figure 5 is the equilibrium diagram of (1) BaO.6Fe203
(BaM) ~- FeO.BaO-7Fe203 (Ba-Fe-X) + liquid + 02 and (2) Fe203 ~- Fe304+x + 02
(Van Hook 1964), and also (3) the stable region of 2FeO.BaO.8Fe203 (Neumann
and Wijn, 1968). The latter authors obtained B a - F e - W under the condition of
/9o2 = 0.05-0.1atm [5-10x 103pa] at 1400°C as a single phase. But Van Hook
(1964) found that BaM is congruently melted in Po2 = 40 atm [4 MPa] as shown in
fig. 6. Figure 7 is the phase diagram by Sloccari (1973), which shows the existence
of the peritectoid reaction, BaO.Fe203 (B)+ BaO.6Fe203 (BaM)~-2BaO-3Fe203,
at 11500-+ 10°C.

1.2.2. SrO-Fe203 systems


The phase relations are quite similar to that of the BaO-Fe203 system. Figure 8 is
the equilibrium diagram in Po2 = ½atm [0.02MPa] given by Goto et al. (1971,

//
1600, //
/
/
/ (5)
Liq ,j
z/

1 /
/ /

/ f/
1400 (I) ', ,/" (3)
....... ,_2 .........
P
I1)

E
I1)
I- (1) (2) (4)

1200

I
BaO 6o 8o Fe203
Mole %
Fig. 6. Phase diagram of a BaO-Fe203 system (Van Hook 1964). P o 2 = 4 0 a t m [4MPa]: (1)
BaO'Fe203 + liq., (2) BaO.Fe203 + BaO.6Fe203, (3) BaO.6Fe203 + liq., (4) BaO.6Fe203 + Fe203, (5)
Fe203 + liq.
FUNDAMENTAL PROPERTIES OF H E X A G O N A L FERRITES 313

Mole % Fe=O 3
50 60 70 80 85.71
I[I 1
1200 ; o,oo~ ? o
(3)
D 000, i 0
D 0000 0
/
I100= 0-00~. 0

D 0000 O.

°
iooo I
a)
r~
F: • oooe o
1-
9O0
i (I) (2)

I
80C

I Ili I
//\\
ol
m

B(:]O/Fe203
Fig. 7. Phase diagram of a BaO-Fe203 system (Sloccari 1973). (1) BaO-Fe203+ 2BaO.3Fe203, (2)
2BaO.3Fe203 + BaO.6Fe203, (3) BaO-Fe203 + BaO-6Fe203, (0) one phase, ((3) two phase.

1974). They insisted from the results of ditterential thermal analysis that there
might exist some solubility range of SrM in 3SrO.2Fe203, as indicated with the
broken line in fig. 8. SrM is decomposed at 1435°C to SrFe]sOz7 (Sr-Fe-W)+
liquid, and at 1465°C, they change into Fe304 + liquid in air. The stable region of
Sr-Fe-W is also indicated by the region (4) in fig. 5 from the results of Goto et al.
(1974). Furthermore, they obtained Sr-Fe-X by heat treatment at 1420°C for two
hours, but it is decomposed by prolonged heating to a mixed phase containing
Sr-Fe-W in air

1.2.3. PbO-Fe203 systems


The phase relations obtained by Mountvala and Ravitz (1962) and Berger and
Pawlek (1957) are considered as reasonable representations under normally
attainable equilibrium conditions. Figure 9 is the diagram by Mountvala and
Ravitz (1962). PbM is incongruently melted to Fe203 + liquid at 1315°C (1255°C) *,
and below 760°C (810°C) * it is decomposed to FezO3+ PbO.2Fe203. Mountvala

* Data by Berger and Powlek (1957).


314 H. KOJIMA

,,./ ~ " (I0)


1500 ,,1480 / 1465
\',,, Liq ~ (8)
, / ,~5 (7)11395(9)
(2) X I/ (5)
o
1300
/ ,,
(3) ~ (/
1195 "\
09
~- I 1 0 0

( I) (4) (6)

900

40 60 80 I00
SrFeO3-x Mole % Fe203
Fig. 8. Phase diagram of a SrO-Fe203 system (Ooto et al. 1971, 1974). Poz=½atm [0.02MPa]: (1)
SrFeO3_x + 3SrO.2Fe203, (2) SrFeO3-x + liq., (3) 3SrO'2Fe203 + liq., (4) 3SrO'2Fe203 + SrO:6Fe203,
(5) SrO'6Fe203 + liq., (6) SrO'6Fe203 + Fe203, (7) SrO.6Fe203 + 2FeO-SrO-8Fe203 +
(2FeO-2SrO. 14Fe203), (8) 2FeO.SrO.8Fe203 + liq., (9) 2FeO.SrO.8Fe203 + Fe304, (10) Fe304 + liq.

2:1 1:2 2:5 1:5 1:6


14oo
/
/
1315
13oo /
/ 1
/ I
/
12oo / I
I
I
S /iq I
o II00 I
(9) I
I I (H)
i
~ I000
945
i
i
i
(8)
9oo ~- ! ,,/---I 9,ol
(4) 5 (61 , "
800 (2)

7ooF.... 7~oq
\\\2 /
1/
I I i (3)
750
(io)
/
760

I- !" L_~ 650


600 " '
0 20 40 60 80 I00
PbO Mole % Fe203
Fig. 9. Phase diagram of a PbO-Fe203 system (Mountvala and Ravitz 1962). (1) PbO + 2PbO.FezO3,
(2) PbO + liq., (3) 2PbO.Fe203 + Fe203, (4) 2PbO.Fe203 + PbO-2Fe203, (5) PbO.2Fe203, (6)
PbO-2Fe203 + PbO.6Fe203, (7) PbO-6Fe203, (8) PbO.2Fe203 + liq., (9) PbO.6Fe203 + liq., (10)
PbO-2Fe203 + Fe203, (11) PbO.6Fe203 + Fe203, (12) Fe203 + liq.
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 315

and Ravitz (1962) also claimed that there is a rather wide solid solution range of
Fe203/PbO = 5.0-6.0 between 800°-945°C and proposed a crystal structure for the
end member of the solid solution, PbO.5Fe203. It would be reasonable to explain
the fact that the lattice parameters were found to be independent of composition
in this solid solution range, but the saturation magnetization of PbO.5Fe203 is
higher than that of PbO'6Fe203. There seems to be no investigation about the
hexagonal phase of P b - M e - W , P b - M e - X etc. in this system.

1.3. Preparation

The most typical way to obtain the ferrimagnetic hexagonal oxides as powder or
in a sintered polycrystalline state is the solid state reaction of heating the mixtures
of constituent oxides or of compounds which are easily changed to oxides by
heating. BaCO3, SrCO3, PbO and c~-Fe203 etc. are generally used as starting
materials, but oxalates, sulfates, chlorides, nitrates or hydroxides are also used for
specific purposes. Proper reaction temperature, atmosphere and cooling con-
ditions should, of course, be chosen according to the phase diagrams. However,
single phase M type compounds, for example, can be usually obtained by heating
in air (PQ = ~ atm [0.02 MPa]) between 800 ° and 1200°C, and just removing from
the furnace.
The formation processes of the reactions have been reported by many authors
from various points of view. Erchak Jr, et al. (1946), and Erchak Jr. and Ward
(1946) investigated the reaction between ferric oxide and barium carbonate and
found the formation of BaO.2Fe203 above 550°C and BaO.6Fe203 above 750°C
by X-ray diffraction. Sadler (1965) studied the reaction kinetics of BaM, and
obtained the activation energy of 73.2 kcal/mol [3.06× 105 J/mol] above 735°C.
Wullkopf (1972, 1975) observed the variations of the amount of reaction products,
length, weight and grain size for the mixed compact of BaCO3 and 5.5Fe203
during the sintering process and collected the data, as shown in fig. 10. Similar
results were also reported by Haberey et al. (1973). Furthermore, Haberey and
Kockel (1976) found that SrM is formed from a mixture of SrCO3 and 6Fe203
through the following two endothermal reactions:

SrCO3 + 6c~-Fe203 + (0.5 - x) × 102 --~ SrFeO3-x + 5.5a-Fe202 + CO2


SrFeO3-x + 5.5ol-Fe203---> SrO-6Fe203 + (0.5 - x) × ½02.

Bowman et al. (1969) investigated the formation mechanism of PbM in the


PbO-Fe203 system and concluded that PbM is formed through the intermediate
compounds of 2PbO-Fe203 or PbO'2Fe203, depending upon the mixing method,
the time and temperature of heating. They used two kinds of co-precipitation
methods for mixing, in which the aqueous solution of ammonium bicarbonate was
added to the solutions containing lead nitrate and ferric nitrate, or lead nitrate
and ferric oxide.
The co-precipitation method was also used to obtain a high coercivity BaM or
316 H. KOJIMA

1.0

0.8
(b) ' ~ \\k\\\\\ M/
0.6
.o
0.4
\/,/,, \/
\/-
\ \
/ \
0.2 /
! !
0 ! I

/
1.0
(0)
•~ 0.8
(D

>, 0.6
P

" 0.4
'l
\
z
"~ 0.2
rr
//. w

0
0 400 800 1200
i./411600
Temperoture (°C)
Fig. 10. Changes of BaO-5.5Fe203 compact during heating (Wullkopf 1972, 1975): (a) variation of
reaction products, F: Fe203, Bc: BaCO3, B: BaO.Fe203, M: BaO.6Fe203, Y: 2FeO.2BaO.6FezO3, W:
2FeO-BaO-8Fe203, Z: 2FeO-3BaO.12Fe203; (b) variation of length L, weight G, grain size D and
saturation magnetization M.

SrM by Mee and Jeschke (1963), Haneda et al. (1974a), Roos et al. (1977) and Oh
et al. (1978). Furthermore, Shirk and Buessem (1970) obtained a high coercivity
BaM from a glass with the composition 0.265BzO3-0.405BaO-0.33Fe203 in mole
ratio. They reported that single domain particles can be crystallized in the
fast-quenched glass with this composition by the heat treatment under the
appropriate condition (see also table 10). A molten salt synthesis utilizing NaC1-
KC1 for BaM and "SrM submicron crystals with high magnetic quality was
proposed by Arendt (1973). Moreover, Okamoto et al. (1975) applied hydro-
thermal synthesis with ol-Fe203 suspension in barium hydroxide aqueous solution
and obtained BaO.2Fe203 crystals, whose space group was reported as P63/m
with the lattice parameter of a = 5.160 A and c = 13.811 A. Kiyama (1976) obtained
BaO.6Fe203, BaO.4.5Fe203 and BaO.2Fe203 with Fe(OH)3 or FeOOH and
Ba(OH)2 under the similar conditions in an autoclave and studied magnetic
properties.
Single crystals of M compounds were obtained by cooling a nearly eutectic melt
(Kooy 1958); under the high oxygen pressure (Van Hook 1964, Menashi et al.
1973); using various kinds of flux (Mones and Banks 1958, Brixner 1959, Linares
F U N D A M E N T A L P R O P E R T I E S OF H E X A G O N A L F E R R I T E S 317

1962, Suemune 1972, Aidelberg et al. 1974); or discontinuous grain growth


(Lacour and Paulus 1968, 1975). B a - Z n - Y , B a - Z n - W , B a - Z n - Z , B a - C o - Z n - W
and B a - C o - Z n - Z etc. were also grown by the flux method (Tauber et al. 1962,
1964, Savage and Tauber 1964, AuCoin et al. 1966, Suemune 1972). BaC12,
BaO-B203, BaO-B2Og-PbO, Na2CO3 and NaFeO2 were recommended as a flux
for these compounds. Takada et al. (1971) found that topotactic reactions among
a - F e O O H or ~-Fe203 and BaCO3 or SrCO3 are effective to obtain a grain
oriented specimen. The crystallographic relationships of the materials are

( 100 )~-F~OOH//(0001)~,_F~203//(0001)S~O.6Fe203,
[010],,-F~OOH//[11,201a -Fe203//[1010]SrO.6F~203 •

Hot press or hot forging processes are also useful to prepare a dense oriented
sintered body (St~iblein 1973, Haneda et al. 1974a).

2. M compound

A Ba 2+ ion in the M compound, BaO-6Fe203 (BaM), can be replaced partly or


completely by Sr 2+, Pb 2+ and a combination of, for instance, Agl++ La 3÷ or
Nal++ La 3+, without changing its crystal structure. Substitutions of Fe 3÷ and 02-
ion in the compound are also possible. In all cases, substituted ions would be
chosen to keep electrical neutrality and to have similar ionic radii with the original
ions (see table 2; a more comprehensive table of ionic radii can be seen, for
example, in the book of Galasso (1970)). BaM was at first the only main
constituent of M-type oxide magnet, produced on an industrial scale but

TABLE 2
Ionic radii of several related ions (Pauling 1960).

Element valence r (A) Element valence r (A) Element valence r (4)

Ag +1 1.26 Ga +3 0.62 P +3 0.44t


AI +3 0.50 Ge +4 0.53? +5 0.35t
As +3 0.58? In +3 0.81 Pb +2 1.207
+5 0.46t Ir +4 0.68? Sb +3 0.76?
Ba +2 1.35 La +3 1.15 +5 0.62?
Bi +3 0.96? Li +1 0.60 Sc +3 0.81
Ca +2 0.99 Mg +2 0.65 Sn +4 0.71
C1 - 1 1.88 Mn +2 0.80? Sr +2 1.13
Co +2 0.72? +3 0.66t Ta +5 0.68t
Cr +3 0.63? +4 0.60? Ti +4 0.68
Cu +2 0.72? Na +1 0;95 T1 +3 0.95
F - 1 1.36 Nb +5 0.70 V +5 0.59
Fe +2 0.74t Ni +2 0.69? Zn +2 0.74
+3 0.64t O - 1 1.40 Zr +4 0.80

? Ahrens (1952).
318 H. K O J I M A

SrO-6Fe203 (SrM) has more recently taken over some part of BaM. PbO.6Fe203
(PbM) is used only as an additional material for oxide magnet purposes at present
(see ch. 7 by Stfiblein for the applications). In the following section the fun-
damental properties of BaM, SrM and PbM are described. The properties of solid
solutions among BaM, SrM and PbM, and substituted M compounds are treated
separately in this chapter.

2.1. BaFele019, SrFe12019 and PbFea2019

2.1. i. Crystal struclure


Adelsk61d (1938) determined the crystal structures of BaM, SrM and PbM,
prepared by heating co-precipitated mixtures from the solutions of nitrates. Figure
11 is a perspective drawing of BaM. The 02 ions form a hexagonal close packed
lattice, so that its layer sequence perpendicular to the [001] direction is A B A B . . .
or A C A C . . . as is shown in the figure. Every five oxygen layers, one O 2 ion is
replaced with Ba 2+, Sr 2+ or Pb 2+ in BaM, SrM or PbM respectively and this occurs
due to the similarity of their ionic radii as given in table 2. Five oxygen layers

(15)
(12)

(11) A

(10) C

(9) A

(8) C

(7)

(6) B

{5)

(4)

(3)

(2)

Fig. 11. Perspective illustration of BaO.6Fe203.


FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 319

make one molecule and two molecules make one unit cell. Each molecule shows
180° rotational symmetry around the hexagonal c-axis against the lower or upper
molecule. The O z- layer containing Ba 2+ is a mirror plane, being perpendicular to
the c-axis. Fe 3+ ions occupy the interstitial positions of the oxygen lattice. The
space group of the compound is denoted as P63/mmc (D4h) using H e r m a n n -
Mauguin's (Sch6nflies') symbols.
Figure 12 illustrates more clearly the layer structure of BaM, where z means
the layer height along the [001] direction and the layer numbers are the same as
in fig. 11. Explanation of the symbols used in the figures are also given here.
Wyckoff's notations are adopted for every site in the crystal (Henry and Lonsdal
1952). The positions of each atom are tabulated in table 3 (Galasso 1970). Figure

~ C~) 0 -2 ion
(6) 0.45 {12) 0,95
(~) Be2+ ion

Fe3+(4f2)IOctahedral site
(5) 0.35 (11) 0,85 (~ Fe3+(2o)J
T

141 0.25 clot 0.75


(~ Fe3+ under the layer

~ Lt~ Fe3+ above the layer


T relative orientation
of magnetic moment
(5) 0.15 (9) 0.65

(2) 0.05 (8) 0.55

(I) Z=O (7) 0.50 (13) 1.00


Fig. 12. The layer sequence of BaO.6Fe203.
320 H. KOJIMA

TABLE 3
Atomic positions of BaFe120~9 (see fig. 12) (Galasso 1970).

Ion Site Coordinate x z

2Ba 2+ 2d I, 2, 2; 2, ½, ¼ - -
2a 0,0,0;0,0,1 - -
2b 0,0,¼;0,0, 3 - -
24Fe 3+ 4fl ±(½, 2, Z; 2, 5,Z ~q_)l I -- 0.028
4f/ ___(1, 2, z ; 2, 3,1 ~+z)
1 0.189
12k +(x, 2x, z; 2x, x, Y.; x, 2, z; x, 2x, l - z; 2x, x, ½+ z; £, x, ½+ z) 0.167 -0.108

4e +(0, 0, z; 0, 0, 1+ z) - 0.150
1 2 2
4f ±(3, 3, z; ~, ½, ½+ z) - -0.050
3802- 6h ±(x, 2x, ¼; 2x, x, 3; x, g, ¼) 0.186 -
--+(x, 2x, z ; 2x, x, 2 ; x, 2, z; x, 2x, ~1- z; 2x, x, ~+z; £, x, 1+z) 0.167 0.050
1 1 1
12k ±(x, 2x, z ; 2 x , x, 2;x, 2, z ; x , 2 x , ~ - z ; 2 x , x , ~ + z ; x , x , ~ )
-

0.500 0.150

13, t h e (110) s e c t i o n of B a M , is a n o t h e r e x p r e s s i o n o f t h e c r y s t a l s t r u c t u r e ,
s h o w i n g a t o m s a n d s y m m e t r y e l e m e n t s in a m i r r o r p l a n e c o n t a i n i n g t h e c - a x i s
( B r a u n 1957). S a n d R a r e t h e b u i l d i n g b l o c k s of t h e crystal, a n d S* a n d R *
i n d i c a t e t h e b l o c k s , o b t a i n e d by r o t a t i n g S a n d R t h r o u g h 180 ° a r o u n d t h e c-axis,
as p r e v i o u s l y i l l u s t r a t e d . It can b e said, t h e r e f o r e , t h a t t h e u n i t cell of B a M is
21
63 6 C 6 63

++3++2,

R~

ooz ~'~hz ~U/~z iIz


[JTo]
Fig. 13. The (110) cross section of BaO-6Fe203.
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 321

expressed as RSR*S*. Moreover, Townes et al. (1967) refined the crystal structure
of BaM by X-ray investigation. They stated two points: o(i) The Fe 3÷ ion in the
trigonal bipyramidal site, 2b, is split into half atoms 0.156 A away from the mirror
plane, 4e. (ii) Some iron octahedra occur in pairs which share a common face to
form Fe209 coordination groups. The former is supported by some M6ssbauer
investigators (see ch. 2 section 4.2.1 M6ssbauer effect). Atomic coordinates,
interatomic distances and structure factors are tabulated in their paper which
gives more accurate results.
Figure 14 shows perspective drawing of the R (BaFe6Oll), S (Fe6Os) and T
(Ba2FesO14) blocks separately. The T block is related only with the Y, Z and U

I- 0

Fig. 14. Perspective drawings of building blocks in the hexagonal compounds, T(Ba2Fe8014), S(Fe6Os)
and R(BaFe6On).
322 H. KOJIMA

j Qlf

j ~ J

Fig. 15. Unit cell of BaO.6Fe20), showing the crystal structure composed of spinel blocks and Ba
layers (Gorter I954).

TABLE 4
Lattice constants, molecular weights and X-ray densities of M-type compounds.

Lattice constant
Molecular X-ray density
Compound weight (g/mol) a (,~) c (,~) c/a (g/cm ~) Ref.

5.893 23.i94 3.936 5.29 (a)


5.89 23.20 3.94 5.29 (b)
BaFe12019 1111.49
5.889 23.182 3.936 5.30 (c)
5.876 23.17 3.943 5.33 (d)

5.885 23.047 3.916 5.10 (e)


SrFe12019 1061.77 5.876 23.08 3.92s 5.11 (c)
5.864 23.031 3.928 5.14 (d)

5.877 23.02 3.917 5.70 (d)


PbFe12019 1181.35
5.889 23.07 3.917 5.66 (c)

(a) Tauber et al. (1963)


(b) Smit and Wijn (1959)
(c) Bertaut et al. (1959)
(d) Adelsk61d (1938)
(e) Routil and Barham (1974)
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 323

c o m p o u n d s but is shown here to illustrate the relationship a m o n g the blocks (see


ch. 2 sections 4.l and 4.3). S has the cubic spinel structure with the [ l l l ] axis
vertical. In other words, M is synthesized by piling up a Ba layer and a spinel
block, whose layer sequence is A B C A . . . . alternatively (see fig. 11). G o r t e r
(1954) showed this simply by fig. 15, where spinel block contains four oxygen
layers. Molecular weight, X-ray density and lattice constants reported by various
investigators are tabulated in table 4.

2. 1.2. Magnetic structure


M-type c o m p o u n d s have a typical ferrimagnetic structure, that is, the orientation
of the magnetic m o m e n t s of the ferric ions in the crystal are generally aligned
along the c-axis in antiparallel with each other. Ndel (1948) and A n d e r s o n (1950)
first considered from the theoretical view point that these alignments of magnetic
ions can be realized by superexchange interaction through oxygen ions and such a
structure has been proved from the experimental results of saturation mag-
netization, neutron diffraction, M6ssbauer effect and nuclear magnetic resonance
etc. Grill and H a b e r e y (1974) calculated the exchange parameters of Fe > ions in
BaM, as shown in table 5. H e r e it can be clearly seen that the closer the angle of
the F e - O - F e b o n d approaches 18(t°, the larger the exchange p a r a m e t e r b e c o m e s

TABLE 5
Distances and angles of the Fe-O-Fe bonds and calculated exchange parameters in
BaFe12Ot,~ (Grill and Haberey 1974).

Distance Angle Exchange Calculated


Bond (A) (degree) parameter value (K//z~)

'~ Fe(b')-OR2Fe(f2) { 1.886 + 2.060 142.41


]' Fe(b')-OR2-Fe(f:) ,~ 1.886 + 2.060 132.95 Jbf2 35.96

{ Fe(f0-Os~-Fe(k) I" 1.897 + 2.092 126.55


,~Fe(f~)-Os2-Fe(k) ~" 1.907+ 2.107 121.00 Jkf~ 19.63

]"Fe(a)-Osz-Fe(f~) + 1.997 + 1.907 124.93 Jaf~ 18.15

{ Fe(f2)-OR3-Fe(k) ]" 1.975 + 1.928 127.88 Jf2k 4.08

]"Fe(b')-OR,-Fe(k) '[' 2.162+ 1.976 119.38


'["Fe(b")-OR~-Fe(k) 1' 2.472 + 1.976 119.38 Jbk 3.69

]"Fe(k)-OR,-Fe(k) 1" 1.976 + 1.976 97.99


~"Fe(k)-Os,-Fe(k) ~" 2.092 + 2.092 88.17
J~k <0. I
I"Fe(k)-Os2-Fe(k) ~ 2.107+ 2.107 90.08
'["Fe(k)-OR~-Fe(k) ~" 1.928 + 1.928 98.05

"["Fe(a)-Os2-Fe(k) 1" 1.995 + 2.107 95.84 J~k <0.1

{ Fe(f2~OR2-Fe(f2) $ 2.060 + 2.060 84.64 Jf2f2 <0.1


324 H. K O J I M A

and when the angle is closer to 90 ° , the p a r a m e t e r becomes negligible small. Thus,
arrows in figs. 11 and 13 etc. are the orientation of the magnetic m o m e n t s of each
ferric ion as a result of superexchange interaction. One can see in these figures
that the S block contains four Fe 3+ of up-spin in octahedral sites and two Fe 3+ of
down-spin in tetrahedral sites. In the R block, there exist three Fe 3+ of up-spin in
octahedral sites, two Fe 3÷ of down-spin in octahedral sites and one Fe 3+ of up-spin
in a trigonal bipyramidal site. Considering the magnetization values of B a M with
the Andersons (1950) indirect exchange theory, the exchange scheme of the
compound is illustrated in fig. 16, proposed by Gorter (1957), and these relations
are tabulated in table 6. Notations of the sublattices commonly used for M6ss-
bauer or N M R investigations are also shown in the last column of the table.

[20 =l14f, ~
J4f2~ 2
14f2~ b[

[4f2//~12k [[[
I
14f2,2kHt
t20--lt4',
Fig. 16. Exchange interaction scheme in the unit cell of the magnetoplumbite structure. Each arrow
represents the magnetic moment of a Fe 3+ ion (Gorter 1957).

TABLE 6
Coordination number and direction of magnetic moment of Fe 3÷ ions in the unit cell of the
magnetoplumbite type crystal.

Number of Wyckof's Direction of magnetic


Coordination number positions notation moment per mole Remarks

6 12 k I" I" I" I' '~ 1' I (a)


(Octahedral site) 4 f2 ~ ~, III (d)
2 a t v (b)

4
4 fl $ ,~ II (c)
(Tetrahedral site)

5
2 b I" IV (e)
(Trigonal bipyramidal site)
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 325

The total magnetization at temperature T, therefore, is expressed as

Ms(T) = 6O-k(T) -- 2o-f~(T) - 2o-~2(T) + O'b(T) + O'a(T), (1)

where, Ok, O-fl, O'f2, O'b and O'a denote the magnetization of one Fe 3+ ion in each
sublattice. Because Fe 3+ has the magnetic moment of 5/~B at 0 K, eq. (1) is
calculated as

Md0 K) = 5 x (6 - 2 - 2 + 1 + 1) = 20/~B. (2)

This value agrees well with the experimental result of BaM as described in the
next section.
The large magnetocrystalline anisotropy of compounds was explained by Smit
(1959) from the effect of Fe 3+ in the trigonal bipyramidal site, where a ferric ion is
surrounded by five oxygen ions. Fuchikami (1965) also showed theoretically the
contribution of Fe B+ in the trigonal site as principally responsible for the uniaxial
anisotropy. On the other hand, Van Wieringen (1967) stressed from the M6ss-
bauer investigations that the contributions to the anisotropy from all of five
sublattices should be more or less equally considered.

2.1.3. Saturation magnetization


The saturation magnetization per unit volume, Ms, per gram o-s, the number of
Bohr magnetons per mole Ns at room temperature or 0 K and the ferromagnetic
Curie temperature 0c for BaM, SrM and PbM reported by various authors are
collected in table 7. Values of NS at 0 K for these three compounds are about
20/-tB and agree well with the theoretically expected values from their magnetic
structures as given by eq. (1) or (2).
The temperature dependences of the saturation magnetization Ms near the
Curie points are shown for BaM and SrM by Shirk and Buessem (1969) in fig. 17.
Figure 18 illustrates the compariso n of Ms for BaM (Rathenau et al. 1952), SrM
(Jahn and Mfiller 1969) and PbM (Pauthenet and Rimet 1959a) in a wider
temperature range. In addition, fig. 19 and fig. 20 are other reported o-~T
relations for BaM and SrM (Shirk and Buessem 1969) and BaM (Tauber et al.
1963, Casimir et al. 1959) respectively. Belov et al. (1965) pointed out that the
spontaneous magnetization of BaM along the c-axis is somewhat greater than that
in the perpendicular direction to it, and found the difference of 0c between both
directions to be as shown in table 8. Similar facts are reported on the paramag-
netic Curie points, as described later. Though the reported results are slightly
different from author to author, the magnetic characteristics of M-type com-
pounds may be summarized as follows:
(i) The saturation magnetization becomes smaller according to the order to
SrM, BaM and PbM; (ii) The Curie temperatures also become smaller in the same
order; (iii) Saturation magnetization versus temperature curves are almost straight
or rather concave with a steep slope in the temperature range of 200 and
600 K.
326 H. KOJIMA

k~ s,
d °

(-q
d o.
+1 I +J
0 o'; ['---

+~ +1 I

t-,- t'--

8
o'3
+1 I
~ tt3
0
t~ tt3

or-- t'q
,4~d
e- p-- ,,o
0

5
'4~ eq
z

o r~
F U N D A M E N T A L P R O P E R T I E S OF H E X A G O N A L F E R R I T E S 327

120

80 --( , •

'o_
×
6O

4 0

20 . . . . . . . •

o I I T
680 700 720 740 760 780
Temperature (K)
Fig. 17. Saturation magnetization versus temperature near the Curie points of BaO.hFc_-(% and
SrO.6Fe203 (Shirk and Buessem 1969): (1) BaM, (2) SrM.

TemperQture (°C)
-200 -I00 0 IO0 200 300 400 500
600
(I)

~. 400

x 300
,,¢
x I
'-" 2 0 0 ~ + ----

ioo - - . . . . . ~ -

0
I' i
i
\ __
0 200 400 600 800
Temperafure (K)
Fig. 18. Saturation magnetization of BaO.6Fe203, SrO.6Fe203 and PbO.6Fe203 as a function of
temperature: (1) BaM (Rathenau et al. 1952), (2) SrM (Jahn and Mfiller 1969), (3) PbM (Pauthenet and
Rimer 1959a).

TABLE 8
Differences of ferromagnetic Curie temperature 0¢ of BaFe12Ot9
along and perpendicular to the c-axis (Belov et al. 1965).

Sample 0c//(°C) 00~ (°C) 0c/j-0ci

Single crystal 440.3 436.5 3.8 °

Polycrystal (oriented) 459.4 453.3 6.1 °


328 H. K O J I M A

I
\,
100
~g
<
E, 80

b
x 60
k
x

40

03

2O

0
0 200 400 600 800
Temperature (K)

Fig. 19. Plot of saturation magnetization per gram versus temperature for BaO-6Fe203 and
SrO.6Fe203 (Shirk and Buessem 1969): (1) BaM, (2) SrM.

Temperature (°C)
2O0 0 200 400
I I I

100

\
,E 80

b-

'O
x 60
b
&

4O
ID

C 20

0
0 200 400 600 800
Temperature (K)
Fig. 20. Plot of saturation magnetization per gram versus temperature for BaO.6Fe203: (1) Tauber et
al. (1963), (2) Casimir et al. (1959).
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 329

2.1.4. Magnetocrystalline anisotropy


Villers (1959a) measured the specific magnetization of a PbM single crystal as a
function of the applied field. From his results along various measured directions
against the c-axis, as shown in fig. 21, a fairly large uniaxial anisotropy can be
estimated for the crystal. In fig. 22, magnetizing curves of B a M (Casimir et al.
1959) and P b M (Pauthenet and Rimet 1959a) perpendicular to the c-axis are
c o m p a r e d at various temperatures. Magnetocrystalline anisotropy constants K~ as
a function of t e m p e r a t u r e are plotted for B a M (Rathenau et al. 1952, Shirk and
Buessem 1969), SrM (Jahn and Mfiller 1969, Shirk and Buessem 1969) and PbM
(Pauthenet and Rimet 1959a, Villers 1959b) in figs. 23, 24 and 25 respectively.
When the direction of the spontaneous magnetization in a hexagonal crystal is
expressed by polar coordinates, 0 and ~b with respect to the crystal axis, assuming
the z-axis is the c hexagonal axis, then the magnetocrystalline energy E is given
by

E = K~ sin 2 0 + / ( 2 sin 4 0 + K3 sin 6 0 + / ( 4 sin 6 0 COS 6 (D q- . . . . (3)

For M-type compounds,/£2,/(3 . . . . are negligible in comparison with K1 and their


70
I
0=0

) 5o
60
otOj,.e_e-e- -
/ - T

Fy
/
r ~

y,
b 4o
×

H
,x 3 0 vl , __.t/#
I

20 / )
v
b

,o/
0
0 5 I0 15 20
H (kOe) [ x l O 6 / ( 4 n -) A / m ]
Fig. 21. Magnetization per gram of PbO.6Fe203 single crystal as a function of the magnetic field
applied at various angles with respect to the c-axis (Villers 1959a).
330 H. K O J I M A

600

20.4 K

0 ~ C ' 0 4 2 ~ " 0 " .7 - - - ~ - - - 0 - - --0- -


0
500 ;.~,. . .77.4 K

,~ 2.3-4.2-9,5-15-20.4 K

400

b
~,l //
,~ 300
×

eoo / 7SI ~ [
• JI ! '
567 K

tOO :4

t ' ~ G-
: ~ !
0 ,-----r - ° - - ; I I i
0 5 I0 15 20 25 30
H (kOe) E x l O e / ( 4 r r ) A/m]
Fig. 22. Magnetization of BaO.6FezO3 and PbO.6Fe203 single crystals at various temperatures as a
function of the magnetic field applied perpendicular to the c-axis: ([]) B a M (Casimir et al. 1959), (O)
PbM (Pauthenet and Rimer 1959a).

Temperofure (°C)
-200 -I00 0 I00 200 500 400 500

%
4
L I I ! r F ! r L !

I
L .........
i

x
~J5
ro
E
!

/
o
L
I
~L
× T -
v
i
~-o i
0 200 400 600 800
Temperafure (K)
Fig. 23. Magnetocrystalline anisotropy constant K~ of BaO-6Fe203 as a function of temperature: (C))
R a t h e n a u et al. (1952), (Q) Shirk and Buessem (1969).
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 331

~
tO
5

-3 4
%
,x, 3 .,
t,3
E
0
"" 2

-- [
X

"-- o
0 200 400
:
600
%800
Temperafure (K)
Fig. 24. Magnetocrystalline anisotropy constant K~ of SrO.6Fe203 as a function of temperature: (O)
Jahn and Miiller (1969), (z~) Shirk and Buessem (1969).

t¢3
E4 4 E
J

%s s%
X
,< t...a

0
.-,%. 2 ,n
E
O

(2) K 2 \°Xo ~'~.~ ID

% O

x 0 0 x
0 200 400 600 800
Temperature (K)

Fig. 25. Magnetocrystalline anisotropy constant KI and K2 of PbO.6Fe203 as a function of tem-


perature: (O), (O), Kj K2, Pauthenet and Rimet (1959a); (A) K~ Villers (1959b).

values were reported only for PbM, as shown in fig. 25 and table 9. The order of
the values of Kj among the three compounds is again, the same as M~, i.e.,
SrM > BaM > PbM, though the difference between SrM and BaM is not clear in
s o m e cases.
Now, because HA defines the maximum coercivity as described in the next
section, we shall briefly compare HA values among these compounds. Since the
magnetocrystalline anisotropy field HA for uniaxial materials is given by

Ha = 2KI/Ms, (4)

we can easily estimate the value of HA from the data of K1 and Ms, described
here. Figures 26 and 27 give curves of HA against temperature for BaM (Rathenau
et al. 1952, Shirk and Buessem 1969) and SrM (Jahn and M/iller 1969, Shirk and
332 H. KOJIMA

x~ o

0
~+~ ~+
©
II II ×

dddd dd
e. e- e. ~ ~ e-

¢=

-ff
d
+l ~ +1 tt~
t-- ~.

+1 +i

Ca

e.

¢e) +1
eqeq ¢q. t'q
Me4 m

xa m.

+1 +1

ff

o
~7

m
F U N D A M E N T A L P R O P E R T I E S OF H E X A G O N A L F E R R I T E S 333

T e m p e r e f u r e (°C)
-200 -I00 0 I00 200 300 400 500
20 , ~ i I! I i I i

"- ..4
< i 5 - -

'~ I0
o
x

o 5

31
i!

0
0 260 400
i
600
l 800
Temper0ture (K)
Fig. 26. Magnetocrystalline anisotropy field HA versus temperature for BaO.6Fe203: (C)) Rathenau et
al. (1952), (A) Shirk and Buessem (1969).

25

20

<

15

......... (3) " - . . ~


\
o
×
""\ \
10
o

'1-
5

0
0 200 400 600 800
Temperature (K)
Fig. 27. Magnetocrystalline anisotropy field HA versus temperature for SrO.6Fe203 and PbO'6Fe203.
SrM: (1) Jahn and M/iller (1969), (2) Shirk and Buessem (1969). PbM: (3) calculated from the data of
Pauthenet and Rimet (1959a).
334 H. KOJIMA

Buessem 1969) respectively. The broken line in fig. 27 shows HA of PbM


calculated from the results of Pauthenet and Rimet (1959a). There are broad
maxima between 400 and 600 K, except Rathenau's results. Beside the data in figs.
23-27, the reported values of the magnetocrystalline anisotropy constant K~ and
anisotropy field Ha are tabulated in table 9. Above all, from these points of view,
SrM eventually shows the best properties as the permanent magnet material
among the three compounds.

2.1.5. ())ercive force


It can be generally said that the coercive force of ferrite magnets such as sintered
or bonded compacts of BaM, SrM or PbM powders, originates in the magnetic
behaviour of single domain particles with uniaxial magnetocrystalline anisotropy.
Now we shall briefly describe general considerations regarding the coercive force
of such materials. The critical diameter of a particle, where a single domain state
can be energetically realized, was obtained by Kittel (1946) and Ndel (1947) as

D = (9/2¢r)(O-w/m~), (5)

cr,~ in the equation denotes the domain wall energy per unit area and ap-
proximately written by

O-w= 4(AK) '/2 . (6)

Here, A expresses the exchange constant. In table I{L the values of D~, ow and
MHc are listed for M compounds.
Since only rotation of magnetization can be expected in a particle of smaller
diameter than D~, assuming its demagnetizing factor as N. we can write the
coercive force as

MHc = 2 K / M ~ - N M ~ . (7)

For the magnetizing process of a powder assembly, Stoner and Wohlfarth (1948)
assumed the coherent rotation mode without any interaction effect, and concluded
that the coercivity factor obtained was 0.48 for the randomly oriented powders;
then MHc is given by

MHc = 0.48(2K/Ms - NMs) . (8)

For M compounds, crystals are usually easy to grow along the c plane, so that N
affects MHc more unfavourably than spherical particles. However, we still cannot
attain the MHc value of eq. (8) experimentally for practical hexagonal ferrite
powders. To explain this, various magnetization reversal mechanisms of incoherent
type have been proposed.
For example, by considering the interaction field of surrounding magnetic dipoles,
Eldridge (1961) simply estimated the effect of mutual interaction. Hence, we can
F U N D A M E N T A L P R O P E R T I E S OF H E X A G O N A L F E R R I T E S 335

e~

8 io< ©
,=
$
!c X
kD
, t¢3 I [,q kD
;o 5
II I II II

O O
.o O 0
,,=
N 'i
1 I
a e~
O
"a
! e~ D.
ff )
o 6
L) )
6
L) ~d
N

8
11 ©
©
©
e~

0 0
~.~ ¢..)

m. o
# trl 6
&-a x
~'~

m O~ uq ~ ~D t~q
o. tt3
,4

o X

d
©
o.
e~ O0 +1
¢- t'4

¢×
E
O

e. o,
~ E g =N
o

o
336 H. K O J I M A

generally express MHc using his results:

MHc = MHc(0)- CpMs, with C = 1.7. (9)

Here MHc(0) and p express the coercive force of ideal coherent rotation and the
volumetric packing factor of the powder assembly. Bottoni et al. (1972) deduced
the partial existence of chains of spheres with fanning mode from the rotational
hysteresis measurements for dry-milled BaM powders. Many researchers, for
example, Hoselitz and Nolan (1969), Ratnum and Buessem (1972), Haneda and
Kojima (1973a, 1974) besides others mentioned below, have been discussing the
causes for the reduction of domain nucleation field and they concluded that
domain wall movement should be also considered for the magnetizing process of
M compound powders, even though they are sufficiently small compared to Dc in
table 10. For instance, lattice defects (Heimke 1962, 1963), different phases
(Richter and Dietrich 1968), stacking faults or deformation twins (Rantnam and
Buessem 1970), local changes in anisotropy (Aharoni 1962), inhomogeneity of
magnetic field in the edge of plate like crystals (Holtz 1970) etc. can be given as
factors to be considered as possible causes.
Experimental results of packing effects on MHc are illustrated in fig. 28 for BaM
(1), (2) by Shimizu and Fukami (1972), BaM (3) by Luborsky (1966) and SrM (4)
by Hagner and Heinecke (1974). We can say from these results that the constant
C is rather smaller than the value of eq. (9) for BaM or SrM, and this may be
caused by their larger Ks and lower Ms values. The last authors especially stressed
the importance of the agglomeration effect to verify the dependence of MHc on
the packing density. To prevent the agglomeration, they used a soft ferrite to
dilute SrM powders and measured MHc above the Curie temperature of the soft

4000

< 3000
g

_o 2 0 0 0 (4)
x

o
I000
o
"1"

0
0 0.2 0.4 0.6 0.8 1.0
Packing factor P
Fig. 28. Effect of the packing factor on the coercive force of BaO.6Fe203 and SrO'6Fe203. (1), (2)
BaM (Shimizu and Fukami 1972), (3) BaM (Luborsky 1966), (4) SrM (Hagner and Heinecke 1974).
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 337

ferrite. But Tfmfisoiu et al. (1976) reported a very high MHc value of 6500 Oe
[5.17x 105A/m], obtained for SrM by the ordinary dispersion method in zir-
conium oxide powder or an epoxy resin, if the packing factor p ranges from 0.1 to
0.01.
Curve (1) in fig. 29 is an example of the change in MHc during the milling
process (after Haneda and Kojima (1974)). Though the shape of the curve
depends on the grinding conditions, MHc generally increases at first and then
decreases with milling time. The first rise is presumably caused by the approach to
the single domain behaviour and the following decrease may be explained b y
various causes introduced by the prolonged milling process, as mentioned before.
Curve (2) in the figure shows MHc obtained by annealing the specimenS of curve
(1) at 1000°C for an hour. The remarkable increase may be mainly due to the
recovery of various defects. The MHc versus particle diameter relations of BaM
are compared in fig. 30. Shirk and Buessem (1971) determined the particle size by
X-ray (1) and electron microscope (2), while Sixtus et al. (1956) measured MHc
against thickness (3) and diameter (4) of the platelets. Powders larger than 10 .4 c m
seem to behave as multidomain particles and below 2 x 10 6cm by X-ray or
6 x 10-6cm by electron microscope, superparamagnetic effects are likely to be
predominant. Since superparamagnetic powders do not show a hysteresis loop,
MHc abruptly tends to zero in this region. If the particle has volume V and Sc
denotes the superparamagnetic critical size, the coercivity factor against V/Sc can
be expressed as in fig. 31, according to the calculated results of Shirk and Buessem
(1971) for the case of BaM. Hence, one can say the Stoner-Wohlfarth model is
valid only under the condition of V ~ 100So, that is roughly larger than 4 x 10-6 cm
in diameter for a BaM particle; the agreement of this estimation with the results
in fig. 30 is reasonably good.

k II
1

0 ~' / --e~e'-- •

0
0 50 100 150 200
Milling time (hr)
Fig. 29. Milling and annealing effect on the coercive force of BaO.6Fe203 (Haneda and Kojima 1974),
(1) as milled powder, (2) after annealing at 1000°C for 1 h.
338 H. KOJIMA

104 I I I I I I I I I I I I I I I I I
i
I

< i0 3 o

O
x

I) (2)
O
i0 2 x\

I0 I I I I I I II I I I I I I [ I I
10-6 10-5 iO-4 IO-3 fO-Z
Porficle diomefer (cm)

Fig, 30. Relation of coercive force and particle size in BaO.6Fe203: (1) diameter as spheres deter-
mined by X-ray, (2) by electron microscope (Shirk and Buessem 1971), (3) thickness of platelets
determined by optical microscope, (4) diameter (Sixtus et al. 1956).

Figure 32 shows the changes of MHc with temperature; i.e., for SrM (1) and
BaM (2) by Mee and Jeschke (1963), BaM (3) by Sixtus et al. (1956), BaM (4) by
Rathenau (1953) and SrM (5) by Jahn and Mdller (1969). All results show maxima
between 200 ° and 300°C. Rathenau tried to explain the maximum of MHc by
assuming a single domain state around 300°C and a multidomain state at room

0.6 i i E ~ i i i E i i i

0.5

0.4
8
0.3

$ o.2
3
0.1

0 E i I I I I I I
I I0 10 2 03
V/Sc
Fig, 31. Coercivity factor against particle size for isotropic powder assemblies (Shirk and Buessem
1971). V: volume of a particle, So: super-paramagnetic critical size; dotted line: Stoner-Wohlfarth
model.
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 339

6 .- o~ .---o l

?-
< 5~ ~" °'~ ''°

! i

-200
i 0 200
I 400
Temperature (°C)
Fig. 32. Temperature dependence of the coercive force for BaO-6Fe2Os and SrO-6Fe203. (1) BaM, (2)
SrM (Mee and Jeschke 1963), (3) BaM (Sixtus et al. 1956), (4) BaM (Rathenau 1953), (5) SrM (Jahn
and M/iller 1969).

9 I r i
i
i
8
i

s j
J

5
-200 0 200 400
Temperature (*C)
Fig. 33. Temperature dependence of ",/~/Ms, which is proportional to the critical diameter of single
domain particle (Rathenau 1953).
340 H. K O J I M A

temperature. He reported the temperature dependence of (KO1/2/Msas in fig. 33,


to which the critical particle diameter Do would be proportional. His explanation
is based on the increase of this quantity with temperature. However, this might
only be one reason, because many authors reported maxima of HA in the same
temperature range as MHc (see figs. 26 and 27).
Craik and Hill (1977) pointed out that the strong domain wall pinning could be
expected at grain boundaries of BaM powders without supposing any imper-
fections and so the role of grain boundaries is important in relation to the
coercivity mechanisms. Thus, the magnetization reversal process in hexagonal
oxide powders are still open to investigation for some details, even for the
particles of critical size. But we may now conclude that all the factors described
here should be considered, though the dominant magnetizing mechanism depends
perhaps on the preparation method.
When the magnetization reversal is achieved by the nucleation and growth of
reverse domain, the angular variation of coercive force, plotted as MHc versus
1/cos 0, shows a linear relation. Here, 0 expresses the angle between the applied
field and the easy axis of the oriented powder assembly. Ratnam and Buessem
(1972) obtained this relation with BaM particles of 100-200 Ixm as shown in fig.
34. However, if particles perfectly follow the Stoner-Wohlfarth model then MHc
can be drawn as the curve S-W in fig. 35; here the value of the effective
anisotropy field HA is assumed to be 12000 Oe [9.5 × 105 A/m], considering the
demagnetizing factor. For practical powders, magnetizing process takes place by
both of these mechanisms and thus curve (2) is observed as the angular depen-
dence of ordinary ball-milled BaM powders. For the sample prepared under
relatively defect-free conditions, for instance by the chemical co-precipitation

240
o

E 200
<
f
F-
~O
--
×
160
/

×
120 /
O
IO
/
v
u
"r
×

40 /
o
. o 0.2 0,4 0,6 0,8 1.0

cos 8
Fig. 34. A n g u l a r dependence of the coercive force for oriented BaO.6Fe203 powders (Ratnam and
B u e s s e m /972).
F U N D A M E N T A L PROPERTIES OF H E X A G O N A L FERRITES 341

12 I /I
~lo i
\ /
K//

tt S-W /

2 , ~ Q O ~ Q ~ O ---''-tD'-''~" ~ ~ ,
I
(2)
0
0 0 40 60 80
0 (degree}
Fig. 35. Analysis of magnetizing process in BaO.6Fe;O; powders from the angular variation of
coercivity (Haneda and Kojima 1973a): (1) co-precipitated powders, (2) ball-milled powders, (S-W)
calculated curve by the coherent rotation model, (K) calculated curve by the magnetizing process with
wall motion.

method, they may be magnetized mostly by coherent rotation, as proved by curve


(1) in fig. 35, where it completely agrees with curve S - W in the range 4 0 ° < 0 <
90 ° . To see the difference more clearly, magnetizing curves are compared with the
theoretical S - W curve in fig. 36. Demagnetizing curves of defect-less powders of
BaM (1) and SrM (3) agree well with the calculated curves for coherent rotation,
S-W, but ball-milled BaM powders (2) behave rather differently from the S - W
model (Haneda and Kojima 1973a, Tfinfisoiu et al. 1976). Ratnam and Buessem

i// I f

id " /'"'1 I 5 <koe~

.g ," I -- ,/i~

Fig. 36. Demagnetizing curves of unoriented powder assemblies: (1) co-precipitated BaM, (2) ball-
milled BaM (Haneda and Kojima 1973a), (3) wet-milled and annealed SrM (Tfinfisoiu et al. 1976).
342 H. KOJIMA

(1972) and Haneda and Kojima (1973a) proposed methods to estimate the ratio of
the two magnetizing mechanisms in a powder assembly from the angular variation
of MHc. The highest MHc values obtained hitherto by special preparation methods
are listed, together with other experimental conditions, in table 10.

2.1.6. Paramagnetic properties


Borovik and Mamaluy (1963) measured the temperature dependence of the
susceptibility per gram above the Curie point for M compounds. Figure 37 shows
the 1/xg-r relations of BaM (1) (Gorter 1954), SrM (2) and PbM (3) (Borovik and
Mamaluy 1963). They also determined N6el's constants for the susceptibility per
atom XA; ~r, CA and X0,A in eq. (10) (N6el 1948) by a graphical method from the
results of fig. 37. Here,

1 1 T o-
- ~ (10)
XA Xo,A CA T-0'

and 0 can be given by

o = nCAA~ (2 + o~+/?), (11)

where CA is the atomic Curie constant for Fe3+; A,/x are the numbers of the Fe 3+
ions in the A and B sublattice; n, a and /3 are the molecular field constants
connected with the exchange interactions of the AB, A A and BB types. If the

T (°C)
600 800 IO00
8 I I I iI i ] ~ t , ~I
I

,~. .o/e
{,9
~. I ..
6 [/I/ l

~s
x
4

3
~o
x- 2 1
/
I
1
I
0
700 900
II00 1300
T{K)
Fig, 37. Temperature dependence of the specific susceptibility above the Curie temperature: (1) BaM
(Gorter 1954), (2) SrM, (3) PbM (Borovik and Mamaluy 1963).
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 343

m o l a r susceptibility XM is used i n s t e a d of XA in eq. (10), t h e n CM can b e o b t a i n e d


i n s t e a d of CA. T h e c o n s t a n t s are t a b u l a t e d in t a b l e 11. B o r o v i k a n d M a m a l u y
c o n c l u d e d that the c o n s t a n t s X0,A, Cr, CA, a n d t h e r e f o r e the m o l e c u l a r field
c o n s t a n t s are r o u g h l y the s a m e for h e x a g o n a l a n d cubic ferrites.
A l d o n a r d et al. (1966) m e a s u r e d XM of P b M a n d f o u n d different values for AiM
with a p p l i e d m a g n e t i c field parallel a n d p e r p e n d i c u l a r to the c-axis, as s h o w n in
fig. 38. T h e y also d e r i v e d t h e following f o r m u l a from the precise m e a s u r e m e n t of
the t e m p e r a t u r e c h a n g e in XM in cgs u n i t b e t w e e n 800 a n d 1200 K along the
c-axis:

1 1 T 4367
XM 38.21 + 6 5 T-655" (12)

10 I 2
o H II C aXiS I .,//
9 fd-
• H .L c oxis / J/"

~7 ! Az

-~5

~3

"2_ I
///1
0 ////
710 720 7:30 740 750
T(K)
Fig. 38. Temperature dependence of the molar susceptibility of PbM along parallel and perpendicular
directions to the c-axis (Al6onard et al. 1966).

TABLE 11
N6el constants of M compounds according to eq. (10) (Borovik and Mamaluy 1963).

CA CM 104 (emu/atom)
l/X0, A X o- x 105
Compound cm3K/atom cm3K/mol [1/(4zr) × 10 4 S I / a t o m ] K atom/cm3

BaM 0.0515 58 5.1 26


SrM 0.0475 51 5.8 27
PbM 0.0526 61 6.8 31

CA: atomic Curie constant of Fe3+ ions,


CM: Curie constant obtained from molar susceptibility.
344 H. K O J I M A

2.1.7. Magnetic aftereffect


Diilken et al. (1969) investigated the relaxation phenomena of polycrystal BaM.
They measured the complex permeability between 77 and 743 K in the frequency
range from 500 Hz to 100 kHz. The real p a r t / x ' and the imaginary part/x" of the
initial permeability as a function of temperature T, measured at i kHz for Fe 2+
free BaM, show linear relations with T, as shown in fig. 39(a). However, /x" of
BaM with 0.24wt% Fe 2+ measured at 10kHz shows three maxima at - 1 8 5 °,
- 1 4 5 ° and +60°C, as shown in fig. 39(b). If k and Q denote Boltzmann's constant
and the appropriate activation energy, then the relaxation time ~- is expressed as,

r = "r= exp(O/kT). (13)

Diilken et al. (1969) determined O for the three maxima as (I) 0.1 eV, (II) 0.2 eV
and (III) 0.6 eV, and the relaxation frequency f = 1/2~-~-= as (I), (II) - 101° Hz and
( I I I ) ~ 1 0 1 3 Hz. These relaxation phenomena can be explained by the diffusion
process between Fe 2+ and Fe 3+, which is often observed in cubic ferrites.

5 I I
--o-c~-c. -o-o-oc - o - o -
F'

10°5

t
h.
Io-) l
5

=._o_oj °-%
F" I t0 -2
(b)
I

!.c-.o--~ ~ "ov---o-c-~-.o~

I 1°°5 F' 510-1l


(a)
lo-' lo-2
5 o'o--c-~ ~c.o.o~5
-200 -I00 0 I00
Temperature (°C)

Fig. 39. Temperature variation of the complex permeabi|ity for polycrystal BaO.6Fe203 (Dtinke] ¢t al.
1969): (a) Fe z+ free B a M at i kHz, (b) B a M with 0.24 w t % Fe 2+ at 10 kHz.
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 345

2.1.8. F M R
Numerous experiments of ferromagnetic resonance absorption in BaM or SrM
have been reported, most of which aim to determine the anisotropy field H A and
the effective spectroscopic splitting factor gen against temperature. Several authors
also especially investigated the effects by the shape, size or surface roughness of
the samples, the existence of domain walls, or the angle between the applied d.c.
field and the easy axis, etc. Figure 40 shows the resonance absorption in BaM at
various temperatures. Smit and Beliers (1955) showed with a thin single crystal
that only the absorption peak excited by a parallel a.c. field is observed at lower
temperature, but the perpendicular field peak appears above room temperature
and splits into two at 200°C (fig. 40(a)). Grosser (1970) investigated the resonance
absorption of an isotropic polycrystal between 21°C (fres= 22.22 GHz) and 465°C
@es = 22.03 GHz) as shown in fig. 40(b), where the resonance field at room
temperature moves to lower fields with increasing temperature near the Curie

2.5 \
\\
A
\
b) I
t3
2.0 \
(3 \ 450* 420*
03 \
..Q \
1.5 \

1.0
Q.
o 4ti~ .- ~/
J~ 0.5

0o 4 8 12 16 20
H (kOe) [xlOS/(4~-) A/m]
50 I
(a) 200"C
a~

o
40 IP
to
..a
30
II 11

g
,/i
I I I

"..7-- 20 I/ '
¢.~

0 IJ
o~
t~
< I0

0
I 13 15 17 19
H (kOe) [x106/(4~) A/m]
Fig. 40. Ferromagnetic resonance absorption in BaO.6Fe203 versus applied d.c. field at various
temperatures: (a) single crystal (Smit and Beljers 1955), (b) isotropic polycrystal (Grosser 1970).
346 H. KOJIMA

points. The ordinate of the figure is illustrated in arbitrary scale but it is


proportional to/x".
Figure 41 is the relation of resonance frequency and applied d.c. field for a BaM
single-crystal, by Silber et al. (1967). The solid lines in the figure express the
calculated results for various angles 00 between the c-axis and the applied field,
using the following resonance conditions:

(o~/y)2 = [Ha(1 - 2 sin 2 0) + Hext cos(0o - 0)]Hext(sin 0o/sin 0), (14)

and

sin 0 cos 0/sin(00- 0) = HexdHa. (15)

Here, ~o, y and 0 represent the resonance frequency, gyromagnetic ratio and
angle between the c-axis and the magnetization vector, respectively. The effective
anisotropy field is generally given by Ha = HA--(NIl- N±)M~. Thus, Silber et al.
determined y from the slope and HA from the intersection with the frequency axis
of the zero-degree curve in fig. 41. The "knees" in the experimental data

0o 44* 59*

ID

°.o .........--- . i'


rv

46

42
0 4 8 2 16 20
Applied field (k0e) "xl06/(4=) A/m]
Fig. 41. Relation of resonance frequency and magnetic field applied at various angles to the c-axis of
BaO.6Fe203 single crystal (Silber et al. 1967).
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 347

represent the change from the single domain state at higher fields to the
multidomain state at lower fields. The resonance frequencies and absorption
intensities in BaM specimens with cylindrical domain structure have been cal-
culated by Sigal (1977) and compared with the experimental results, changing the
direction of magnetic field with the easy axis. Typical experimental results are
summarized in table 12.
When the d.c. field is cycled between _-_25kOe [+_1.99 x 106A/m] for BaM
milled powder, Hempel and Kmitta (1971) found a microwave absorption peak at
about 2 kOe [1.59 x 105 A/m] with f - - 9.4 GHz, which shows a pronounced hys-
teresis apart from the F M R occurring at about 15 kOe [1.19 x 106 A/m]. They also
observed that the low field losses and the line width of the F M R decrease with
increasing annealing temperature of the sample. While Roos et al. (1977) found
microwave absorption of chemically co-precipitated, polycrystalline BaM at H
--HAl2, when the sample was previously saturated; however, no absorption peak
in negative fields was found for the conventionally prepared BaM. They obtained
three peaks at H = -6.6 kOe [-5.25 x 105 A/m], -8.8 kOe [-7.00 x 105 A/m] and
-15.3 kOe [-1.22 × 106 A/m] with f = 11.9 GHz, and concluded from these results
that the magnetization in most particles of the chemically precipitated sample is
reversed by coherent rotation.
Besides such basic investigations, several studies relevant to the applications to
microwave devices have been reported. For example, Dixon Jr. and Weiner (1970)
obtained H a = 17-5 kOe [13.5-4.0 x 105 A/m] over the frequency range of f =
23.2-37.6 GHz with BaZnxTixFe12 2xO19. De Bitetto (1964) also investigated the
F M R of BaO.x(TiCoO3)-(6-x)Fe203 and SrO-xA1203.(6-x)Fe203 systems and
determined HA as 17.5-6.6 kOe and 19.3--53.4 kOe [15.442.5 x 105 A/m] respec-
tively.

2.1.9. N M R
Streever (1969) obtained the NMR line shapes of Fe 57 in single-crystal powders at
4.2K and resintered single-domain powder at 77 K for BaM, by plotting the
spin-echo amplitudes versus frequency as shown in figs. 42(a) and (b). Each line
can be assigned on the basis of the changes in relative intensities and shift in
external magnetic fields. M6ssbauer data are also useful to analyze the signals and
the reverse is naturally also true. The relations of the sublattices and the notations
in the figure are identical to those used in table 5. L/itgemeier et al. (1977)
measured relaxation rates of Fe 57 nuclei in BaM between 1.2 and 4.2K and
proved that the relaxation time is much shorter for the wall signals than for the
domain signals due to the influence of the wall excitations. The temperature
dependence of NMR frequencies is summarized in fig. 43, after the results of
Streever and Hareyama et al. (1970). Though the temperature changes of signal
III (sublattice d) and V (b) show some differences between the two papers, the
frequency changes of each signal are almost continuous over the whole tem-
perature range. The temperature dependence of signal I (a) from Fe 3+ in octa-
hedral a sites is more remarkable than those of signal II (c), III (d), IV (e) and V
(b), while signal IV (e) appears at a lower frequency than the others. These results
348 H. K O J I M A

T A B L E 12
Summarized results of F M R experiments for M-type compounds.

Compound Experimental condition fr~s(GHz) T(°C) ge~

single cryst, disc, d = 1 m m , 23.93 -196 2.00 (assumed)


t = 40 tzm; disc _1_c-axis, 23.98 ' 20 1.98
BaM Hr~s and ha in basal plane, 23.98 112 2.00
angle between Hres and ha 23.98 155 2.01
is 45 ° 23.98 200 2.02

single cryst, disc, d = 0.083 cm, 58.2 R.T. 1.87


t = 0.030 cm; Hr~s//disc, 58.4
BaM Hre~//c-axis, hr~A_H ~ 58.6
58.8
59.0

BaM single cryst, sphere, d = 0.4 m m 53.0-55.3 25 1.96


Hresffc-axis, h~//c-axis

BaM 1.91
SrM oriented polycrist, slab R.T. 1.91

BaM single cryst, disc, 20 2.05


d = 2.60 cm, t = 0.01 cm

single cryst, sphere, 51-65


BaM d = 0.036 cm R.T. 2.02 ± 0.01

BaM single cryst, sphere, 66.6, 120 4.2 (K) 1.99 - 0.03
d = 0.05 cm 66.6, 120 20 (°C)

single cryst, sphere, d = 0.038~.05 cm,


BaM Hre~~~c-axis I>52.0 R.T.

22 21
200
BaM 300
35O
40O

* Spin wave line width


(a) Smit and Beliers (1955) (f) Silber et al. (1967)
(b) W a n g et al. (1961) (g) Kurtin (1969)
(c) Mita (1963) (h) Dixon and Weiner (1970)
(d) D e Bitetto (1964) (i) Grosser (1970)
(e) Burlier (1962)
F U N D A M E N T A L P R O P E R T I E S OF H E X A G O N A L F E R R I T E S 349

T A B L E 12 (continued)

Hr~ (kOe) [x 106/(4~-)A/ml


HA (kOe) A H (Oe)
hrr//'Hres hrr-l-Hres [X 106/(47r)A/m] [x llY/(4~)A/m] Remarks Ref.

14.3 -10 16.2 47rMs = 6.67 (kG)


15.05 17.15 17.0 4.80
15.40 17.55 17.3 3.90 (a)
15.40 17.15 17.3 3.50
15.35 16.85 17.3 3.12

3.844 18.4 0-13 p = 5.13 (g/cm3)


3.886 3'4 = 2.62 (MHz/G)
3.964 (b)
4.042
4.140

18.7-28.2 17.55 <12 (c)

17.5 1600
19.3 1600 (d)

17.0 3' = 2.87 (MHz/G) (e)

2.0-7.0
16.2 (f)

16.1 ~<10 4~'M~ = 5.2 (kO)


17.1 46 AHk* = 2 ± 1.5 (Oe) (g)

17.0 4~'Ms = 4.8 (kG) (h)

17.0 Ms = 375 (G)


17.0 248
15.1 177 (i)
13.1 142
10.5 106
350 H. KOJIMA

I0
(b) l
I
I

m
=
o_
E

:>
4

I0
i
/,
(a) ~
8

6 III
lI

0
/% /
58 62 66 70 74 78
Frequency (MHz)

Fig. 42. Plot of the spin-echo amplitudes versus resonance frequency for BaO.6Fe203 (Streever 1969):
(a) single-crystal powders at 4.2 K, (b) resintered single-domain powder at 77 K.

78
......... ~ _ ~ i__~__ .~
74
O. . . . . . .

I 70
2; I "7

66
I° '~ ~.,~

g62 r

b_

58 ~

]7
54
0
I
I00 200
°""°~'
500
Temperofure ( K )
Fig. 43. Temperature dependence of NMR frequency: (O) Streever (1969); (O) Hareyama et al.
(1970).
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 351

can be understood from the differences of F e - O - F e bonds for each Fe 3+ ions (see
table 5).

2.1.10. M6ssbauer effect


Van Loef and Van Groenou (1964) investigated the M6ssbauer effect with BaM
single crystals and oriented crystallites enriched in Fe 57. Changing the y-ray
direction parallel or perpendicular to the c-axis and with or without the magnetic
field of 13 kOe [10.3 x 105 A/m], they observed spectra of I + V (sublattice a + b),
II (c), III (d) and IV (e); (note that different labelling methods are used by other
investigators). The temperature dependence of the sublattice magnetization,
anisotropy and dipole fields, and the exchange constant of the Bloch wall as a
function of relative magnetization are discussed in relation with the experimental
results. Furthermore, in contrast to other authors, an especially high hyperfine
field of the sublattice e was reported by Van Loef et al. (1964). Zinn et al. (1964)
also observed M6ssbauer spectra in BaM and explained the temperature depen-
dence of the saturation magnetization by the temperature changes of internal
fields for each sublattice, although they considered only two kinds of sublattices as
a whole. Kreber et al. (1975) did further M6ssbauer study with BaM, an oriented
and enriched polycrystal between 4.2 and 200 K, and nonenriched fine powder at
300 and 870 K. They resolved the spectra as I (sublattice a), II + I I I + V (c + d + b)
and IV (e) at low temperature, and I (a), I I I + V (d + b), II (c) and IV (e) above
200 K, respectively. Based on the observed large quadrupole splitting of sublattice
e and its distinct change near 80 K, they proposed a model in which Fe 3+ occupies
randomly one of the two equivalent sites of Wyckhoff's notation 4e instead of 2b
at low temperature. Here, 4e sites are located on the trigonal axis and 0.156A
away from 2b site on the mirror plane (see also section 2.1.1 on crystal structure).
Using their model, they showed that the temperature dependence of the splitting
values, which they had observed, and also the anisotropy behaviour obtained by
the M6ssbauer data of Rensen and van Wieringen (1969) can be understood by
the molecular orbital calculation after Trautwein et al. (1975).
As for SrM, Van Wieringen and Rensen (1966) observed spectra for powders,
enriched in Fe 57 at room temperature. Four subspectra are assigned as I (sublat-
tice a), II + V (c + b), III (d) and IV (e). Internal field Hi, quadrupole splitting
and isomer shift 6 of four subspectra as a function of temperature are illustrated
in figs. 44(a), (b), (c) and 45. Van Wieringen (1967) reviewed M6ssbauer data of M
compounds and derived the temperature dependence of saturation magnetization
from the values of Hi in fig. 44(a). Figure 44(d) shows the result together with
directly measured values, expressed with the open circles, which fit quite satis-
factorily with the deduced curve. Comparing the results of BaM by Zinn et al.
(1964) or Van Loef and Franssen (1963) with the results by Van Wieringen and
Rensen (1966) it can be concluded that the M6ssbauer data of SrM are qualita-
tively the same as those of BaM, but the main quantitative difference concern
rather lower values of Hi and e in SrM. M6ssbauer effect measurements of PbM
single and polycrystals between 15 and 780 K were reported by Zinn et al. (1964).
Differing from BaM, an intense spectral line at v -- 0, arising from the paramag-
352 H. KOJIMA

120 ;' I
iI (d)
E "'---o l
90
'
II
'~o 6 0 - -

~.~. 3o

(n

E 0
E

60 -I

-j (b)

E
E !

uo

6OO
(o)
,---1
500

400 ....... ~---~r-- - - ~ .. "~"

0
3OO

o 200

z I00

0
0 200 400 600 800
T (K)
Fig. 44. Variations of the internal field (a), quadrupole splitting (b), isomer shift (c) and saturation
magnetization (d) of SrO.6Fe203 against temperature (Van Wieringen and Rensen 1966, Van
Wieringen 1967).
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 353

f'
6 . 0 x 10 5

5.0

I J

I
I I[(c)
I E(d)
I
IV(e)
I I
V(b)
I 4 I J I I I I I I I i I
-I0 -5 0 5 I0
v ( mm/sec
Fig. 45. Assignment of the M6ssbauer spectrum of SrO.6Fe203 at room temperature (Van Wieringen
and Rensen 1966, Van Wieringen 1967).

netic phase was observed in PbO.xFe203 (x = 4-6) near the Curie temperature.
A signal corresponding to the paramagnetic phase of 30% was obtained at 700 K,
though further investigation regarding the effect of the PbO.2Fe203 paramagnetic
phase in the sample seems to be necessary to explain these phenomena.
As described hitherto, two kinds of labelling for the M/Sssbauer absorption
peaks of M compounds seem to be used. That is, Kreber and Gonser (1973)
considered one of the lines V (b) with III (d) as an unresolved one, while Van
Wieringen and Rensen (1966) recognized peak V (b) overlapped with peak II (c).
Since the N M R data at low temperature by Streever (1969) and at higher
temperature by Hareyama et al. (1970) proved that the hyperfine fields and their
temperature dependence of III (d) and V (b) are almost the same as shown in fig. 43,
the assignment of Kreber and Gonser is suggested to be reasonable. However, when
we compare the results of N M R and M6ssbauer measurements, we should, of course,
realize that the spectra observed by these two experimental procedures are based on
different physical origins. In conclusion, Van Wieringen (1967) summarized the
results of M6ssbauer studies with M compounds as follows:
(1) The iron ions are trivalent in all sublattices; (2) All five sublattices may be
expected to contribute to the crystal anisotropy; (3) The rapid drop in the
magnetization with increasing temperature is entirely due to Fe 3+ in sublattice a
(signal I).
The interest regarding the M6ssbauer studies with M compounds also concerns
the determination of ion distributions in the substituted compounds as discussed
in detail below. For example, Rensen et al. (1971) investigated A1, Cr, ZnTi,
354 H KOJIMA

ZnGe, ZnSn, ZnZr, CuTi, CoTi, CoCr and NiTi substituted M compounds and
found that AI firstly enters the 2a site (sublattice b) and then the 12k site (a); Cr
also starts by entering the 2a site. In addition, they concluded that there is no
indication of Fe 3+ ions in an asymmetric 4e site at low temperature, contrary to
the results of Towns et al. (1967) or Kreber et al. (1975). Kreber and Gonser
(1973, 1976) reported that As ions occupy preferentially the 2b site (sublattice e),
while Ti 4+ + Co 2+ firstly occupy 4fl (c) o r 4f 2 (d) sites and then randomly the 2a (b),
12k (a), and lastly 2b site (e). If Ba 2+ ion in BaM is replaced with a trivalent ion,
Fe 3+ may partly change its valency t o Fe z+ to keep the neutrality. LaFe12Oa9 was
investigated from this point of view by Drofenik et al. (1973) and also by Van
Diepen and Lotgering (1974). In both papers the authors described the difficulty
in observing the Fe z+ sublattice, but the latter authors discussed the preference of
Fe z+ ion from the intensity ratio of subspectra. Moreover, Drofenik et al. (1973)
determined the Curie point as 697 K from the appearance of the paramagnetic
Mrssbauer absorption peak. The hyperfine field at Fe 57 nuclei in different sublat-
tices in LaFe12019 as a function of temperature are similar to BaM or SrM but the
change in the 12K sublattice (I) is found to be more convex than that of SrM as in
fig. 46. This agrees with the more convex o-s-T curve of LaM as compared to that
of SrM.

600

<

"~ 4 0 0

0
×
\\,,,
O
-~ 200 \\ g
\\ g
\\ g
"7-

It
II
0 ',
0 200 400 600 800
T (K)
F i g 46 Temperature dependence of hyperfine field in LaO 6Fe203 (Drofenik et al. 1973)

2.1.11. D o m a i n observation
In the earlier days, domain structures were mainly discussed qualitatively in
connection with the magnetic properties of the materials, magnetic anisotropy,
magnetizing process and so forth. However, as the sophistication of techniques in
domain observation and sample preparation improved, more quantitative in-
vestigations were attempted. Further, some of the results acquired in this research
F U N D A M E N T A L P R O P E R T I E S OF H E X A G O N A L FERR1TES 355

field have led to the development of the industrial application of bubble domains.
Kooy and Enz (1960), for example, observed the plate-like domains in thin layers
of BaM, which were peeled off from the surface of single crystals. They discussed
the results of the experimental domain period under the influence of magnetic
field and compared them with theoretically calculated results. Specific wall energy
and critical diameter of cylindrical domain were also reported in the paper as
~rw= 2.7erg/cm 2 [2.7x 10-3j/m 2] at 50°C and Dc = 0.3 Ixm for a 3 Ixm platlet.
Figure 47 illustrates the domain period versus the magnetic field applied parallel
to the easy axis for (a) BaM (Kooy and Enz 1960), (b) SrO-4.5Fe203-1.5A1203 and
(c) SrO-4.2Fe203.1.8A1203 (Rosenberg et al. 1967). The curve for BaM is the

300

(c)
250 ] 1 250

(b) I

_jl
200 '~ 2 0 0
E ii
o
L
~o 150
x
J 2 ~5o

c~ I 0 0 + I00 ~
.4- I c:T
c~
5O 50

0 0
0 0.5 1.0 1.5 0 200 400
H (kOe) [xlOS/(4"rr) A / m ] H (Oe) [xlO3/(4rr) A/m]
12

I0

'-' 8
10
x
6

+ 4
cE

0 J J 1__
0 I 2 3 4
H (kOe) [x106/(4"n") A/m]
Fig. 47. Domain period as a function of the applied field: (a) BaM, thickness: 3 x 10 .4 cm (Kooy and
Enz 1960); (b) SrO'4.5Fe203'l.SA1203; (c) SrO-4.2Fe203.1.8A1203 (Rosenberg et al. 1967). D1 + D2:
width of antiparallel two domains.
356 H. KOJIMA

result of t h e t h e o r e t i c a l curve fitting. ( U n f o r t u n a t e l y , t h e crystal t h i c k n e s s is not


c l e a r for (b) o r (c), b u t t h e s a m e a u t h o r s (1966) r e p o r t e d s e p a r a t e l y t h e D - t
r e l a t i o n s of t h e s a m e c o m p o u n d s ; as d e s c r i b e d later.) F i g u r e 48 shows t h e c h a n g e s
of d o m a i n structures on t h e b a s a l p l a n e of B a M p l a t e l e t b y t h e F a r a d a y effect
( K o j i m a a n d GotO 1965). C y l i n d r i c a l d o m a i n s a r e c l e a r l y seen in t h e h i g h e r field,
just b e f o r e t h e s a t u r a t e d state, as s h o w n in (d) a n d (e).

Fig. 48. Changes of domain structures observed by Faraday effect on the basal plane of BaO-6Fe203
(Kojima and Got6 1965): (a) demagnetized state, (b), (c) H = 3600 Oe [2.86 × 105 A/m] parallel to the
c-axis, (d), (e) H = 3900 Oe [3.09 x 105 A/m] parallel to the c-axis. 0: set angle of the analyzer from the
crossed nicosol state.
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 357

Moreover, it was proved that maze, parallel or honeycomb domains can be


realized as a remanent structure for B a M (Kojima and G o t 6 1965) and for P b M
(Kacz6r and G e m p e r l e 1961, Palatnik et al. 1975) and for BaM, SrM and P b M
(Kojima and G o t 0 1970). H e n c e a specified domain structure among these three
can be brought into existence at will by choosing certain angle ranges of applied
magnetizing field direction against the c-axis before the remanent state is
produced in the crystal. Utilizing these facts, HA and K1 were determined for
small single platelets by Kojima and G o t 6 (1964) as shown in table 9(d). The
effects of various kinds of defects on the domain nucleation process were
discussed by Kojima and G o t 0 (1965), Wells and R a t n a m (1971) for BaM, and
Tfinfisoiu (1972) for A1 substituted SrM, Sr(Fe7.2A14.8)O19. Kojima and G o t 6 found
that the h o n e y c o m b domains are m o r e likely to appear in a defect-less crystal
which can easily be changed to a crystal showing maze remanent domains by
water quenching or pricking with a needle.
M o r e quantitative relations have been obtained on the variation of domain
width D with crystal thickness t. For simple slab domains, the relevant theoretical
equation is given by Kittel (1946) as,

D = (O-w/1.7Ms)l/2t m • (16)

For the domains with reverse spikes near the surface, Kaczdr (1964) modified the
relation as:

D = [0.3751/Ms)(O'wtX/Tr)l/212/3t2/3, (17)
where
[,, = 1 + 8 r r 2 M d K 1 .

G o t 5 (1966) obtained an experimental formulae D = 0.392t °665 for B a M and


D = 0.441t °64° for SrM in the thickness range 2 t x m < t < 2 0 txm. On the other
hand, Kacz6r and G e m p e r l e (1960) reported D oc t 1/2 for t < 10 txm and D o c t 2/3
for t > 10 Ixm on the domains of PbM. Rosenberg et al. (1966), moreover,
obtained the relations for S r O - ( 6 - x)Fe2OyxA1203; D oc t 0.586 at x = 0, D oc t 0-616
at x = 1.0, D oct 0-665 at x = 1.5, D oct °.418 at x = 1.8, D oct 0.ass at x = 1.9 and
D o c t 0.431 at x ~ 2.0. For thinner crystals observed by G o t 6 (1966), the results can
be rewritten as D = [0.386(t- 1.460)] 1/2 for B a M and D = [0.432(t- 1.1260)] 1/2 for
SrM respectively. H e n c e the one half Power law may be valid in this case, though
the physical meanings of the constants are not clear. However, if we consider the
existence of a surface layer where the Kittel model might b e c o m e unstable, giving
a kind of lattice distorted layer, and also the resolution limit in m e a s u r e m e n t s
with an optical microscope, the numerical values in these formulae seem to be
reasonable. For thicker crystals, the two third power law related to spike domains
seems to be valid. Some deviation from the law could be understood by the
complicated domain structures and the resulting ambiguity in the measured values
of domain widths.
358 H. KOJIMA

The temperature variation of the domain width D in BaM was studied by


Kojima and Goff) (1962). They obtained the temperature coefficient of D as
8.9× 10-4/°C and that of O'w as - 2 . 3 × 10 3/°C. Gemperle et al. (1963) also
performed similar experiment with PbM and found the thermal hysteresis of the
domain width, as Shimada et al. (1973) observed with honeycomb domains.
Kacz6r (1972) discussed these results from a theoretical point of view and showed
the free energy decreases linearly to zero as T/Oc changes from 0.2 to 1.0, and the
domain width almost doubles in the same temperature range. Furthermore,
undulating Bloch walls, for which Goodenough (1956) first gave a theoretical
explanation, can be seen on the surface of crystals with medium thickness, for
instance, 10 fxm < t < 50 txm for BaM. Szymczak (1971) reported the temperature
dependence of domain width, wave amplitude and wave length in these domains.
The stability with temperature for honeycomb domains was investigated by
Gemperle et al. (1963) with PbM and by Shimada et al. (1973) with BaM. The
latter authors observed the increase of the nearest neighbour distance among the
cylindrical domains during the temperature rise. The honeycomb domains rever-
sibly change to a mixture of honeycomb and stripe domains by the heat treatment
from 600 K to room temperature. It was pointed out that the equilibrium distance
theoretically predicted by Kaczdr and Gemperie (1961) would be realized only in
such a mixed domain structure. Regarding the same phenomenon, Kozlowski and
Zietek (1965) showed from a theoretical consideration that the deviation from
Kacz6r and Gemperle's equation in these experiments would rapidly increase for
thinner specimens.
Grundy (1965) observed Kittel type slab domains in PbM of 1000-2000
thickness by Lorentz microscopy and determined the Bloch wall thickness as
250 _+150 A. Grundy and Herd (1973) used the same technique and applied it in
an investigation of the nucleation mode of bubble domains. They gave the
material length l = Crw/(4~-M2) as 0.03-0.04 Ixm for BaM and PbM. Wall mobility
constant ,1 of 0.7× 102cm/s/Oe for BaFe12019 and 1.6× 102cm/s/Oe for
BaFeu.aA10.7019 were reported by Asti et al. (1968).

2.1.12. Optical properties


The absorption coefficient a (=2~-K/nA) and Faraday rotation &F of BaM
measured at 300 K as a function of wavelength from 1 ~m to 8 p~m by Zanmarchi
and Bongers (1969) are shown in fig. 49. It is seen that &V changes sign between
2 ~m and 3 ~m. Drews and Jaumann (1969) measured the absorption coefficient K,
refractive index n, Faraday rotation &F, Faraday ellipticity ~/v and Kerr rotation
against air ~bw and against glass &KC for the same material in the shorter
wavelength region of 0.4 ~m-1.7 ~m. The results are illustrated in fig. 50. Kahn et
al. (1969) also added the data of the polar Kerr spectra for PbM, showing° a
negative peak at 4.43eV (2799,~) and a positive peak at 5.5 eV (2254A).
According to their conclusion, charge transfer transitions occurring at about 4 eV
(3100 A) and 5 eV (2480 A), associated with Feoct and Fetet complexes, respec-
tively, are responsible for the principal magnetooptical spectra. Blazey (1974)
reported on the wavelength-modulated reflectivity spectra of BaM with the
minima at 2.2eV (5636A) and 2.6eV (4769 A) corresponding to the internal
150

I00

'E 300
¢j

13
200 E
rj

50
0
I00 "o
v

LL
-g.
0

0 '-I00
0 2 4 6 8 I0
X (Fm)
Fig. 49. Absorption coefficient a(=2~'K/nA) and Faraday rotation ~bF of BaO.6Fe203 in infrared
region at room temperature (Zanmarchi and Bongers 1969).
xlO 2
I0.0

7.5 1~F
o I-- xlO 2
o~ 5.0 - - "F]F -2.0

LL 2.5 --I.0 o
E
g,
~,, o o 2
- 2.5 1.0 -8-

6
-5.0 2.0

I
3 m

r- i o- 2

2 - -
L
10.4
I
0 0.5 1.0 1.5 2.0
X (/~m)
Fig. 50. Optical properties of BaO-6Fe~O3 as a function of wavelength. K: absorption coefficient, n:
refractive index, OF: Faraday rotation, rTF: Faraday ellipticity, CbKL:Kerr rotation against air, ~bKo: Kerr
rotation against glass, t: 4.5 p~m, / 4 : 1 5 kOe [1.19 x 106 A/m].

359
360 H. K O J I M A

transition of Fetet and at 3.9 eV (3179 A), 4.3 eV (2883 A) and 4.8 eV (2583 A),
these being assumed to be charge transfers to Feoct.

2.1.13. Magnetostriction
The saturation magnetostriction in a hexagonal crystal is given by Mason (1954) in
the form,

a = /~A[(OZI~I q- a2~2) 2 -- (O/1]~ 1 q- O~2][~2)@3~3]q- ~.B[(I -- O~2)


x ( I - fl~) ~ - ( o ~ # , + o~=/?~)=] + a d o - o d ) B 1 - (o~/3, + o~fl=),~,8~]
+ 4AD(alB1 + o12f12)a3B3, (18)
where al, a2, O~3 and ill, fi2, r3 are the direction cosines of the magnetization
vector and measuring direction. Here, the direction cosines are taken with respect
to the Crystal axes, the z-axis coinciding the c-axis. Kuntsevich et al. (1968)
determined the constants AA, As, Ac and AD in eq. (18) for BaM in measurements
with the following geometry: •A:O/1 =/31: 1, AB:OZl= /3== 1, Ac:OZl = J~3=
1, h D : a l = / 3 1 = a 3 = / 3 3 = l/X/2. Thus, they found the constants at room
temperature as AA = -- (15 --+0.5) X 10-6, AB = + (16 + 0.5) X 10-6, Ac =
+ (11 -+ 0.5) X 10-6 and AD = -- (13 --+0.5) X 10-6. For polycrystals, they also obtained
the longitudinal and transverse magnetostrictions, hi! = -(9-+ 0.5)x 10 6 and h~ =
+ (4.5 -+ 0.5) x 10-6. However, these values do not coincide with the values derived
by the simple averaging of the formula for a single crystal. The authors explained
these results from the effects of the defects in the crystals and the interference of
grains during deformation.

K E 6
I
E
8.6

15.0 6.4

8.4 6.3

6.2
14.5
8.2 6.1
Z
6.0
/
8.0 "14.0' 5.9
I00 200 300
T (K)
Fig. 51. Temperature dependence of Young's modulus E, rigidity modulus G and bulk modulus K
(x 1011 dyn/cm 2) [× 101° N/m 2] (B.P.N. Reddy and P.J. Reddy 1974a).
F U N D A M E N T A L P R O P E R T I E S OF H E X A G O N A L F E R R I T E S 361

2.1.14. Mechanical properties


Fundamental studies of the mechanical properties of hexagonal ferrites are quite
few. Clark et al. (1976) referred to the following mechanical data for their BaM
specimen at room temperature. Density: 5g/cm 3, porosity: 5%, Young's
modulus: l a x 106kg/cm2 [1.38x 1011N/m2], Poisson's ratio: 0.28, compressive
strength: 4.5 x 103kg/cm2 [4.41x 108N/m2], tensile strength: 5.6x 102kg/cm2
[5.52x 107 N/m2]. Reddy and Reddy (1974a) measured the elastic modulus of
sintered BaM with a density of 4.8910 g/cm3. Figure 51 shows the relations of
Young's modulus E, rigidity modulus G and bulk modulus K versus temperature.
These moduli decrease with temperature, in contrast with those of Ni-Zn or
Mn-Zn cubic ferrites. Hodge et al. (1973) investigated the compressive defor-
mation of sintered BaM with 18% porosity in creep and press forging modes in
the temperature range 1000° to 1200°C. Figure 52 is the true strain rate against
l I T plot for an isotropic compact at 5% true strain. The activation energy for
creep was estimated from the experiments as 123-+6kcal/mol [(5.15-+0.25)x
105 J/moll.

-2
(x
13

--<..
..3 , ~ " ~
-6

6.8 7.0 7.2 7.4 7.6 7.8


I/T ( K -I x l O -4)
Fig. 52. Arrhenius plot of creep by compressive deformation of isotropic BaO.6Fe203 (Hodge et al.
1973). (1) 421.8 kg/cm 2 [4.14 x 104 N/m2], (2) 281.2 kg/cm 2 [2.76 × 104 N/m2], (3) 140.6 kg/cm 2 [1.38 x
104 N/mZ].

2.1.15. Heat capacity


Reddy and Reddy (1974b) determined the heat capacity of a BaM sintered rod in
the temperature range 80 to 303 K. Cp is plotted versus temperature in fig. 53,
giving values of 52.42 cal/mol deg [2.194 x 102 J/mol deg] and 107.56 cal/mol deg
[4.502× 102J/mol deg] at 80 and 303K, respectively. Furthermore, the Debye
temperature was estimated from the calorimetric measurements at 166 K and
from the elastic measurements as 158 K, both of which are rather low compared
with the temperatures of Ni-Zn or Mn-Zn ferrites.
362 H. K O J I M A

10

O0 -8
E
'-"a

90
Cp "-S
8 0 .x.

I0 70 "o
"5

8 / 60 -6
u

"e 6 50 °
%

2 :,.,"
0
0 I O0 200 300
T(K)
Fig. 53. Heat capacity and thermal expansion coefficient of sintered BaO-6Fe203. Cp: heat capacity
(Reddy and Reddy 1974b), (1) o~(Reddy and Reddy 1973), (2) a parallel to the magnetized direction, (3) a
perpendicular to the magnetized direction (Clark et al. 1976).

2.1.16. Thermal expansion


Reddy and Reddy (1973) also measured the thermal expansion coefficient o~ of
sintered B a M by a two terminal capacitance dilatometer. The result is illustrated
by curve (1) in fig. 53. They derived the following experimental equation in the
t e m p e r a t u r e range of 80 to 300 K:

a = 0.572 x 10 7 T - 1.437 x 101° T 2 + 0.175 x 1012 T 3 . (19)


Clark et al. (1976) gave the values c~ of B a M parallel and perpendicular to the
magnetized direction, which are shown as curves (2) and (3) in fig. 53. They also
reported the thermal conductivity of their specimen having 5% porosity as
4 W / m deg at r o o m temperature. A b o v e r o o m temperature, Davis (1965) obser-
ved the change of lattice constants in B a M up to 1000°C by X-ray diffraction, but
the value is fairly high and no noticeable change was found around the Curie
point. K 6 m o t o and Kojima (1976) showed the t e m p e r a t u r e variations of the a and
c-axis in B a M in fig. 54, where the sharp kinks can be seen at 450°C for both axes,
but the value is rather low. In table 13, the expansion coefficients of B a M by
several authors are compared. The cause of the discrepancy in the list may be
explained by the difference of porosity, grain size, degree of crystal orientation,
chemical composition in the specimens, or accuracy of measuring methods.
FUNDAMENTAL P R O P E R T I E S OF H E X A G O N A L FERRITES 363

A5
7
o
X
--4
O
<l

O 3
1D

0
0 200 400 600
Temperature (°C)

Fig. 54. Variations of the lattice constants in BaO.6Fe203 with temperature (K6moto and Kojima
1976).

T A B L E 13
T h e r m a l e x p a n s i o n c o e f f i c i e n t o f B a M (x 10-6/deg).

T (°C) R.T. 0° 27 ° R.T.

a 8.5 (a) 8.4 c°) 8.8 ~) 11.1 (d)

T (°C) 25-150 150-250 25-250 25-640

~(e) 9.5 10.5 9.8 10.2

T(°C) 10 - 100 - 200 - 250 - 300 - 350 - 450 - 550

an (f) 5.7 6.8 8.2 10.5 18.6 20.7 11.5


ao (f) 10.8 11.0 12.1 13.8 17.2 17.2 7.3

* By X-ray diffraction
(a) H a b e r e y et al. (1973) (d/Davis* (1965)
Co)R e d d y a n d R e d d y (1973) (e) B u e s s e m a n d D o f f (1957)
(c>C l a r k et al. (1976) If) K 6 m o t o * a n d K o j i m a (1976).
364 H. KOJIMA

• 2.1.17. Electric and dielectric properties


Zfiv&a (1963) observed the electrical conductivity of BaM and PbM single crystals
and found that the relations among the conductivities are approximately expres-
sed by 100"BaM ~" OrpbM and 10or//~ o-1 for both compounds. Here, o-B~ and O-pbM
mean the values of the conductivities with each ferrite, and o-//and o-~ denote the
values measured along and perpendicular to the c-axis, respectively. Figure 55(a)
and (b) show the conductivities of BaM and PbM, rewritten from these results by
Wijn (1970). Z~veta also estimated the activation energy for the temperature
dependence of the conductivity and, moreover, reported on the thermoelectric
force for the same specimens. He finally concluded that an electron hopping
mechanism between Fe ions might reasonably be assumed and so the conductivity

TIE I0-I
T

b id z

io-3

/o)

TE 10-3
T ",,,.\
b 164
II C

id 5

I
2 3 4 5 6 7
I/T (xlO-3/deg K)
Fig. 55. Temperature dependence of the electrical d.c. conductivity of the single crystals: (a)
BaFe12019, (b) PbFe12019. The different marks indicate the measuring points of different crystals
(Zfiveta 1963).
F U N D A M E N T A L P R O P E R T I E S OF H E X A G O N A L F E R R I T E S 365

10-2

Ld ~

10-4

IE
0-5

b
10-6

to-T

io-e

j0-9
2 5 4 5 6 7 8 9
I/T (xlO-3/deg K)
Fig. 56. Temperature dependence of the conductivity of BaFe12019 polycrystal containing 0.022% Fe 2+
by weight. The figures stand for the used frequency in Hz (Haberey and Wijn 1968).

104' ' I '

I0 .c,-''~"

I ---
! IOM ' i!

/// o
1

,o
-150 -I00 -50 0 50 I00 150
T (°C)
Fig. 57. Temperature dependence of the real part of the dielectric constant for the same specimens in
fig. 56 (Haberey and Wijn 1968).
366 H. KOJIMA

is related to the presence of Fe 2+. D u l l e n k o p f and Wijn (1969) r e p o r t e d similar


experiments in the frequency range of 1 to 8 G H z and b e t w e e n - 5 5 ° and 100°C.
T h e specimens were polycrystals containing Fe z+ up to 0.5% by weight. T h e
m o d e l of a thermally activated h o p p i n g process again received support in this
work. Figure 56 shows the t e m p e r a t u r e d e p e n d e n c e of the conductivity for the
p r o p o s e d composition of BaFe12.59019.39
3+ polycrystal, d e t e r m i n e d by chemical
analysis, at various frequencies r e p o r t e d by H a b e r e y and Wijn (1968). T h e y also
described the t e m p e r a t u r e d e p e n d e n c e of dielectric constants E' for the same
specimen, as shown in fig. 57.

TABLE 14
Properties of the magnetoplumbite type compounds substituted Ba ion.

o's (Gcm3/g)
Compound a (A) c (A,) [x 4zr x 10-7 Wbm/kg] NB (/xB/mol)

LaFe2÷Fe3+nOig* - 65 (R.T.) 19.5 (0 K)


13 (300 K)

69 (R.T.)
Na0.5La0.sFe12019 - 113 (0 K) 21.5 (0 K)

Ca0.88La0.i4Fe12019* 5.877 22.91 52 (R.T.) -


73.4 (77 K)

Ag0.sLa0.sFe12Oi9 5.85 22.85

LaFea2019* 96.2 (0 K) 19.2 (0 K)

LaFei2019*

LaFe12019*

Tl0.sLa0.sFea2019

(a) Aharoni and Schiber (1961) (e) Lotgering (1974)


(b) Summergrad and Banks (1957) (f) Van Diepen and Lotgering (1974)
(c) Ichinose and Kurihara (1963) (g) Drofenik et al, (1973)
(d) Laroria and Shinha (1963)
F U N D A M E N T A L P R O P E R T I E S OF H E X A G O N A L F E R R I T E S 367

2.2. Substituted M compound

2.2.1. BaO-SrO-PbO-Fe203 systems


As already discussed, replacement of Ba 2+ ion with Sr 2+, Pb 2+ or both of them is
possible with any mixing ratio without changing the crystal structure. For
example, Ba0.75Sr0.zsFea2019 (Borovik a n d Yakovleva 1962a, 1962b) and
Sr0.7sPb0.25Fe12019 (Borovik and Yakovleva 1963) attracted special interest owing to
the improved properties as permanent magnet materials. Hunty (1963) however,
found no improvement of these mixed ferrites. Kojima and Miyakawa (1965)

T A B L E 14 (continued)

KI (erg/cm 3) /(2 (erg/cm 3) Other data


[x 10 ~J/m 3] [x 10-1J/m 3] 0c (K) Composition in the paper Ref.

2,5 x 10 6 -- 695 NB-T


- G-La205 (a)

o--T
3.0 x 106 - 713 _+ 10 -
o-j/, o - l - H

- - 718 CaO 13.70 o-s-La203 (c)


Fe203 82.78
La203 2.00
FeO 1.52
(mol %)

708 Br = 1600 G
MHc = 2450 Oe (d)
BHc = 1100 Oe

(10-13) x l0 s (0 K) 0 (293 K) LaM 86.2 o-//,o-±-H


(5-10) x l0 s (77 K) Fe304 10.3 o--T (e)
LaFeO3 3.4 K1-T
(wt %)

Fe304 10 M6ssbauer spect.


(wt %) Hhf-T (f)

697 -+ 2 Fe304 M6ssbauer spect. (g)


removed by e-T
10% H2804 8-T
Hi-T

M6ssbauer spect. (g)


368 H. KOJIMA

investigated the changes of lattice constants, saturation magnetization and Curie


points for the Ba-Sr, B a - P b and Sr-Pb binary hexaferrites systems and found that
there are no sharp kinks on the curves relating the properties and the com-
position. These facts suggested that those systems form complete solid solutions
between both end m e m b e r s and no special c o m p o u n d can exist at any mixing
ratio. Therefore, it is m o r e plausible that the i m p r o v e m e n t s due to the sub-
stitutions in the Borovik's papers might be due to the acceleration of the solid
state reaction or sintering process etc., and not to the intrinsic properties of a
specific compound. Though such non-intrinsic but practically useful effects in the
p e r m a n e n t magnet properties of hard ferrites introduced by many kinds of
additives were once actively studied the results will n o t be mentioned further,
since the details will be discussed in ch. 7 by Stfiblein.

2.2.2. Other substitutions of B a 2+ ions


Ferrimagnetic properties of the magnetoplumbite type compounds are naturally
based on the superexchange interaction of F e - O - F e , so that no remarkable effect
on the magnetic properties would be expected by the replacement of a Ba ion.
However, the properties could be influenced to some extent by the changes in the
distance or the angle between F e - O - F e arising from the substitution, leading to
differences among BaM, SrM and PbM.
Table 14 shows the properties of several examples for Ba 2+ substituted ferrites.
In some cases m a r k e d * in the table, the partial coexistence of Fe 2+ may be
unavoidable in order to keep electrical neutrality?. The examples are so few that
no general tendency can be deduced from the table. As described before, when
discussing the M6ssbauer effect, Drofenik et al. (1973) concluded from the
t e m p e r a t u r e dependence of the hyperfine fields for L a M that a m o r e convex
magnetization against t e m p e r a t u r e curve than SrM comes from the difference of
the behaviour of the 12k sublattices. It is also pointed out in this p a p e r that in
LaFet2019 and T10.sLa0.sFe12019 no Fe 2+ sublattice was observed in the M6ssbauer
experiment and the averaged hyperfine fields could originate from a fast electron
exchange between Fe 2+ and Fe 3+, or in the process of electron sharing by the six
ions in the 12k sites of these compounds.
Van Diepen and Lotgering (1974) insisted from a similar study that the
substitution of Ba 2+ with La 3+ causes a valency change of Fe 3+ to Fe 2+ at the 2a or
4f2 site, although they also found no subspectrum of Fe 2+.

2.2.3. Substitution of Fe 3+ ions

Substitution with A13+, Ga 3+ and Cr 3+ ion. T h e replacement of Fe 3+ in B a M or


SrM with AP +, G a 3+ or Cr 3+ has been intensively investigated. Because of the
resemblance of their ionic radii with Fe 3+ (shown in table 2) especially the former
two ions are easily replaced at any substitution ratio without changing the crystal
structure. Figure 58(a) and (b) show the variations of the lattice constants a and c

t Replacements with combinations of a few ions are also cited in the next section.
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 369

23.2

22.8

o,=I
22.6
(j,

22.4

(b
22.2

i
22.0 i

0 2 4 6 8 I0 12
X
Fig. 58. (a) Lattice parameter a, and (b) c as a function of substituted ratio. (1) BaAlxFe12-,Om, (2)
BaGaxFe12-xOm,(3) BaCr, Fe12-xO19,(4) SrCrxFem-xO19(Bertaut et al. 1959), (5) SrAlxFe12-xOt9(Goto
and Takahashi 1973).

for (1) BaAlxFe12_x019, (2) BaGaxFe12-xO19, (3) BaCrxFe12-xO19, (4) SrCrxFelz-xO19


(Bertaut et al. 1959) and (5) SrAlxFem_xO19 (Goto and Takahashi 1973). Almost
the same results as given on line (1) for the Ba-A1-M system by Vinnik and
Zvereva (1969) and also on line (5) for SrAlxFe12-~O19 were reported by Florescu
et al. (1973). The parameters decrease linearly with the substitution ratio in all
cases. Numerical values for the end members were given by Adelsk61d (1938) and
Bertaut et al. (1959) as in table 15. This table also contains the results for
Ca/Sk112019 by Wisnyi (1967); Ca(AlFe)12019 by MacChesny et al. (1971);
BaGaa2019, SrGa~2Oa9 and LaMgGanO19 by Verstegen (1973).
370 H. KOJIMA

TABLE 15
Lattice parameters of some substituted M-type compounds with A1, Ga or Gr.

Formula a (4) c (4) Ref.

BaA112019 5.577 22.67 Adelsk61d (1938)


5.66 22.285 Bertaut et al. (1959)

SrAl12019 5.557 21.945 Adelsk61d (1938)

CaA1~2OI9 5.566 22.010 Wisnyi (1967)

Ca(A1Fe)12019 5.792-+0.004 22.56_+0.04 MacChesneyet al. (1971)

BaGa12019 5.818 23.00 Bertaut et al. (1959)


5.850 -+0.004 23.77 _+0.02 Verstegen (1973)

SrGa12Oa9 5.796 -+0.004 23.77 _+0.02 Verstegen (1973)

LaMgGa11019 5.799-+0.003 22.71 -+0.01 Verstegen (1973)

BaCrsFe4Oa9 5.844 22.82 Bertaut et al. (1959)


SrCr6Fe6019 5.844 22.77 Bertaut et al. (1959)

Van Uitert (1957), and Van Uitert and Swanekamp (1957) reported 4¢rMs of
BaAl~Fe12_xO19, (1) fired below 1400°C and (2) fired above 1400°C, and (3)
SrAlxFe12_xOi9 after Rodrigue (1963) as shown in fig. 59(a). Figure 59(b) illustrates
(1) BaAlxFe12_xO19; (2) BaGa~Fe12_~O~9 (Mones and Banks 1958); (3)
SrAlxFe,2_xO~9 (Rodrigue 1963) and (4) ditto (Gotto and Takahashi 1973). Lines
(3) and (4) agree quite well. In figs. 60(a) and (b), Bertaut et al. (1959)
summarized NB at 0 K per mole in Bohr magneton and 0c for (1) BaAl~Fe~2_xO~9,
(2) BaGa~Fe12_xO19, (3) BaCrxFe~2_xO~9; the m a r k A in (a) relates to
SrCrxFe12-~O19. Albanese et al. (1974) estimated 0c for BaAlxFe~2 xO,9 and
SrAlxFe~2-xO19 from the M6ssbauer experiments, as shown in table 16. The lines
D~ and D2 in fig. 60(a) were drawn on the hypotheses that AP + ions enter the 2a
and 12k sites, or G a 3+ the 4fl and 12k sites, respectively (after Bertaut et al.
(1959)). With these results, and in addition referring to the X-ray investigations,
they came to the conclusion that AI 3÷ and Cr 3+ first occupy 2a sites and then 12k
sites, while G a 3+ enters first 4fl and then 12k sites. These assumptions are
thoroughly supported by Rensen et al. (1971) by M6ssbauer studies, but in
contrast with these results, Suchet (1971) expressed a strong preference of AI 3÷ to
the 4f~ and 2a sites. Glasser et al. (1972) observed M6ssbauer and E S R spectra for
CaAl12-xFe~O~9 with x <~ 4.8, i.e., mainly on the A1 rich side. They recognized a
m a r k e d preference of Fe 3÷ for the tetrahedral 4fl sites from M6ssbauer in-
vestigations and interpreted the majority of E S R resonances by supposing Fe 3÷
ions in an axially distorted site.
T h e relations of a, Ms, 0c, K, MHc and the critical radius for a single-domain
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 371

800 1 [ I
.(3) I i

60 O l ~ ' ~
r...<-., i Ib)
A
,.i
o

400
qb (.3

200
121

5000 I
(0)
3)
4000
!
I
i

~o 3 0 0 0
~', ,,,
x.
2000

I== 1000

0 I
0 2 4 6 8 10 12
X
Fig. 59. (a) Saturation 4~rM~, and (b) Curie temperature 0c as a function of substituted ratio. (a) 4~-Ms:
(1) BaAlxFe12-xO19 fired below 1400°C, (2) ditto fired above 1400°C (Van Uitert and Swanekamp
1957), (3) SrAlxFelz-xO19 (Rodrigue 1963); (b) 0c: (1) BaAlxFelz-xO19, (2) BaGa~Fet2 xO19 (Mones and
Banks 1958), (3) SrAlxFel2-xOl9 (Rodrigue 1963), (4) ditto (Goto and Takahashi 1973).

particle Rc etc. versus x for BaCrxFel2 xO19; and BaAlxFel2-xOi9, BaGaxFe12-xO19


and BaCrxFe12_xOx9 were given by Haneda and Kojima (1971 and 1973b). In fig.
61, curves (3), (4) and (5) indicate H A for BaAlxFe12-xO19, BaGaxFei2-xO19 and
BaCrxFei2 xO19 (Haneda and Kojima 1973b) besides the marks • and A illustrat-
ing HA for SrAlxFe~2_xO19, as obtained by De Bitetto (1964) and Rodrigue (1963).
Curves (1) and (2) are drawn with the calculated values of eqs. (20) and (21) by De
Bitetto who gave the experimental formulae:
372 H. K O J I M A

800 I
(b)
700

A
•v" 6 0 0

500 --

400

300

25 I

(a)
20'~

\2a \\ D2
,

D,,\ \ )--'~,X~L
0 2 4 6 8 I0 12
X
Fig. 60. (a) Saturation at 0 K and, (b) Curie temperature versus substitution ratio (Bertaut et al. 1959):
(1) BaAlxFe12-xO19, (2) BaGaxFeu-xO19, (3) BaCrxFeu-xO19, (A) SrCrxFelz-xO19, D1, D2: calculated
values.

T A B L E 16
Curie temperature for the compounds
BaAlxFe12-xO19 and SrAlxFelz_xOl9 (AI-
banese et al. 1974).

BaAlxFei2-xO19 SrAlxFe12-xO19

x Tc (K) x T~ (K)

0 723 -+ 5 0 723 -+ 5
2.5 595 ---5 1 658 -+ 5
4 518---5 2 590---5
FUNDAMENTAL PROPERTIES OF H E X A G O N A L FERRITES 373

7O , I
(2)
I
6O I
I
// °/(I)
5O
z~
/,/
-g / •/
40 //~e'~ 131
~D
o
X
I// /
0

"1-

~ (51

I0

0
0 I 2 3 4
x

Fig. 61. Anisotropy field HA as a function of substitution ratio: (1) and (2) see text, (0) SrAlxFe12-xO19
(De Bitetto 1964), (A) ditto (Rodrigue 1963), (3) BaAlxFe12-xOxg, (4) BaGaxFex2-xO19, (5) BaCrxFe~z_xO19
(Haneda and Kojima 1973b).

HA(x)/H(o) = (12 - x)/612 - x/2 + (x/6)zss], (20)

or

HA(x)/H(o) = (12 - x ) / ( 1 2 - 3 x ) . (21)

Since D e B i t e t t o e x p e r i m e n t a l l y o b t a i n e d a r e l a t i o n for K as a f u n c t i o n of x for


SrAlxFe~z-xO19 in t h e f o r m ,

K(x~/K(o) = 1 - x / 1 2 , (22)

w e can say that t h e m a g n e t o c r y s t a l l i n e e n e r g y m a y b e r o u g h l y a s s u m e d to h a v e


a b o u t t h e s a m e s t r e n g t h for all t h e five sublattices. A s s u m m a r i z e d in fig. 62, t h e
M s - x d e p e n d e n c e is r e l a t e d to t h e v a l i d i t y of eq. (20) o r eq. (21). D e B i t e t t o
c l a i m e d t h a t an e m p i r i c a l fitting f o r m u l a ,
374 H. KOJIMA

1.0

0.8

1~ O.6

0,4

0.2

0 I 2 3 4 5 6
x
Fig. 62. Variations of normalized saturation magnetization with composition in MAlxFe12-xO19(De
Bitetto 1964): (0) Ba (Bertaut et al. 1959), (A) Ba (Mones and Banks 1958), (V) Ba (Van Uitert and
Swanekamp 1957), (V) Pb (Bozorth and Kramer t959), (D) Ba, (C)) Sr (Du Pr6 et al. 1958), (&) Ba
(Vinnik and Zvereva 1970).

M(x)/M(o) = ~[2 - x/2 + (x/6)zss], (23)

which is shown by curve (2) in fig. 62, is m o r e plausible than most other authors'
results given by curve (1), i.e. by

M(x)/M(o) = 1 - (x/4). (24)

D a t a by Vinnik and Z v e r e v a (1970) for the B a - A 1 - M system m e a s u r e d at 4.2 K


were a d d e d to fig. 62. T h e y concluded that the preference of A13+ seems to agree
with the results of Bertaut et al. (1959) and the deviation at x = 3.3 m a y be due to
the differences of applied field. H o w e v e r , we would like to point out here that the
occupied site with a substituted ion might be different according to the pre-
paration conditions, starting materials and various other factors, although the
substitution ratio m a y be perfectly identical.
HA was d e t e r m i n e d by D e Bitetto using F M R m e a s u r e m e n t s and the line width
A H was also given for the S r - A I - M system between x = 0-1.7. A l b a n e s e et al.
(1974) reported H i - T and Hi(T)/I-L(o)--T/Oc relations by M6ssbauer investigations
for A1 or G a substituted B a M and SrM. H e i n e c k e and Jahn (1972) studied the
angular variation of the nucleation field Hn for single d o m a i n particles of
SrA13.sFes.2019 and c o m p a r e d the Hn-O relation with the 1/cos 0 law or the
c o h e r e n t rotation model. Interesting investigations of the f o r m a t i o n m o d e of the
domain structures on the c-plane of PbA1Fel~O19 were r e p o r t e d by Williams and
S h e r w o o d (1958) using a motion picture technique. Tanasoiu (1972) o b s e r v e d
r e m a n e n t domain structures on the c-plane of SrAI4.8FeT.2OI9, the main magnetic
properties of which were previously d e t e r m i n e d as 4~rMs = 315 G [315 × 10 -4 T],
To= 5 1 8 K and Crw= 5 . 8 e r g / c m 2 [5.8× 10-3j/m2]. H e f o u n d that the r e m a n e n t
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 375

structure after thermal demagnetization consists of a great number of domains


with irregular shape and are very unstable, while the remanent state by a.c.
demagnetization consists of only a few domains.
MacChesney et al. (1971) studied the ferrimagnetic phase of the CaO-A1203-
Fe203 system and, as for the hexagonal magnetic phase, they found the com-
position range extends from CaAI~20~4 to at least CaA14.2FetsOls.6. The lattice
constants of an idealized composition Ca(A1Fe)12019 were already referred to in
table 15. The magnetic moment at room temperature for the group varies from 1
to 10 emu/g [12.56 to 125.6 x 10-~ Wbm/kg] according to the Fe content, but they
show rather high anisotropy fields of about 60 kOe [4.78 x 10 6 A/m] regardless of
the composition.
1 / x - T relations up to 1000 K were determined by Florescu et al. (1973) for
SrAlxFe~2-xO19 from x = 2.0 to 9.0 and they estimated N6el's parameters by the
graphical method (N6el 1948) from these results. The photoluminescence of
BaGa~2019, SrGa120~9 and LaMgGauO~9 with and without Mn 2+ phosphors was
reported by Verstegen (1973) who found that the emission strongly resembles that
of Mn 2+ in MgGa204.

Substitution with Sc 3+ and In 3+. The partial replacement of Fe 3+ with Sc3+ or In 3+


ions in M-type compounds were investigated mainly by research groups in the
U S S R with an interest in their specific magnetic ordering. Perekalina and Che-
parin (1967) measured the relations of a, c-x, ~rs-T and K ~ - T for BaScxFe~2 xO19
single crystals with x = 1.2-1.8, as shown in table 17 and figs. 63 and 64. Maxima

80

7O
• .~,^ "" o

"~
r
6o

'~~4-O0~5_xS~=~~~
× 0~2,03~"
0~
C31

10

0 r I,o
0 200 400 600
Temperature (K)
Fig. 63. Temperature dependence of the specific magnetization BaScxFe12-xOa9single crystals (Pere-
kalina and Cheparin 1967): (1) x = 1.2, (2) x = 1.4, (3) x = 1.8.
376 H. KOJIMA

o< o<

+1 +1

tt3 t'¢3 t'e'j "~"

'~" ¢xl t'xl ('xl ~:7~ t'~ , ~ ,.~j tt3


. . . . . . 0~
x ~., trq

tt3 "~" ,,..~


" X
uS,.-~o~ ~ u E e 6

b,

©
X

,,~ ~ t¢3 t ~ t'q e¢3


r~

.=

. . . . . . . . ¢¢3

~D

.=
~mmmm~m~ e~
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 377

9 r
(I).

8 /,
I
7 \_

6
E
5
o
X
4
tO
E
-2 3
03

%2
X

-I

-2 I I
0 200 400 600
Temperature (K)
Fig. 64. T e m p e r a t u r e d e p e n d e n c e of t h e a n i s o t r o p y c o n s t a n t K1 for BaScxFelz-xO19 ( P e r e k a l i n a a n d
C h e p a r i n 1967): x = 1.2, x = 1.4, x = 1.8.

in the O-s-T relations, which shift towards higher temperatures with increasing Sc
content, and the change of K1 for BaScl.sFe10.2019 to negative values between 125
and 355 K are in contrast with pure M-type compounds. It was suggested that the
t e m p e r a t u r e dependences of ~rs can be explained by N6el'S theory (N6el 1948).
Aleshko-Ozhevkii et al. (1968, 1969) determined the magnetic structure by
neutron diffraction studies at 77 K in magnetic fields up to 5 k O e [3.89 x 105 A/m]
for BaScxFe12-xO19 with x = 1.2-1.8. They showed conclusively that Sc 3+ ions
occupy 4fl and 2b sites, so that antiphase conical helixes were formed in certain
blocks of the crystal unit cell. In a later p a p e r these authors estimated (i) the
periods along the c-axis; (ii) the cone vertex half angle; and (iii) the phase angle of
the projection of theo total magnetic m o m e n t s in the basal o
plane which were,
respectively: (i) 141A, (ii) 30 °, (iii) 150 ° for x = 1.8; (i) 91 A, (ii) 20 °, (iii) 135 ° for
x = 1.4; and (i) 70 A, (ii) 12 °, (iii) 120 ° for x = 1.2. Alesko-Ozhevskii and Yamzin
(1969) added a detailed explanation of the intensity anomalies of the satellites in
the neutron diffraction patterns of BaScxFe12-xO19 on the basis of the super-
position of ferro- and antiferromagnetic reflections. Koroleva and Mitian (1971)
378 H. KOJIMA

suggested the possibility of the appearance of these magnetic antiphase helical


structure based on Moriya's rule (Moriya 1960). The same behaviour of o-s-T and
K1-T plots for single crystals of BaInxFe12_xO19 for x > 1.9 and PbInl.gFemO~9
were observed by Perekalina et al. (1970). Their main data are added to table 17,
which can also be understood by the model of two magnetic sublattices with a
weak exchange interaction. Bashkirov et al. (1975) carried out M6ssbauer studies
for BaIn3FegO19 and showed the presence of an angular magnetic structure, which
was presumably caused by the almost full occupation of 2b and 4fi sites with In 3+.
Magnetic order-disorder transition of the compound was observed between 300
and 250 K with an external field of H = 16 k O e [1.28 x 10 6 A / m ] and below 100 K
in the absence of an external field. The difficulty of realizing the saturated state
even in 50 k O e [3.98 x 106 A/m] for BaInz4Fe9.6Oa9 was reported by Perekalina et
al. (1970). These anomalous magnetic properties are due to the fact that collinear
structures are not realized even at helium temperatures.

Substitution of Fe 3+ with paired ions; Co-Ti system


Smit et al. (1960) measured the torque curves of BaCox/zTix/2Felz-x019 for x = 0.9,
1.0 and 1.5 at T = 90 K and H = 32 k O e [2.55 x 10 6 A/m] after thermal demag-
netization and suggested from the results that the characteristic details of these
curves were due to the effects of incomplete alignment, both on a microscopic and
an atomic scale. Rodrigue (1963) pointed out that the effective internal anisotropy
fields of B a M could be varied from virtually zero to 17.5 k O e [1.39 × 106 A/m] by
substituting with Co-Ti, and this is quite advantageous for applications to
microwave devices. In fig. 65, the data in this p a p e r were cited in (a), (b) and (d).
De Bitetto (1964) determined the relation of H A with x by r o o m t e m p e r a t u r e
microwave ferrimagnetic resonance measurements. The line widths A H of these
oriented polycrystalline compounds were all about 2 k O e [1.6× 105A/m]. His
results are also plotted in figs. 65 (a), (b) and (c). Lines (1) and (2) in (a) are drawn
based on G o r t e r ' s assumption (Casimir et al. 1959): (i) a random distribution of
both Co and Ti ions among only the nine octahedral sites (12k, 2a and 4f2), or (ii)
both ions are substituted preferentially for some Fe 3+ ions with only the eight
favourably oriented moments (12k, 2a, 2b). It seems that D e Bitetto's data are
situated on line (1), G o r t e r ' s results on line (2), but that Smit and Wijn's and
Rodrigue's data are just between the two lines. The microscopic structure of
C o - T i substituted samples were reported by Bonnenberg and Wijn (1968) using
an electron microprobe analyser. They found, for instance, that in the com-
position BaComTimFe~2019, the concentration of Fe, Co and Ba, Ti showed the
same distribution tendency in pairs but the opposite tendency in the sintered
grains. Investigations of the cation distribution on an atomic scale were reported
by K r e b e r and Gonser (1976). They compared the electric quadrupole splitting E
and the area of the M6ssbauer spectra for BaCox/2Tix/2Fe12-xO19 sublattices from
x = 0 up to x = 11 at temperatures above the corresponding Curie points of each
compositions. Their conclusions are as follows: The Fe 3+ ions in 4fl and 4f2 sites
are replaced for x ~<4 and the remaining octahedral 2a and 12k sites are
substituted randomly for 4 < x < 8. Finally, for x > 8, a weak substitution in the
F U N D A M E N T A L P R O P E R T I E S OF H E X A G O N A L F E R R I T E S 379

800 I , _,
I Idl I

400"
L i i
%

~,0< (c)
x

E
.2 z.0

%
-× 1.0

~g

18~
(b)
12 E

0
X

I
r.o~ ~ ~ _ (a)
§ o.8 - ...... --~

0.6
0 I 2
x
Fig. 65. Normalized magnetic saturation at room temperature, anisotropy field, anisotropy constant
and Curie temperature against composition in Ba(TiCo)xFe12-xO19. (a) M(x)/M(01: (©) De Bitetto (1964)
at room temperature, (11) Gorter (Casimir et al. 1959) at 77 K, (A) Smit and Wijn (1959 at 20 K, (3)
Rodrigue (1963) at room temperature, (1) and (2) calculated values (see text); (b) Ha: ((2)) De Bitetto
(1964), ( 0 ) Rodrigue (1963); (c) KI: De Bitetto (1964); (d) 0o: Rodrigue (1963).

2b site occurs. When comparing these conclusions with the above mentioned
results in fig. 65, it should be noticed that the latter observations correspond to
the paramagnetic region.

Other combinations with H-IVpairs. D u e to the interest in the effect of substitut-


ing ions on the magnetic properties of M compounds, attempts to replace Fe 3+ by
non-magnetic ions are still being reported occasionally. Since equal amounts of
380 H. KOJIMA

M2++ M 4+ are simplest for electrical neutrality in these substituted compounds, a


number of papers has been published regarding such a substitution. For instance,
Tauber et al. (1963) investigated BaZnx/2Irx/zFe~2 xOi9 single crystals with x = 0-0.6
and found that the lattice constant a increased from 5.893-+0.001 A to 5.920-+
o o
0.001A but that c remained constant at 23.194_+0.002A with increasing I r
content. The o-~-T relations for x = 0, 0.32, 1.04 and 1.2, determined by the
chemical analysis of total Ir, were given in their paper. Other main results
obtained using FMR are summarized in table 18. They showed that the sub-
stitution of Ir 4+ causes the spin system in the compound to be pulled down toward
the basal plane. Similar FMR measurements have been done by Dixon et al.
(1971) with single crystals of BaZnx/zTix/zFelz-x019 for nominal compositions,
x = 2.0-5.0. They succeeded in decreasing the resonance line width A H to 36 Oe
[2865 A/m] with x = 3.0. The variation of A H versus x at 37.6 G H z was illustrated
in their paper. A summary of their results is given in table 18.
Rensen et al. (1971) discussed ZnTi, ZnGe, ZnSn, ZnZr, CuTi, CoZr or NiTi
substitution pairs. They referred briefly to the cation distributions for each
compound from M6ssbauer effect results. Magnetization, anisotropy field and
M6ssbauer spectra of BaZnxTiyMnzFe12_x_y_z019 were investigated between 4.2
and 300 K by Mahoney et al. (1972). The o-~-T curves for various nominal x
values, H~-x and e - x diagrams and tables of magnetic moments at 0 K and cation
distributions were given in their paper. They concluded that Ti 4+ mainly, and
Mn 3+ completely occupied 12k and 2a sites, and Z n 2+ w a s primarily assigned to 4fi
sites on crystal chemical grounds, but partly distributed on to 12k or 2a sites to fit
with the observed magnetic moment. Haneda and Kojima (1973b) measured HA,
MHc and Rc as a function of x for BaZnx/zGex/zFe12-x019 and BaZnzx/3Vx/3Felz-x019,
BaZn2x/3Tax/3Fe12-~019 and BaZn2x/3Nbx/3Fe12-~O19, also involving M2++ M 5+ pairs.
The main problem concerned the changes of MHc with x for these compounds and
this was also considered in connection with the changes of H A and also Rc for
sintered specimens.

2.2.4. Effect of substitutions on the temperature dependence of magnetization


An attempt to improve the temperature dependence of the magnetization of the
magnetoplumbite structure, which is inferior by one order than for Alnico
magnets, was perhaps reported f o r the first time by Heimke (1960). H e showed
that the Ms-T relation could be changed for the M-type compound
BaTiO3"5Fe203+ 1.5% CaSiO3, that is, its magnetization varies almost parallel
relative to the temperature axis below 0°C, as illustrated by line (1) in fig. 66(a)
(using the scale of the right-hand axis). For an explanation of this result, he
proposed two kinds of exchange schemes other than the one reported by Gorter
(1957). Esper and Kaiser (1972) continued this analysis and obtained interesting
results on the changes in the temperature dependence of O-s for M compounds.
Their o-s- T relations are shown in fig. 66(b) as curves (2) and (3), with (1) as the
normal relation of BaM. Their typical substituted compound has the composition
Ba(Sb203)0.25(Fe203)5.75 and the sample of curve (2), having almost the same
temperature coefficient a(o-s)= 0.2% per degree and about ~O-s of BaM, was
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 381

-g

o X ~ X

rO

{IN]{[{ {{
O

8 tt3

© ÷1
r
eq k~) tt~ ~ ee) tt3
0J

~2
t--
e~

¢e5

© ,<1
e'~

¢-
o
t~, ,,o

O
r.¢)

cq ¢¢~ ,~1- tt3 rqrd

.+ +
¢- © i--.

N
+ +
¢
~z
382 H. K O J I M A

I00
..L,~ i

80 (c)
60

40
i

20

E
I
..Q

b--
I0 I00 --. I
X

Bo -4) 1 (b )

--.<\-..~
b~ 20 ~.

I00 5000

N
E
8O
i~)-~ I
~_,p~.~,-~-~-~.
(0 ) - 4000

60 (5) _ ~ . , 5000 ~o
×

4O 2000

2O

0
i
I
400
\
600
i
800
I000

o
0 200
Temperature (K)

Fig. 66. Temperature dependence of the saturation magnetization o's for some substituted M com-
pounds. (a): (1) 4~rMs of BaTiO3'5Fe203 with the scale of the right ordinate (Heimke 1960), (2)
"4+ 2+ 3+ 5+ 2+ 3+
BaTlo.6Feo.6Felo.8Ol9, (3) BaSbo.sFel.oFelo.5019 (Lotgering 1974); (b): (1) BaM, (2) and (3) Sb-substituted
BaM (Esper and Kaiser 1972); (c): (1) SrM, (2) As-substituted SrM (Esper and Kaiser 1972).
FUNDAMENTAL PROPERTIES OF HEXAGONAL FERRITES 383

p r e p a r e d through the heat treatment in air and below 1300°C. The sample of
curve (3) received a secondary heat treatment above 1200 ° in N2 or above 1300°C
in air, after the normal solid state reaction. Sb205, i.e., Sb 5+ was used as a starting
material in both cases. It was claimed that the anomalous t e m p e r a t u r e depen-
dence of curve (3) was caused by the partial replacement of Fe 3÷ in 4f2 sites with
non-magnetic trivalent ions having ionic radius smaller than Fe 3÷. In this case, the
crystal structure or the magnetic coupling scheme gradually turns to the W-
structure from M-type, though the sample of curve (2) still has a distorted M
structure. In conclusion As and P were r e c o m m e n d e d as the most suitable
replacing ions to realize a low t e m p e r a t u r e coefficient around room temperature.
A valve c~(o-s)= 0.1% per degree was obtained with As substituted SrM as
illustrated with curve (2) in fig. 66(c), in contrast to curve (1) for the normal SrM.
On the other hand, Kreber and Gonser (1973) concluded from the M6ssbauer
experiments at r o o m t e m p e r a t u r e and 77 K for the same composition, that As
ions occupy preferentially 2b sites, although they did not describe the preparation
conditions. Lotgering (1974) showed in curve (2) in fig. 66(a) the saturation
magnetization measured parallel to the easy direction, o-//, of oriented poly-
crystalline BaFe2+0.6Ti4+0.6Fe10.8019 and curve (3) for or// of BaFe2+1.0SbS+0.sFe10.sO19,
both of which behave very similarly with t e m p e r a t u r e as curve (1) obtained by
H e i m k e (1960). Lotgering reported Ti 4+ and Sb s+ to occupy 4f2 sites, which was
supported by the X-ray and neutron diffraction investigations. H e stressed the
effect of Fe 2+ on a 4f2 site on the t e m p e r a t u r e dependence of the magnetic
anisotropy for the substituted compounds, which would be caused by the effect of
an Fe 3+, Ti 4+ or Sb 5+ ion on the neighbouring 4f2 site.
As described before, Grill and H a b e r y (1974) calculated the exchange
parameters of BaM (see table 5) and mentioned that the t e m p e r a t u r e coefficient
of the magnetization at room temperature cannot be decreased without decreas-
ing the magnetization, and moreover, the saturation magnetization would only be
increased with a diamagnetic substitution on 4f2 sites, measuring the magnetic
properties of Bal-xDxPxFe12-x019 with D = K ~+, Bi 3÷ and P = Cu 2+, Ni 2+, Mn 4+ be-
tween 180 and 770 K, Grill (1974) practically proved that all the compounds have
lower ~r~, 0c and almost the same c~(o-s) values as those of pure BaM. On the other
hand, H a n e d a et al. (1974b) studied the effect of replacements of Fe 3+ in M type
compounds with (Cu 2+ + Ge4+), (Cu 2+ + Si4+) and moreover with (Cu 2+ + Vs+),
(Cu 2+ + Nb 5+) or (Cu 2+ + TaS+). They obtained ~(Br) = 0.1% per degree without
reducing other p e r m a n e n t magnetic properties and insisted that ce(Br) can be
improved by changing the temperature coefficient a(MHc) or C~(BHC) by such
substitutions, instead of improving or(O-s), as proposed by most other investigators.
Esper and Kaiser (1975) found that the anisotropy field HA, and accordingly MHc,
and also their temperature coefficients can be varied with Sb-substituted BaM by
different heat treatments. HA can be changed from 6.3 to 12.6kOe [500-
1000 kA/m] by varying the preparation conditions, and c~(MHC) was determined to
be negative for HA = 6 . 3 k O e [500kA/m] and positive for HA = 12.6kOe
[1000 kA/m]?. The results of the studies on the mechanisms of the i m p r o v e m e n t
t o~(MHc)=+ 0.28%/°C and o~(BHc)- -0.09%/°C near room temperature for the normal BaM were
estimated by Haneda et al. (1974b).
384 H. KOJIMA

of these t e m p e r a t u r e coefficients are fairly confused, but Lotgering's consideration


(Lotgering 1974) seems to be most reasonable. H e proposed that the effects of
partially existing reduced Fe 2+ ions could be considered to contribute to the
temperature change of the anisotropy with any combinations of substituted ions.

2.2.5. Substitution with anions


Q u i t e a few researches have been published on the substitution with anions in M-
type compounds. It is presumably more difficult to obtain the specimens with
definite concentration of substituted anions than in the case of cations. Moreover,
if it is possible to replace O 2- with only anions, this will dilute the ferromagnetic
ions, so that the magnetic properties would be unfavourable for most industrial
purposes. H e n c e some examples described here concern the substitution of 02-
with F 1- and some other kinds of cations for the purpose of valency compen-
sation. Because of the resemblance in their ionic radii, F 1- seems to be the only
suitable anion (see table 2). Frei et al. (1960) obtained an M-type c o m p o u n d
BaFz.2FeO-5Fe203, showing specific gravity p = 5.25 g/cm 3 and saturation mag-
netization at r o o m temperature o-s = 72 emu/g [9 x 10 -3 Wbm/kg] in comparison
with p = 5.25 g/cm 3 and O-s= 67 emu/g [8.4 x 10 -3 Wbm/kg] for BaM. F r o m these
experiments, they suggested that the Fe e+ ions are on 4f2 sites and F 1- ions near
Ba 2+ ions. C o m p o u n d s intended to give the formula BaCoxFe12-xO19-xFx with
x = 0, 0.3, 0.4, 0.45, 0.47, 0.5, 1.0 and 2.0 were prepared by Robbins and Banks
(1963) and recognized as single phase magnetoplumbite type by X-ray diffraction
for the compositions x <~ 0.47. The estimated magnetization values for randomly
oriented powders at 0 K and H = 6 k O e [4.8 x 105 A/m] were: 76, 76.5, 85, 95.2
and 100 emu/g [9.55, 9.61, 10.68, 11.96, 12.57 x 10 -5 Wbm/kg] for x = 0, 0.3, 0.4,
0.45, 0.47 respectively. They also confirmed from torque m e a s u r e m e n t s that the
cone angle of the easy magnetization was 10 ° at 77 K for x = 0.3; and 40 ° at r o o m
t e m p e r a t u r e and 58 ° between 200 and 77 K for x = 0.47.
Robbins (1962) investigated the preferential site of Co 2+, Ni 2+ and Cu 2+ in the
hexagonal compounds BaAla2019, BaGa12Oa9 and BaFe12019. Ba 2+ was replaced by
La 3+ or O 2- by F ~-, for the charge compensation in these compounds. Referring to
the results of Banks et al. (1962) regarding the same compounds we can deduce
the following tendency for the preference of bivalent ions: In the
BaMZ+Al12_xO~9_xFx system, the order of the tetrahedral preference is Co2+>
Zn2+> Ni 2+, while in the BaM2+Galz_xO19_xFx system, it is Co2+> G a 3+ > N i 2+.
These results were derived from X-ray investigations and diffuse reflectance
measurements for powder compacts. Tables 19 and 20 give lists of the lattice
parameters after Robbins (1962) for the BaA112019 and BaFe12019 systems, res-
pectively. Examples of reflectance measurements are illustrated in fig. 67, with
curve (1) for BaNi0sGa1150185F0s, curve (2) for BaCo0.sGalLsO18sF05 and curve (3)
for BaNi1.sAl10.50175Fls. Assignments of the absorption peaks shown as T or O in
the figure have been obtained from the literature values. As for
BaM2+xFe~2_,O~9_~F~, Robbins obtained M phase single crystals with a flux of a
Fe203 and NazCO3 mixture for the composition x ~< 0.5 for Ni 2+ and Co 2+, and the
concentration x = 0-2 for Cu 2+. The bivalent cations in these substituted systems
FUNDAMENTAL PROPERTIES OF H E X A G O N A L FERRITES 385

T A B L E 19
Lattice parameters of substituted Bamll2019 with M 2+ and
F 1- (Robbins 1962).

Formula x a (A) c (A)

BaAl12019 - 5.577 22.66

0.1 5.585 22.64


0.4 5.597 22.64
BaNixAla2-xO19-xFx 0.6 5.607 22.64
0.8 5.610 22.64

B aNio.sZn2A19.sO16.sF2.5 - 5.617 22.76

0.1 5.586 22.63


0.4 5.598 22.63
BaCox Al12- x O19-xF x 1.0 5.603 22.63
2.0 5.616 22.70
2.5 5.622 22.81

BaCoZnAlloO17F2 - 5.613 22.63


BaCoo.sZn2A19.5Ox6.~F2.5 - 5.616 22.70

T A B L E 20
Lattice constants of M 2+ and F 1" substituted M compounds
(Robbins 1962).

Formula x a (,~) c (A)

0.5 5.888 23.15


0.6 5.892 23.14
BaNixFe12-xO19-xFx
1.0 5.891 23.19
2.0 5.896 23.21

BaFe22+Nio.ssFels.4201s.42Fo.ss - 5.882 32.83

0.35 5.883 23.15


BaCuxFe12-xO19-xFx 0.63 5.884 23.16
1.14 5.885 23.19

0.3 5.883 23.12


0.4 5.886 23.12
BaCoxFe12-xO19-xFx
0.45 5.886 23.13
0.47 5.887 23.15
386 H. K O J I M A

8O

7o /s)

6O r~i_ / /~IT~- (21


50

t j... / r,J

I// t . . . . 1~
iX . . . .

g I

.//," /i
,~ '
(_./ i~ V1,
+%,.,d;
Io
200 400 600 800 I000
.X. ( m,u.m )
Fig. 67. Examples of the reflectance spectra for the pellets with some anion substituted M compounds
(Robbins 1962). (1) BaNi0.sGall.5Ois.sF0.5, (2) BaCo0.sGall.sO1s.sF0.5, (3) BaNii.sAl10.5017.sF1.5. T and O
are absorption peaks from relevant ions in tetrahedral or octahedral site.

also exhibited a preference for 4fl sites, as was confirmed from the results of
magnetic measurements. The magnetic properties at 0 K are listed in table 21,
where the saturation values were estimated from the o-s-T relations in a rather
higher temperature range compared with the literature value o f BaFei2019 as a
standard. Though the Curie temperatures seem to be higher than the other
reported values by about 50°C, the increase of O-s and the decrease of 0¢ with
increasing substitution ratio are clearly seen. For the higher substitution ratio of
Co 2+, x > 0 . 5 , the compound changes to a mixture of M and W phases, then
rapidly becomes single W phase with increasing x values, which was confirmed by
X-ray and magnetic torque measurements in this paper.

T A B L E 21
Magnetic properties of M 2+ and F 1- substituted M compounds (Robbins 1962).

0% at 0 K (emu/g) NB at 0 K 0c KI at R.T. (x 10 6 erg/cm 3)


Formula x [x 4~- x 10-7 Wbm/kg] (/xB) (°C) [x 105 J / m 3]

BaFez2019 - 100 20 500 3.3

BaNixFe12-xOm-xFx 0.6 113 22.50 475 2.4

0.35 108.0 21.66 460 3.1


BaCuxFe~2-xO19-xFx 0.63 113.5 22.72 450 2.9
1.14 118.0 23.72 418 -

0.3 105 21.0 470 - -


BaCoxFel2-xOx9-xFx 0.4 108.5 21.7 470- -
FUNDAMENTAL PROPERTIES OF H E X A G O N A L FERRITES 387

References
Adelsk61d, V., 1938, Arkiv Kemi. Mineral. Bottoni, G., D. Candolfo, A. Cecchetti, L.
Geol. 12A, 1. Giarda and F. Masoli, 1972, Phys. Status
Aharoni, A., 1962, Rev. Mod. Phys. 34, 227. Solidi A32, K47.
Aharoni, A. and M. Schieber, 1961, Phys. Rev. Bowman, W.S., Saturno, F.H. Norman and J.H,
123, 807. Horwood, 1969, J. Can. Ceram. Soc. 38, 1.
Ahrens, L.H., 1952, Geochim. et Cosmochim. Bozorth, R.M. and V. Kramer, I959, J. Phys. et
Acta, 2, 155. Rad. 20, 393.
Aidelberg, J., J. Flicstein and M. Schieber, Brady, L.J., 1973, J. Material Sci. 8, 993.
1974, J. Cryst. Growth, 21, 195. Braun, P.B., 1957, Philips Res. Rep. 12, 491.
Albanese, G., M. Carbucicchio and A. Deriu, Brixner, L,H., 1959, J, Amer. Chem. Soc. 81,
1974, Phys. Status Solidi A23, 351. 3841.
A16onard, R., D. Bloch and P. Boutron, 1966, Buessem, W.R. and A. Doff, 1957, The Ther-
Compt. rend. B263, 951. mal expansion of ferrites at temperatures
Aleshke-Ozhevskii, O.P., R.A. Sizev, V.P. near the Curie point, Proc. 13 Ann. Meeting,
Cheparin and I.I. Yamzin, 1968, Zh. ETF Met. Powd. Ass., Chicago, Ill., USA, II, 196.
Pis'ma, 7, 202. Burlier, C.R., 1962, J. Appl. Phys. 33, 1360.
Aleshke-Ozhevskii, O.P. and I.I. Yamzin, 1969, Bye, G.C. and C.R. Howard, 1971, J. Appl.
Zh. Eksp. Teor. Fiz. 59, 1490. Chem. Biotechnol. 21, 319.
Anderson, P.W., 1950, Phys. Rev. 79, 350. Casimir, H.B.G., J. Smit, U. Enz, J.F. Fast,
Appendino, P. and M. Montorsi, 1973, Ann. H.P.J. Wijn, E.W. Gorter, A.J.W. Duy-
Chim. (Rome) 63, 449. vesteyn, J.D. Fast and J.J. deJong, 1959, J.
Arendt, R.H., 1973, J. Solid State Chem. 8, 339. Phys. et Rad. 20, 360.
Asti, G., F. Conti and C.M. Maggi, 1968, J. Clark, A.F., W.M. Haynes, V.A. Deason and
Appl. Phys. 39, 2039. R.J. Trapani, 1976, Cryogenics, May, 267.
AuCoin, T.R., R.O. Savage and A. Tauber, Cochardt, A., 1967, J. Appl. Phys. 38, 1904.
1966, J. Appl. Phys. 37, 2908. Craik, D J . and E.H. Hill, 1977, J. Phys. Col-
Banks, E., M. Robbins and A. Tauber, 1962, loque, 38, C1-39.
J. Phys. Soc. Japan, 17, Supplement B-I, 196. Davis, R.T., 1965, Thesis, Thermal expansion
Bashkirov, Sh.Sh., A.B. Liberman and V.I. anisotropy in BaFe120~9, The Penn. State
Sinyavskii, 1975, Zh. Eksp. Teor. Fiz. 59, Univ. College of Mineral Ind.
1490. De Bitetto, D.J., 1964, J. Appl. Phys. 35, 3482.
Batti, P., 1960, Ann. Chim. (Rome) 50, 1461. Dixon, S. Jr. and M. Weiner, 1970, J. Appl.
Belov, K.P., L.I. Koroleva, R.Z. Levitin, Phys. 41, 1375.
Yu.V. Jergin and A.V. Pedko, 1965, Phys. Dixon, S. Jr., T.R. AuCoin and R.O. Savage,
Status Solidi, 12, 219. 1971, J. Appl. Phys. 42, 1732.
Beretka, J., and M.J. Ridge, 1958, J. Chem. Drews, U. and J. Jaumann, 1969, Z. angew,
Soc. A10, 2463. Phys. 26, 48.
Berger, W. and F. Pawlek, 1957, Arch. Eisen- Drofenik, M., D. Hanzel and A. Moljik, 1973,
hiittenwesen, 28, 101. J. Mater. Sci. 8, 924.
Bertaut, E.F., A. Deschamps and R. Pauthenet, Dtilken, H., F. Haberey and H.P.J. Wijn, 1969,
1959, J. Phys. et Rad. 20, 404. Z, angew. Phys. 26, 29.
Blazey, K.W., 1974, J. Appl. Phys. 45, 2273. Dullenkopf, P. and H.P.J. Wijn, 1969, Z.
Bonnenberg, D. and H.P.J. Wijn, 1968, Z. angew. Phys. 26, 22.
angew. Phys. 24, 125. Du Pr6, F.K., D.J. De Bitetto and F.G.
Borovik, E.S. and Yu.A. Mamaluy, 1963, Fiz. Brockman, 1958, J. Appl. Phys. 29, 1127.
Metal. Metalloved. 15, 300. Eldridge, D.F., 1961, J. Appl. Phys. 32, 247 S.
Borovik, E.S. and N.G. Yakovleva, 1962a, Fiz. Erchak, M. Jr., I. Fankuchen and R. Ward,
Metal. Metalloved. 13, 470. 1946, J. Amer. Chem. Soc. 68, 2085.
Borovik, E.S. and N.G. Yakovleva, 1962b, Fiz. Erchak, M. Jr. and R. Ward, 1946, J. Amer.
Metal. Metalloved. 14, 927. Chem. Soc. 68, 2093.
Borovik, E.S. and N.G. Yakovleva, 1963, Fiz. Esper, F.J. and G. Kaiser, 1972, Int. J. Mag-
Metal. Metalloved. 15, 511. netism, 3, 189.
388 H. KOJIMA

Esper, F.J. and G. Kaiser, 1975, Physica, 80B, Haneda, K. and H. Kojima, 1974, J. Amer.
116. Ceram. Soc. 57, 68.
Fahlenbrach, H. and W. Heister, 1953, Arch. Haneda, K., C. Miyakawa and H. Kojima,
Eisenhtittenw. 29, 523. 1974a, J. Amer. Ceram. Soc. 57, 354.
Florrescu, V., M. Popescu and C. Ghizdeanu, Haneda, K., C. Miyakawa and H. Kojima,
1973, Int. J. Magnetism, 5, 257. 1974b, Improvement of temperature depen-
Frei, E.H., M. Schieber and S. Strikman, 1960, dence of remanence in ferrite permanent
Phys. Rev. 118, 657. magnets, AIP Conf. Proc. 24, 770.
Fuchikami, N., 1965, J. Phys. Soc. Japan, 20, Hareyama, K., K. Kohn and K. Uematsu, 1970,
760. J. Phys. Soc. Japan, 29, 791.
Galasso, F.S., 1970, Structure and Properties of Heimke, G., 1960, J. Appl. Phys. 31, 271 S.
Inorganic Solids (Pergamon Press, Oxford) p. Heimke, G., 1962, Bet. Deut. Keram. Ges. 39,
226-234. 326.
Gemperle, R., E.V. Shtoltz and M. Zeleny, Heimke, G., 1963, Z. angew. Phys. 15, 217.
1963, Phys. Status Solidi, 3, 2015. Heinecke, U. and L. Jahn, 1972, Phys. Status
Glasser, F.P., F.W.D. Woodhams, R.E. Meads Solidi hl0, K93.
and W.G. Parker, 1972, J. Solid State Chem. Hempel, K.A. and H. Kmitta, 1971, J. Phys.
5, 255. Colloque, 32, C1-159.
Goodenough, J.B., 1956, Phys. Rev. 102, 356. Henry, N.F.M. and K. Lonsdal, 1952, Inter-
Gorter, E.W., 1954, Philips Res. Rep. 9, 295 national Tables for X-ray Crystallography
and 403. (Kynoch Press, Birmingham) p. 304.
Gorter, E.W., 1957, Proc. IEE, 104B, 255 S. Hodge, M.H., W.R. Bitler and R.C. Bradt, 1973,
Got& K., 1966, Japan. J. Appl. Phys. 5, 117. J. Amer. Ceram. Soc. 56, 497.
Goto, Y. and T. Takada, 1960, J. Amer. Ceram. Holz, A., 1970, J. Appl. Phys. 41, 1095.
Soc. 43, 150. Hoselitz, K. and R.D. Nolan, 1969, J. Phys. D,
Goto, Y. and K. Takahashi, 1971, J. Japan Soc. Ser. 2, 2, 1625.
Powder and Powder Metallurgy, 17, 193 (in Hunty, P., 1963, Fiz. Metal. Metalloved. 16, 132.
Japanese). Ichinose, N. and K. Kurihara, 1963, J. Phys. Soc.
Goto, Y. and K. Takahashi, 1973, Japan. J. Japan, 18, 1700.
Appl. Phys. 12, 948. Jahn, L. and H.G. Mfiller, 1969, Phys. Status
Goto, Y., M. Higashimoto and K. Takahashi, Solidi, 35, 723.
1974, J. Japan Soc. Powder and Powder Jonker, G.W., H.P.J. Wijn and P.B. Braun, 1954,
Metallurgy, 21, 21 (in Japanese). Philips Techn. Rev. 18, 145.
Grill, A., 1974, Int. J. Magnetism, 6, 173. Kacz6r, C., 1964, Zh. Eksperim, Teor. Fiz. 46,
Grill, A. and F. Haberey, 1974, Appl. Phys. 3, 1787.
131. Kacz6r, J., 1972, Phys. Lett., 41A, '461.
Grosser, P., 1970, Z. angew. Phys. 30, 133. Kacz6r, J. and R. Gemperle, 1960, Czech. J. Phys.
Grundy, P. J., 1965, Brit. J. Appl. Phys. 16, 409. BIO, 505.
Grundy, P.J. and S. Herd, 1973, Phys. Status Kacz6r, K. and R. Gemperle, 1961, Czech. J.
Solidi A20, 295. Phys. Bll, 510.
Haberey, F. and A. Kockel, 1976, IEEE Trans. Kahn, F.J., P.S. Pershan and J.P. Remeika, 1969,
Mag. MAG-12, 983. J. Appl. Phys. 40, 1508.
Haberey, F. and H.P.J. Wijn, 1968, Phys. Status Kittel, C., 1946, Phys. Rev. 70, 965.
Solidi, 26, 231. Kiyama, M., 1976, Bull. Chem. Soc. Japan, 49,
Haberey, F., M. Velicescu and A. Kockel, 1973, 1855.
Int. J. Mag. 5, 161. Kojima, H. and K. Got& 1962, J. Phys. Soc.
Hagner, J. and U. Heinecke, 1974, Phys. Status Japan, 17, Supplement B-l, 201.
Solidi A22, K187. Kojima, H. and K. Got& 1964, Determination of
Haneda, K. and H. Kojima, 1971, Phys. Status critical field for magnetoplumbite type oxides
Solidi A6, 259. by domain observation, Proc. Int. Conf. Magn.
Haneda, K. and H. Kojima, 1973a, J. Appl. Nottingham, p. 727.
Phys. 44, 3760. Kojima, H. and K. Got& 1965, J. Appl. Phys. 36,
Haneda, K. and H. Kojima, 1973b, Japan. J. 538.
Appl. Phys. 12, 355. Kojima, H. and K. Got& 1970, Bulletin. Res. Inst.
FUNDAMENTAL PROPERTIES OF H E X A G O N A L FERRITES 389

Sci. Meas., Tohoku University, 19, 67 (in Mones, A.H. and E. Banks, 1958, J. Phys.
Japanese). Chem. Solids, 4, 217.
Kojima, H. and C. Miyakawa, 1965, Bulletin, Mori, S., 1970, J. Phys. Soc. Japan, 28, 44.
Res. Inst. Sci. Meas., Tohoku University, 13, Moriya, T., 1960, Phys. Rev. 120, 91.
105 (in Japanese). Mountvala, A.J. and S.F. Ravitz, 1962, J.
K6moto, O. and H. Kojima, 1976, Hard Mag- Amer. Ceram. Soc. 45, 285.
netic Materials, ed., S. Iida et al. (Maruzen, N6el, L., 1947, Compt. rend. 224, 1488.
Tokyo) Voi. 3, p. 110 (in Japanese). N6el, L., 1948, Ann. Phys. (Paris) 3, 137.
Kooy, C., 1958, Philips Techn. Rev. 19, 286. Neumann, H. and H.P.J. Wijn, 1968, J. Amer.
Kooy, C. and U. Enz, 1960, Philips Res. Rep. Ceram. Soc. 51, 536.
15, 7. Oh, H.J., T. Nishikawa and M. Satou, 1978, J.
Koroleva, L.I. and L.P. Mitina, 1971, Phys. Chem. Soc. Japan, 1, 42, (in Japanese).
Status Solidi AS, K55. Okamoto, S., H. Sekizawa and S.I. Okamoto,
Kozlowski, G. and W. Zietek, 1965, J. Appl. 1975, J. Phys. Chem. Solids, 36, 591.
Phys. 36, 2162. Okazaki, C., B. Kubota and S. Mori, 1955,
Kreber, E. and U. Gonser, 1973, Appl. Phys. 1, National Techn. Rep. 1, 23 (in Japanese).
339. Okazaki, C., S. Mori and F. Kanamaru, 1961a,
Kreber, E. and U. Gonser, 1976, Appl. Phys. National Techn. Rep. 7, 21 (in Japanese).
10, 175. Okazaki, C., T. Kubota and S. Mori, 1961b, J.
Kreber, E., U. Gonser and A. Trautwen, 1975, Phys. Soc. Japan, 16, 119.
J. Phys. Chem. Solids, 36, 263. Palatnik, L.S., L.I. Lukashenko, L.Z. Luby-
Kuntsevich, S.P., Yu.A. Mamaluy and A.S. anyy, V.I. Lukashenko and Yu.A. Mamaluy,
Miln6r, 1968, Fiz. Metal. Metalloved. 26, 610. 1975, Fiz. Metal. Metalloved. 40, 61.
Kurtin, S., S. Foner and B. Lax, 1969, J. Appl. Pauling, L., 1960, The Nature of the Chemical
Phys. 40, 818. Bond, 3rd ed. (Cornell Univ. Press, Ithaca,
Lacour, C. and M. Paulus, 1968, J. Cryst. N.Y.) p. 514.
Growth, 3, 814. Paulus, M., 1960, Compt. rend. 250, 2332.
Lacour, C. and M. Paulus, 1975, Phys. Status Pauthenet, R. and G. Rimet, 1959a, Compt.
Solidi A27, 441. rend. 249, 1875.
Laroria, K.K. and A.P.B. Shinha, 1963, J. Pure Pauthenet, R. and G. Rimet, 1959b, Compt.
and Appl. Phys. 1, 215. rend. 249, 656.
Linares, R.C., 1962, J. Amer. Ceram. Soc. 45, Perekalina, T.M. and V.P. Cheparin, 1967, Fiz.
307. Tverd. Tela. 9, 217.
Lotgering, F.K., 1974, J. Phys. Chem. Solids, Perekalina, T.M., M.A. Vinnik, R.I. Zevereva
35, 1633. and A.P. Schurova, 1970, Zh. Eksp. Teor. Fiz.
Luborsky, F.E., 1966, J. Appl. Phys. 37, 1091. 59, 1490.
Ltitgemeier, M., A.V. Gonzales, W. Zietek and Rathenau, G.W., 1953, Rev. Mod. Phys. 25,
W. Zinn, 1977, Physica, 86--88B, 1363. 297.
MacChesney, J.B., R.C. Sherwood, E.T. Keve, Rathenau, G.W., J. Smit and A.L. Stuyts, 1952,
P.B. O'Connor and L.D. Blitzer, 1971, New Z. Phys. 133, 250.
room temperature ferrimagnetic phases in Ratnam, D.V. and W.R. Buessem, 1970, IEEE
the system CaO-AI203-Fe203, Ferrites, Proc. Trans. Magn. MAG-6, 610.
Int. Conf. Japan, p. 158. Ratnam, D.V. and W.R. Buessem, 1972, J.
Mahoney, J.P., A. Tauber and R.O. Savage, Appl. Phys. 43, 1291.
1972, Low temperature magnetic properties Reddy, B.P.N. and P.J. Reddy, 1973, Phys.
in BaFe12-x-y-zZnxTiyMnzO19, AIP Conf. Status Solidi A17, 589.
Proc. no. 10, part 1, p. 159. Reddy, B.P.N. and P.J. Reddy, 1974a, Phil.
Mason, W.P., 1954, Phys. Rev. 96, 302. Mag. 30, 557.
Mee, C,D. and J.C. Jeschke, 1963, J. Appl. Reddy, B.P.N. and P.J. Reddy, 1974b, Phys.
Phys. 34, 1271. Status Solidi A22, 219.
Menashi, W.P., T.R. AuCoin, J.R. Shappirio Reed, J.S. and R.M. Fulrath, 1973, J. Amer.
and D.W. Eckart, 1973, J. Cryst. Growth, 20, Ceram. Soc. 56, 207.
68. Rensen, I.G. and I.S. van Wieringen, 1969,
Mita, M., 1963, J. Phys. Soc. Japan, 18, 155. Solid State Commun. 1, 1139.
390 H. KOJIMA

Rensen, J.G., J.A. Schulkes and J.S. van Smit, J. and H.G. Beljers, 1955, Philips Res.
Wieringen, 1971, J. Phys. Colloque, 32, Rep. 10, 113.
C1-924. Smit, J. and H.P.J. Wijn, 1959, Ferrites (Philips
Richter, H.G. and H.E. Dietrich, 1968, IEEE Technical Library, Eindhoven) p. 177-210.
Trans. Magn. MAG-4, 263. Smit, J., F.K. Lotgering and U. Enz, 1960, J.
Robbins, M., 1962, Thesis, Fluoride-compen- Appl. Phys. Supplement, 31, 137 S.
sated cation substitution in oxides, Polytech- St~iblein, H., 1973, Z. Werkstofftechnik, 4, 133.
nic Inst. Brooklyn, New York. St~iblein, H. and W. May, 1969, Ber. Deut.
Robbins, M. and E. Banks, 1963, J. Appl. Phys. Keram. Geselschaft, 46, 69.
34, 1260. Standley, K.J., 1972, Oxide Magnetic Materials
Rodrigue, G.P., 1963, IEEE Trans. Microwave (Clarendon Press, Oxford) p. 149-160.
Theory and Techniques, September, 351. Stoner, E.C. and E.P. Wohlfarth, 1948, Phil.
Roos, W., H. Haak, C. Voigt and K.A. Hem- Trans. A240, 599.
pel, 1977, J. Phys. Colloque, 38, C1-35. Streever, R.S., 1969, Phys. Rev. 186, 285.
Rosenberg, M., C. Tfinfisoiu and V. Florescu, Suchet, J.P., 1971, Compt. rend. S6ri C, 271,
1966, J. Appl. Phys. 37, 3826. 895.
Rosenberg, M., C. Tanasoiu and V. Florescu, Suemune, Y., 1972, J. Phys. Soc. Japan, 33, 279.
1967, Phys. Status Solidi, 21, 197. Summergrad, R.N. and E. Banks, 1957, J. Phys.
Routil, R.J. and D. Barham, 1974, Can. J. Chem. Solids 2, 312.
Chem. 52, 3235. Szymczak, R., 1971, J. Phys, Colloque, 32,
Sadler, A.G., 1965, J. Can Ceram. Soc. 34, 155. C1-263.
Savage, R.O. and A. Tauber, 1964, J. Amer. Takada, T., Y. Ikeda and Y. Bando, 1971, A
Ceram. Soc. 47, 13. new preparation method of the oriented fer-
Schieber, M.M., 1967, Experimental Mag- rite magnets, Ferrites, Proc. Int. Conf. (Univ.
netochemistry in Series of Monographs on Tokyo Press, Tokyo) p. 275.
Selected Topics in Solid State Physics, ed. Tfinfisoiu, C., 1972, IEEE Trans. Magn. MAG-
E.P. Wohlfarth (North-Holland, Amsterdam) 8, 348.
v01. 8, p. 198-217. Tfinfisoiu, C., P. Nicolau and C. Micea, 1976,
Shimada, Y., K. Got6 and H. Kojima, 1973, IEEE Trans. Magn. MAG-12, 980.
Phys. Status Solidi A18, K1. Tauber, A., R.O. Savage, R.J. Gambino and
Shimizu, S. and K. Fukami, 1972, J. Japan Soc. C.G. Whinfrey, 1962, J. Appl. Phys. 33, 1381.
Powder and Powder Metallurgy, 18, 259 (in Tauber, A., J.A. Kohn and I. Bandy, 1963, J.
Japanese). Appl. Phys. 34, 1265.
Shirk, B.T. and W.R. Buessem, 1969, J. Appl. Tauber, A., S. Dixon, Jr. and R.O. Savage, Jr.,
Phys. 40, 1294. 1964, J. Appl. Phys. 35, 1008.
Shirk, B.T. and W.R. Buessem, 1970, J. Amer. Tebble, R.S. and D.J. Craik, 1969, Magnetic
Ceram. Soc. 53, 192. Materials (Wiley-Interscience, London) p.
Shirk, B. and W.R. Buessem, 1971, IEEE 359-368.
Trans. Magn. MAG-7, 659. Townes, W.D., J.H. Fang and A.J. Perrotta,
Sigal, M.A., 1977, Phys. Status Solidi A42, 1967, Z. Kristallogr. 125, 437.
775. Trautwein, A., E. Kreber, U. Gonser and F.E.
Silber, L.M., E. Tsantes and P. Angelo, 1967, J. Harris, 1975, J. Phys. Chem. Solids, 36, 325.
Appl. Phys. 38, 5315. Van Diepen, A.M. and F.K. Lotgering, 1974, J.
Sixtus, K.J., K.J. Kronenberg and R.K. Tenzer, Phys. Chem. Solids, 35, 1641.
1956, J. Appl. Phys. 27, 1051. Van Hook, H.J., 1964, J. Amer. Ceram. Soc.
Sloccari, G., 1973, J. Amer. Ceram. Soc. 56, 47, 579.
489. Van Loef, J.J. and P.J.M. Franssen, 1963, Phys.
Slokar, G. and E. Lucchini, 1978a, J. Mag. Lett. 7, 225.
and Mag. Mat. 8, 232. Van Loef, J.J. and A.B. van Groenou, 1964,
Slokar, G. and E. Lucchini, 1978b, J. Mag. On the sub-lattice magnetization of BaFe12019,
and Mag. Mat. 8, 237. Proc. Int. Conf. Magn. Nottingham, p. 646.
Smit, J., 1959, J. Phys. et Rad. 20, 370 (see Van Uitert, L.G., 1957, J. Appl. Phys. 28,
Casimir et al. 1959). 317.
FUNDAMENTAL PROPERTIES OF H E X A G O N A L FERRITES 391

Van Uitert, L.G. and F.W. Swanekamp, 1957, Wijn, H.P.J., 1952, Nature, 170, 707.
J. Appl. Phys. 28, 482. Wijn, H.P.J., 1970, Landolt-B6rnstein Numeri-
Van Wieringen, J.S., 1967, Philips Tech. Rev. cal Data and Functional Relationships in
28, 33. Science and Technology. New Series, III/4b.
Van Wieringen, J.S. and J.G. Rensen, 1966, Z. ed. K.-H. Hellwege (Springer, Berlin) p. 552-
angew. Phys. 21, 69. 681.
Verstegen, J.M.P.J., 1973, J. Solid State Chem- Williams, H.J. and R.C. Sherwood, 1958, J.
istry, 7, 468. Appl. Phys. 29, 296.
Villers, G., 1959a, Compt. rend. 248, 2973. Wisnyi, L.G., 1967, Powder Diffraction File,
Villers, G., 1959b, Compt. rend. 249, 1337. 7-85.
Vinnik, M.A. and R.I. Zvereva, 1969, Kristal- Wullkopf, H., 1972, International J. Magnetism,
lografiya, 14, 697. 3, 179.
Wang, F.F.Y., K. Ishii and B.Y. Tsui, 1961, J. Wullkopf, H., 1975, Physica, 80B, 129.
Appl. Phys. 32, 1621. Zanmarchi, G. and P.F. Bongers, 1969, J. Appl.
Wells, R.G. and D.V. Ratnum, 1971, IEEE Phys. 40, 1230.
Trans. Magn. MAG-7, 651. Zfiveta, K., 1963, Phys. Status Solidi, 3, 2111.
Went, J.J., G.W. Rathenau, E.W. Gorter and Zinn, W., S. Hfifner, M. Kalvius, P. Kienle and
G.W. van Oosterhout, 1951/1952, Philips W. Wiedemann, 1964, Z. angew. Phys. 17,
Techn. Rev. 13, 194. 147.
chapter 6

PROPERTIES OF
FERROXPLANA-TYPE
HEXAGONAL FERRITES

M. SUGIMOTO
Saitama University, Faculty of Engineering
225 Shimo-ohkubo, Urawa 338
Japan

Ferromagnetic Materials, Vol. 3


Edited by E.P. Wohlfarth
© North-HollandPublishing Company, 1982

393
CONTENTS

1. C h e m i c a l c o m p o s i t i o n s , crystal s t r u c t u r e a n d spin o r i e n t a t i o n . . . . . . . . 395


1.1. BaMezFe16Oz7 ( W - t y p e ) . . . . . . . . . . . . . . . . . . . . 395
1.2. BaeMeeFel~Oz2 ( Y - t y p e ) . . . . . . . . . . . . . . . . . . . . 397
1.3. Ba3MeeFe24041 ( Z - t y p e ) . . . . . . . . . . . . . . . . . . . . 399
1.4. BazMe2Fe28046 ( X - t y p e ) . . . . . . . . . . . . . . . . . . . . 400
1.5. Ba4MezFe36060 ( U - t y p e ) . . . . . . . . . . . . . . . . . . . . 401
1.6. O t h e r c o m p o u n d s . . . . . . . . . . . . . . . . . . . . . . 401
2. P r e p a r a t i o n ~nd f o r m a t i o n kinetics . . . . . . . . . . . . . . . . . 402
3. SatuJ atic n r~ a g n e | i z a t i o n . . . . . : . . . . . . . . . . . . . . . 4O4
4. M a g n e t o c r ~ stalline a n i s o t r o p y a n d r e l a l e d p h e n c m e n a . . . . . . . . . . . 412
5. M6ssb~ u e r effect . . . . . . . . . . . . . . . . . . . . . . . . 421
6. M a g n e l ostriction a n d N M R . . . . . . . . . . . . . . . . . . . . 425
6.1. M a g n e t o s l r i c t i o n . . . . . . . . . . . . . . . . . . . . . . 425
6.2. N M R . . . . . . . . . . . . . . . . . . . . . . . . . . 427
7. H i g h f r e q u e n c y m a g n e t i c lzrol:crties . . . . . . . . . . . . . . . . . 428
8. E l e c t r i c p r o r erties a n d o l h e r effects . . . . . . . . . . . . . . . . . 434
8.1. Cc n d u c l i v i l y . . . . . . . . . . . . . . . . . . . . . . . . 434
8.2. E ieleclric c c n s t a n t . . . . . . . . . . . . . . . . . . . . . 435
8.3. J a h r - T e l l e r effect . . . . . . . . . . . . . . . . . . . . . . 435
8.4. M a g n e t o - o p t k al effect . . . . . . . . . . . . . . . . . . . . 436
8.5. E ' o m a ! n c bse r v a l i o n a n d c h e m ; c a l analysis . . . . . . . . . . . . . 436
References . . . . . . . . . . . . . . . . . . . . . . . . . . . 439

394
1. Chemical compositions, crystal structures and spin orientation

A group of compounds consisting of isotropic materials with spinel structure and


higher anisotropic materials with hexagonal structure has been developed by
Jonker et al. (1956/57). They are generically referred to as the ferroxplana-type
compounds. The general chemical formula of these compounds is denoted by
m(Ba2++ Me2+)-nFe203, where Me 2+ represents the divalent metal ions Mn, Fe,
Co, Ni, Cu, Mg and Zn. In the triangle of fig. 1 in ch. 5 the symbols W, Y, Z, X
and U represent the compounds with chemical composition BaMezFe16027,
Ba2MezFe12022, Ba3Me2Fe24041, BazMezFe28046 and Ba4Me2Fe36060, respectively. If
Me 2+ ions in the W-structure, as an example, are substituted by Zn 2+ ions, the
composition may be conveniently indicated by the short notation of Zn2W. In the
case of substitution of both Zn 2+ and Fe 2+ ions for Me 2+ ions it may be
represented by ZnFeW.

1.1. BaMe2Fe16027 (W-type)

The unit cell of the W-type compound is built up by superposition of four spinel
blocks (S-block) and two blocks containing Ba ions (R-block) as shown in table 1
of ch. 5. Figure 1 shows a cross section of the W-structure having a hexagonal
packing, which is closely related to the M-structure (Albanes e et al. 1976b). The
only difference is that the successive R-blocks are interspaced by two S-blocks
instead of one, as is the case in the M-structure. The crystal structure of R-blocks
with chemical composition BaFe6Oll as well as that of S-blocks with chemical
composition Me2Fe408 are represented in fig. 14 of ch. 5.
In table 1 the number of ions and the coordination of the different cation
sublattices in the W-structure are shown (Albanese et al. 1976b). The cations
occupy seven different sublattices of 12K, 4e, 4fw, 4fw, 6g, 4f and 2d, in the
nomenclature used by Braun (1957). The spin orientation according to the
generally accepted collinear model is also indicated. This magnetic classification
has been justified by the assumption that the magnetic behaviour of the nearest
neighbour cations is determined by superexchange interaction.

395
396 M. S U G I M O T O

W- type Structure

4fly
® 4e
© 4f
e6g
e 12K
4fv I
• 2d
R
©o 2-
0 Ba 2+
I S*
I S*

R*

Fig. 1. Unit cell of the BaMe2Fe16027, Me2W, compound. T h e anions of O 2 , the divalent barium
cation Ba 2÷, the metallic ions in the sublattices 4fw, 4e, 4f, 6g, 12K, 4fvi and 2d are indicated. T h e
coordination figures of the metallic ions in the different lattice sites are shown (Albanese et al. 1976b).

TABLE 1
N u m b e r of ions, coordination and spin orientation for the various cations of a W-type
c o m p o u n d (Albanese et al. 1976b).

N u m b e r of
Magnetic ions per
sublattice Sublattice Coordination formula unit Block Spin

K 12K octahedral 6 R-S up


fw 4e tetrahedral 2 S down
4fry tetrahedral 2 S down
fvI 4fvi octahedral 2 R down
a 6g octahedral 3 S-S up
4f octahedral 2 S up
b 2d hexahedral 1 R up
PROPERTIES OF FERROXPLANA-TYPE HEXAGONAL FERRITES 397

1.2. Ba=Me2Fe12022(Y-type)

T h e u n i t cell of t h e Y - t y p e c o m p o u n d is b u i l t u p b y t h e s u p e r p o s i t i o n of t h r e e
S - b l o c k s a n d t h e so c a l l e d t h r e e T - b l o c k s as s h o w n in fig. 2, in w h i c h t h e d i f f e r e n t

Y - type Structure

S
3 oVl
(I) 6cvl
3 bvi
o 1Bhvt
T o 6Civ
® 6clv
O 0 2-
• BCl 2+

Fig. 2. Unit cell of the Ba2Me2Fe12Oz~, Me2Y, compound. The anions of O 2-, the divalent barium
cation Ba 2÷, the metallic ions in the sublattices 3avi, 6cvl, 3bvl, 18hw, 6cw and 6c~v are indicated. The
coordination figures of the metallic ions in the different lattice sites, together with their spin
orientation, are shown (Albanese et al. 1975b).
398 M. S U G I M O T O

lattice sites are distinguished by different symbols (Albanese et al. 1975). The
Y-structure has the crystal symmetry characterized by the space group R3m. As
shown in fig. 14 of ch. 5, the T-block with Ba2Fe8014 composition is formed
by four oxygen layers having hexagonal packing and plays an unique role in the ,

TABLE 2
Number of ions, coordination and spin orientation for the various
metallic sublattices of Y-structure? (Albanese et al. 1975b).

Number of
ions per
Sublattice Coordination Block unit cell Spin

6cw tetrahedral S 6 down


3aw octahedral S 3 up
18hvi octahedral S-T 18 up
6cw octahedral T 6 down
6c•v tetrahedral T 6 down
3by1 octahedral T 3 up

t Sublattices having the same crystalline symmetry but belonging to


different blocks are marked by an asterisk.

TABLE 3
Strength of the superexchange interactions between Fe 3+ ions in the
Y-structure (Albanese et al. 1975b).

Interacts with n Fe 3+ ions Strength of the


Each Fe 3+ ion in superexchange
lattice position n lattice position interaction1

3avi 6 6cw 25
6 18hvl 0(+)
6Civ 3 6cIv 0(+)
3 3avi 25
9 18hw 26
18hvl 1 3avr 0(+)
3 6clv 26
2 6Cvi 30
1 6C~v 25
4 18hvi 0(+)
0Cvi 6 18hvl 30
3 6CTv 9(+)
1 3bvi 1.7
6C}'v 3 6cw 9(+)
3 18hvl 25
3 3bvI 77
3bvl 6 6c•v 77
2 6Cvi 1.7

t The cross indicates interactions between sublattices with parallel spins.


PROPERTIES OF FERROXPLANA-TYPE HEXAGONAL FERRITES 399

Y - s t r u c t u r e . In fig. 2, t h e c o m m o n faces of t h e o c t a h e d r a inside t h e T - b l o c k a r e


h a t c h e d . T h e p r e s e n c e of a n i o n o c t a h e d r a with c o m m o n faces is g e n e r a l l y
r e s p o n s i b l e for t h e l o w e r stability of t h e s t r u c t u r e , d u e to t h e h i g h e r p o t e n t i a l
e n e r g y of t h e s y s t e m as c o m p a r e d with s i t u a t i o n s w h e r e o n l y c o r n e r s a r e s h a r e d
a n d t h e c a t i o n s a r e thus f u r t h e r a p a r t . This fact f a v o u r s t h e M e 2+ ions having a
m a r k e d p r e f e r e n c e for t h e o c t a h e d r a l c o o r d i n a t i o n of t h e 6Cvi a n d 3bw lattice
sites.
T a b l e 2 shows t h e spin o r i e n t a t i o n for v a r i o u s s u b l a t t i c e s t o g e t h e r with t h e i r
c o o r d i n a t i o n in t h e Y - s t r u c t u r e . A l b a n e s e et al. (1975b) c a l c u l a t e d t h e s t r e n g t h of
t h e v a r i o u s s u p e r e x c h a n g e i n t e r a c t i o n s b e t w e e n F e 3+ ions in t h e Y - s t r u c t u r e .
T a b l e 3 shows t h e results o b t a i n e d on t h e basis of t h e a s s u m p t i o n t h a t t h e
i n t e r a c t i o n e n e r g y follows an e x p o n e n t i a l d e p e n d e n c e on t h e a n i o n - c a t i o n dis-
t a n c e s a n d a c o s 2 0 law for t h e a n g u l a r d e p e n d e n c e . F r o m this t a b l e it a p p e a r s
t h a t t h e s t r o n g e s t s u p e r e x c h a n g e i n t e r a c t i o n is t h e o n e b e t w e e n t h e t e t r a h e d r a l
ions 6Cfv a n d t h e o c t a h e d r a l ions 3bvi inside t h e T - b l o c k , a n d t h e o n l y a p p r e c i a b l e
p e r t u r b i n g i n t e r a c t i o n a p p e a r s to b e t h e o n e b e t w e e n t h e 6C?v a n d 6Cvr ions, b o t h
b e l o n g i n g to t h e T - b l o c k .

1..3. Ba3Me2Fe24041 (Z-type)

T h e c r y s t a l l i n e s t r u c t u r e of t h e Z - t y p e c o m p o u n d s is s h o w n in fig. 3 ( A l b a n e s e et
al. 1976a). T h e unit cell is f o r m e d by the s u p e r p o s i t i o n of f o u r S-blocks, two
T-blocks and one R-block, and the divalent and trivalent cations are distributed
a m o n g ten different lattice sites. T a b l e 4 shows t h e n u m b e r of ions r e l a t i v e to t h e
v a r i o u s c a t i o n sublattices t o g e t h e r with t h e i r spin o r i e n t a t i o n .

TABLE 4
Number of ions, coordination and spin orientation of the various
metallic sublattices of Z-structuret (Albanese et al. 1976a).

Number of
ions per
Sublattice Coordination Block unit cell Spin

12kv~ octahedral R-S 12 up


2dv five-fold R 2 up
4fw octahedral R 4 down
4f-}i octahedral S 4 up
4etv tetrahedral S 4 down
4fiv tetrahedral S 4 down
12k~}i octahedral T-S 12 up
2avl octahedral T 2 up
4evl octahedral T 4 down
4f]~v tetrahedral T 4 down

t Sublattices having the same crystalline symmetry but belonging to


different blocks are marked by an asterisk.
400 M. SUGIMOTO

Z - type S t r u c t u r e

© 12Kvx
R ® 2dv
4f-Vl
® 4f~ 1
@ 4elv
S ~ 4fly
12K~,I
I) 2clvl
• 4evl
T © 4fly
© 0 z-
BCI 2÷

R.

T*

I S*
Fig. 3. Unit cell of Ba3Co2Fe24041, Co2Z, compound. The coordination figures of the metallic ions in
the main lattice sites are shown (Albanese et al. 1976a).

1.4. Ba2Me2Fez8046 (X-type)

T h e u n i t cell of the X-type c o m p o u n d s consists of f o u r a l t e r n a t e layers of the


M - s t r u c t u r e a n d W - s t r u c t u r e , b e l o n g i n g to the space s y m m e t r y g r o u p R3m.
F i g u r e 4 shows a cross section of the Zn2X structure, in which the spin o r i e n t a t i o n
PROPERTIES OF FERROXPLANA-TYPE HE XAGONAL FERRITES 401

X - type S t r u c t u r e
f~

Y3~_
(

A /X A ZX

Q -- 0 2 - @ -- BQ2÷

O - Oct. • - Tet. [] - Tri.

Fig. 4. A collinear spin model for Zn~X (Tauber et al. 1970).

as well as the coordination figures of all cations are represented (Tauber et al.
1970).

1.5. Ba4Me2Fe3606o (U-type)

The unit--cell of the U-type c o m p o u n d is a rhombohedral structure belonging to


space group R3m, and is formed by three molecules. The structure is built up by
the superposition of two M-blocks and one Y-block along the c-axis. Figure 5
shows a cross section of the Zn2U structure and the spin orientation of all cations
(Kerecman etoal. 1968). Referring to hexagonal structure, its lattice p a r a m e t e r s
are c = 113.2 A, a = 5.88 A and the X-ray density is 5.36.

1.6. Other compounds

Kohn et al. (1964a,b) reported a new hexagonal ferrite with Ba4Zn2Fes2Os4 com-
position. T h e new structure is made up by the sequence of composite TS- and
RS-block. Interleaving the TS- and RS-block at various ratios leads to a structure
with unit cells ranging from 18 to 138 oxygen layers; these give rise to hexagonal
unit cells with c parameters ranging up to about 1600 A. The expected empirical
formulae for this group of compounds are listed in table 5. Kohn and Eckart
(1971) discovered a new c o m p o u n d w~th composition, BasZn2Ti3Fe12031, which is a
hexagonal structure with a = 5.844 A and c =-43.020 A and contains 18 oxygen
layers, indicated by the symbol Zn2-18H.
402 M. SUGIMOTO

U type Structure
-

( )

0
_
-
0 2-
Oct. •
_
-
A

Be2*
Tet.
iA

[]- Tri.
M

Fig. 5. A collinear spin model for Zn2U (Kerecman et al. 1968).

TABLE 5
Chemical compositions and crystallographic properties for new ferroxplana-type compounds (Robert et
al. 1964).

Structural Number of Lattice


unit oxygen parameter Primitive Space
Chemical composition (blocks) layers c (A) symmetry group

Bal0ZnaFes60102 [(TS)4T]3 28 x 3 203 Rhombohedral P63/mmc


Ba12ZnmFe680124 [(TS)sT]3 34 x 3 247 Rhombohedral
Ba14Zni2Fes00146 (TS)6T 40 97 Hexagonal P63/mmc
Ba16Zn14Fe920168 [(TS)TT]3 46 x 3 334 Rhombohedral
Ba4Zn2Fe52084 [RS2(RS)3]3 22 × 3 154 Rhombohedral R3m

2. Preparation and formation kinetics

The processes for producing the ferroxplana-type compounds are very similar to
those for M-type compounds. However, more accurate procedures based on the
phase equilibrium are necessary to obtain the ferroxplana-type compounds,
PROPERTIES OF FERROXPLANA-TYPE HEXAGONAL FERRITES 403
..
because their chemical compositions are"very complex. In particular, it is very
difficult to produce a crystallographicall'y pure compound containing various
amounts of ferrous iron. Neuman and Wijn (1968) shed light on the chemical
equilibrium between the oxygen partial pressure o f gas atmosphere and the
formation of Fe2W phase. In order to obtain a homogeneous W-compound, the
samples must be sintered at 1250° to 1400°C in an atmosphere with a partial
oxygen pressure between 2 x 10 .4 and 2 x 10-1 atm.
Figure 6 shows the formation temperature and stability range of the phases for
the W, Y and Z compounds. The spinel phase appeared as the first major reaction
product of the raw oxides at about 555°C. The M phase was detected next by X-ray
analysis and followed by the formation of a Y, Z, W phase in turn. Castelliz et al.
(1969) also studied the kinetics of phase formation as well as the stability of the
phases.
Lotgering (1959) evolved a new method for making a sintered crystal-oriented
ferroxplana compound. The advantage of crystal orientation is evident from the
fact that the permeability of its compounds can be about 3 times larger than that
of non-oriented compounds. This method differs essentially in the formation
mechanism from that for the M-type compound. A paste or thick suspension
consisting of BaFe12019 powder and raw oxides such as ZnO or CoO is poured
into a die and then introduced into a static magnetic field. The crystal orientation
is made topotactically by compressing the suspension into a pellet. The orien-
tation preserved through the firing at 1100° to 1300°C. Licci and Asti (1979) tried
to produce topotactically the crystal-oriented CoZnY compound. The hot-press-
ing method was performed to obtain a crystal-oriented Co2W compound by
Okazaki and Igarashi (1970).
A large and nearly perfect single crystal of the ferroxplana-type compounds,

t i [ I F i [ i r J

0.8
Z

o.z,
.=- \/ / \
0.2

rr Co 4

~;oo 600 Boo ~ooo ~2oo ~oo


T('C)
Fig. 6. F o r m a t i o n temperature a n d s t a b i l i t y r a n g e o f W - , Y-, Z - a n d M - p h a s e s (Neckenbiirger et al.
1964).
404 M. SUGIMOTO

which is useful for microwave devices, can be grown relatively easily by the flux
method. Tauber et al. (1962, 1964) investigated many kinds of flux materials useful
for growing single crystals of ferroxplana compounds. It was found that BaO-
B203 flUX must be less volatile and less viscous than the NaFeO2 flux to obtain a
single crystal with lower ferromagnetic resonance linewidth. In general, the W-
and Y-type compounds are easily melted at a lower temperature. The difficulty in
growing the crystals of Z-, X-, and U-type compounds may be attributable to the
high-melting composition necessitated by the large concentration of Fe203.
Stearns et al. (1975) and Glass et al. (1980) have grown single crystal films of ZnzY
or Zn2W by the isothermal dipping method of liquid phase epitaxy using a
PbO-BaO-B203 flux.
Many investigations on microstructures of sintered samples have been presented:
Huijser-Gerits and Rieck (1970, 1974, 1976) studied thoroughly the influence of
sintering conditions on microstructure; Drobek et al. (1961) observed the micro-
structures by the Use of electron microscope; both Cook (1967) as well as Landuyt
and Amelinckx (1974) observed the stacking sequence by electron microscope.

3. Saturation magnetization

Many attempts have been made to improve the saturation magnetization by the
substitutions of various kinds of metal ions for cations occupying the octahedral
and tetrahedral sites in the oxygen framework structure of the ferroxplana-type
compounds. This may be attributed to their unique crystallographical structure as
well as their importance as promising materials for technological application in
the field of permanent magnets and microwave devices.
Smit and Wijn (1959) proposed a formula on the basis of the over-simplification
that the number of Bohr magnetons at saturation for the W-type compound is
simply equal to the sum of the corresponding number for M- and S-structures, i.e.:

(ns)w = (nB)M+ 2(nn)s. (1)

This concept implies that we treat as different the two S-blocks which are
perfectly equivalent. This drastic consequence frequently leads to a discrepancy
with the experimental values. For example, the formula gives the value of
(ns)w = 20/XB for the ZnzW compound, while Albanese et al. (1976a) and Savage
and Tauber (1965) determined it experimentally as 35/x8 and 38.2/XB at 0 K
(= 123 G cm3/g and 134 G cm3/g, respectively). The assumption of Smit and Wijn is
applicable to the Y-type and Z-type compounds, but problems slightly analogous to
that for the W-type compound still remain.
The W-structure is characterized by the presence of two additional spinel blocks
instead of one, as in the M-structure. Such a structure creates the possibility of
changing the magnetic properties by a suitable substitution of the cation. Uitert
and Swanekamp (1957) attempted to improve the saturation magnetization of
W-type compounds by the substitution of non-magnetic ions for cations in
PROPERTIES OF FERROXPLANA-TYPE H E X A G O N A L FERRITES 405

tetrahedral and octahedral sites, and showed that the saturation magnetization is
generally apt to decrease with increasing amount of substitution. In fig. 7, zinc
ions seem likely to occupy the tetrahedral sites in BaMe2Fe16027 and a much
greater fraction of A1, G a and In ions appears to occupy the octahedral sites. An
anomalous behaviour of curve (1) at around zero saturation can be attributed to a
lack of homogeneity in the samples. Albanese et al. (1977) reported that the
saturation magnetization of BaZn2AlxFe16_xO27 is reduced when the amount of
substitution of AP + for Fe 3+ in a-sites is increased, and a compensation point of
superexchange interaction results at x = 4.
In fig. 8 the saturation magnetization, o-s, for a number of simple and mixed
W-type compounds is plotted as a function of temperature. It appears from the
figure that almost straight lines are found over a large temperature range, and that
zinc ions give higher saturation magnetization at low temperatures. In Zn2W
compounds, the Z n 2+ ions may occupy two tetrahedral sublattices 4e and 4f~v
(Albanese et al. 1976b), and in Mg2W compounds 90% of the Mg 2+ ions may

4000

E
-x 3000 7
\ ,~ \ \
g

"~ 2000
w

I/1
3"
\
i1

.~ I000

E
- 1000
2 6 B

Fig. 7. Effect of various substitutions on room-temperature saturation magnetization for BaNi2Fe16OzT.


X denotes the number of metal ions replaced per formula unit. (1) AI for Fe, (2) Ga for Fe, (3) In or Cr for
Fe, (4) Zn for Ni, (5) BaZn2GaFelsOz7, (6) BaZn2Ga3Fe13027 and (7) BaZn2AI3Fe13027 (Van Uitert and
Swanekamp 1957).
406 M. S U G I M O T O

120 I i I I t I '

100 ~ ' ~ "


, "'-Lx,.
~ Me2= ZnFe~ Me 2 W

~' ao . 0Yg
N
o
"k
I.-~ 60
'4"
F

LO Ni F' """'/-c~'~'-.XX~D"'~IL~_

2o

0 l J 1 I r , ,a,~ ~a,z
-273 -200 0 200 400 600
T(°C)
Fig. 8. Saturation magnetization as a function of temperature for a n u m b e r of c o m p o u n d s with
W-structure, m e a s u r e d on polycrystalline specimens at a field of 6600 Oe (Smit and Wijn 1959).

,Zn
I i I

60

MeY
E

40

x
/Mg .d

E 20 Ni

0 I I I I

-273 -200 0 200 400


T [°0)
Fig. 9. Saturation magnetization as a function of temperature for a n u m b e r of c o m p o u n d s with
Y-structure measured on polycrystalline specimens at a field of 11000 Oe (Smit and Wijn 1959).
PROPERTIES OF FERROXPLANA-TYPE H E X A G O N A L FERRITES 407

occupy the octahedral sites, and 10% of Mg the tetrahedral sites (Smit and Wijn
1959). Many other reports on the saturation magnetization of W-type compounds
have been presented. A study on Fe2W substituted by Ni 2+ and F- (Banks et al.
1962), Fe2W substituted by Co 2+ and F- (Robbins et al. 1963), (NiZn)W and
(Nil.6Co0.4)W (Hodges et al. 1964) and (Co165Fe0.35)W (Yamzin et al. 1966).
An expected improvement in magnetic moment in the Y-type compounds was
proposed by Albanese et al. (1975b). If all cations occupying 3by1 sites in a T-block
can be substituted by non-magnetic ions, the magnetic moment might be mar-
kedly improved. This is due to the fact that 3bvi sites alone link the upper and
lower parts of the unit cell through the strong interaction with six ions 6cw.
Furthermore, the inversion symmetry around 3bvi sites might be broken by the
partial substitution of iron ions in 6Cvi sites. However, such a drastic change in the
magnetic order, has so far not been reported. Figure 9 shows the saturation
magnetization of a number of simple Y-type compounds as a function of tem-
perature. In the case of Y-type compounds, zinc ions also give the highest
saturation magnetization. Albanese et al. (1975b) reported that in
Ba2Mg2Fe~2022(Mg2Y), Mg2+ ions mainly occupy 3bvi and 6Cw sublattices of the
tetragonal sites inside the T-block, and this occupation leads to the weakening
effect of the superexchange interaction caused by a critical competition of the two
exchange interaction 3bvr-6CTv and 6Cvr-6C•v. The temperature dependence of the
saturation magnetization parallel and perpendicular to the c-axis of
(Ba0.05Sr0.95Zn)Y single crystals is shown in fig. 10 (Enz 1961). We can see from
these curves that this compound has a preferential plane. According to Albanese
et al. (1975b), since the Co 2+ ions in Ba2Co2Fe12022(Co2Y) have a marked pref-
erence for octahedral coordination, 0.9 Co 2+ ions may occupy only the spin-down
octahedral sublattice 6Cw, while the residual 1.1 Co 2+ ions probably distribute
themselves among the 3avl, 18hvb 3by1 sublattices.
In Ba2Zn2-2xCu2xF12022,(Zn2-2xCU2x)Y, the substitution of Zn by Cu resulted in
a linear decrease of the saturation magnetization, o-~, as well as a flattening of the
o-~ vs. T curve (Albanese et al. 1978). This suggests that for all the compositions
nearly 28% of Cu ions enter the spin-down sublattices and the residual 72%
occupy the spin-up sites. Among the spin-up sublattices, the 3bvi sites at the
centre of the T-block play an important role for the equilibrium of the superex-
change interactions in this compound, as already described. In addition, other
experiments on Zn2Y by Savage and Tauber (1964), Mn2Y and (MnZn)Y by both
Tauber et al. (1964) and Dixon et al. (1965) have been performed.
In fig. 11 the saturation magnetization is plotted as a function of temperature
for polycrystalline specimens of Z-type compounds. Zn2Z shows the highest
saturation magnetization. The distribution of Co 2+ ions in Co2Z was deduced by
Albanese et al. (1976a) such that 1.08 Co 2+ ions per unit formula occupy the
spin-up sublattice, while the residual 0.92 enter the 4fvi and 4ev~ sublattices which
are the only spin-down octahedral lattice sites. The Curie temperatures as seen in
fig. 11 are in agreement with their values obtained from M6ssbauer measure-
ments. Petrova (1967) measured the saturation magnetization and Curie tem-
perature of (Co2-xZnx)Z.
408 M. SUGIMOTO

BO

( Boo.05 Sr o.g5 Zn)Y


40 (a)
H.Lc o - - - ' _ ~
~ c - T~290 ~K
0
80 I I i

F:

40 (b)
%

-4" 0

BO

40 t ~ (c)
E
0~ - Hllc T=I2OO°K

80 i -r r

0
0 5 10 15 20
H ( k Oe ) [ x 106/(z./T) A / rn ]
Fig. 10. Magnetization curves of single crystal of (Ba0.0sSr0.95Zn)Y(Enz 1961).

80
\
E Me2Z
i
C o " ~ Me=Zn
60

x /.0

~ 2o

0
-273 -200 -100 0 100 200 300 /+00 500
T(*C)
Fig. 11. Saturation magnetization as a function of temperature for compounds with Z-structure,
measured on polycrystalline specimens at a field of 11000 Oe for Co2Z and Zn2Z and 18000 Oe for
Cu2Z (Smit and Wijn 1959).
PROPERTIES OF FERROXPLANA-TYPE HEXAGONAL FERRITES 409

Figure 12 shows the temperature dependence of the saturation magnetization


for Ba2Zn2Fe~O46(Zn2X), Ba2Co2Fe2sO46(Co2X) and BaeZn2Fe36060(Zn2U) single
crystals (Tauber et al. 1970 and Kerecman et al. 1968). If we assume that the Zn 2+
ions in Zn2X are equally distributed over the sublattices (spin-up and spin-down),
the magnetic moment can be calculated as n u = 50.0/XB ( H = % T = 0 ) by
reference to the collinear Gorter-type spin model of fig. 4. This calculated value is
in excellent agreement with the experimental value 50.4/xB ( H = w, T = 0).
However, crystal chemistry would require most of the Zn 2+ ions to be on
tetrahedral site, leading to:

20 x 5 - 8 x 5 = 60/x B .

Tauber et al. (1970) explained that this roughly 18% difference between experi-
mental and calculated result in ZnzX may arise because of the spin system is not
colinear or because of non-stoichiometry of the crystals. In the case of Co2X, the
calculated value 47#B ( H = ~, T = 0) is in good agreement with the experimental
value 46/XB. As pointed out by Tauber et al. (1970), the agreement may be
fortuitously given by the stoichiometry. The Curie temperatures of 705-+ 3 K for
Zn2X and 740_+ 4 K for CozX are the highest values among the ferroxplana-type
compounds containing Zn 2+ ions.
The magnetic moment of Zn2U was calculated as 60.5/xB from O-s ( H = o0,
T = 0 K). The value of the magnetic moment obtained from the sum of those for
M- and Y-blocks, 58.4 (Gorter 1957) and 59.2 (Vinnik 1966), is in good agreement
with the experimental value. However, a simple Gorter-type model (where all

100 _• I l I ' r ' 500

4OO
13r}
.x 80

.a
300 ~:
~ 6O
(:3
C)
×
x

200 -.t
= 40 d~ (Co~X) x

c9
v
E
2O 100 :~

0
0 200 /-.00 600
T (K)
Fig. 12. Saturation magnetization as a function of temperature for Zn2X, Co2X and Zn2U (Tauber et
al. 1970, K e r e c m a n et al. 1968).
410 M. S U G I M O T O

Z n 2+ ions are tetrahedral in the layers of Y and M but not all in spinel blocks)
gives 60tXa, seemingly in better agreement. The magnetic m o m e n t s in spinels
containing large amounts of Zn are lowered by the formation of angles between
the m o m e n t s of octahedral ions. T a u b e r et al. described that since this effect is
more pronounced in Zn2W than in Zn2Y, the experimental m o m e n t for Zn2U is
still less than the value predicted from theory. The Curie t e m p e r a t u r e of Zn2U
was given as 673 + 2 K.
Figure 13 shows the saturation magnetization of BasZn2Ti3Fe12031(Zn2-18H)
and BasMg2Ti3FeI2031(Mg2-18H) as a function of temperature. Zn2-18H could be
saturated in 7 k O e fields below 15 K in the easy plane and followed a o's = o-0 + x H
law. A magnetic m o m e n t nB ( H = ~, T = 0) = 14.1 ¥ 0.6/XB and Curie t e m p e r a t u r e
Tc = 310~-5 K were extracted from the magnetization data. Mg2-18H crystals
could not be saturated in 15.5 k O e field applied parallel to either the (0001) plane
or [0001] axis below 120 K. A b o v e 120 K the magnetization followed a o%=
cro+ x H law. At 300 K, Tc was measured as 391-7-3 K and nB ( H = ~, T = 0 ) =
7 . 8 * 5tXa was obtained by extrapolation from 120 K. Tauber et al. (1971) dis-
cussed the magnetic m o m e n t of these compounds as follows: A ferrimagnetic
resultant according to the following site arrangement per formula unit,
9°~--~--4tet~--4 °~t was predicted. When this alignment was used to compute the

40

~30
E
B
%
×

~ 20
X

I O0 200 300 ~00


T(K)
Fig. 13. Saturation magnetization as a function of temperature for Znz-18H and Mg2-18H (Tauber et
al. 1971).
P R O P E R T I E S OF F E R R O X P L A N A - T Y P E H E X A G O N A L F E R R I T E S 411

moment for Zn2-18H assuming the following cation distribution,

( 7 . 4 1 F e 3+, l T i 4+, 0 . 5 9 Z n 2+) ~ <-- ( 1 . 4 1 Z n 2+, 2 . 5 9 F e 3+) ~-- ( 2 F e 3+, 2.0Ti4+),

the resultant magnetic moment came to:

(37.05/-tB)--~ - ~-- (12.95/xB) - ~-- (10/xB) = (14.1/.t~).

Zn 2+ ions occupation in octahedral sites in Y-blocks was discussed by Townes a n d


Fang (1970). For Mg2-18H, one of many possible distributions giving the magnetic
moment is:

(6.78Fe 3+, 1.22Ti 4+, 2Mg)---~ ~--(4Fe 3+) ~--(1.22Fe 3+, 1.78Ti 4+, 1Mg)
× (33.9/xB)2-~- ~ ( 2 0 / x B ) - *--(6.1/.tB) = (7.8/.tB).

In table 6 the saturation magnetizations and Curie temperatures of the fer-


roxplana-type compounds are shown.

TABLE 6
The saturation magnetization of ferroxplana-type compounds at absolute zero and at 20°C, and Curie
point.

o-0 0-20
Compounds (gauss) (gauss) 47rMs T¢
symbol (cm3/g) (cm3/g) (gauss) (°C) Ref.

MnzW 97 59 3900 415


Fe2W 98 78 5220 455 Smit and Wijin (1959)
NiFeW 79 52 3450 520
ZnFeW 108 73 4800 430

Mn~Y 42 31 2100 290


CoaY 39 34 2300 340
Ni2Y 25 24 1600 390 Smit and Wijin (1959)
Zn2Y 72 42 2850 130
Mg2Y 29 23 1500 280

Co2Z 69 50 3350 410


CuzZ 60 46 3100 440 Smit and Wijin (1959)
Zn2Z 55 58 3900 360

Zn2X 107 - 4700 432 Tauber et al. (1970)


Cu2X 95 - 3900 467 Silber and Tsantes (1969)

Zn2U - - 3700 400 Kerecman et al. (1968)

Znz-18H 37 - - 37 Savage et al. (1974)


CuNi-18H 20 _+2 - - 291
412 M. SUGIMOTO

4. Magnetocrystalline anisotropy and related phenomena

One of the reasons for the great scientific and technical interest in the fer-
roxplana-type compounds is their large crystalline anisotropy. In addition, this
anisotropy can be modified over a large range of values by the substitution of the
transition metal ions for other divalent ions. A simplified description of the
crystalline anisotropy energy density, applicable to most ferroxplana-type com-
pounds, is given in ch. 5.
The anisotropy constant, K1 = HAMs~2, and the anisotropy field, HA, as a
function of temperature for BaZn2Fe16027(Zn2W) are shown fig. 14 (Albanese et
al. 1976b). The value of the anisotropy field was deduced from the saturation
magnetization. Although Ks of Zn2W is lower than that of BaFea20~9, the
saturation magnetization of Zn2W at room temperature is substantially higher
than that of any other hexagonal compounds. For this reason, Zn2W has been of
interest for special applications as permanent magnet materials (Lotgering et al.
1980). According to Albanese et al. (1977), /{-1 of Zn2W decreases linearly with
increasing substitution of AI for Fe in its structure. The experimental values of/{1,

~o.3
X

%U 2 (a)

% i
X

0 100 200 300 /*00 500 600


T(K)

_<.~15 I HA I

x (b)

3= 0 ~.
~00 200 300 zoo soo 600
T(K)
Fig. 14. Temperature dependence of the first anisotropy constant KI and of magnetic anisotropy field
HA for Zn2W (Albanese et al. 1976b).
PROPERTIES OF FERROXPLANA-TYPE H E X A G O N A L FERRITES 413

/£2 and K3 are shown in fig. 15 as a function of temperature for a


BaZnl.38Co0.62Fea6027, (ZnCo)W, single crystal. In addition, there were studies on
the crystalline anisotropy of (CoxFe2-x)W by Bickford (1960b), by Lotgering et al.
(1961) and Perekalina (1964); on the crystalline anisotropy field of Zn2W, Ni2W,
NiZnW, (ZnxNi2-x)W, (Nil.sCo0.4)W and (Nil.26Co0.74)W by Hodges and Harrison
(1964), (Ni2_xCo~)W by Rodrique et al. (1962).
The dependence of the anisotropy on the Cu content and on temperature for
Ba2Zn2-2xCu2,Fe12022, (Zn2-2xCu2x)Y, polycrystalline samples was evaluated by
Albanese et al. (1978) assuming the Stoner and Wohlfarth model (1948) using a
p a r a m e t e r H A defined as the field at which the magnetization reaches 95.4% of its
saturation value in the planar phase or, 91.3% for the axial configuration. The
temperature dependence of HA for (Zn2-2xCu2~)Y is shown in fig. 16. The
introduction of the Cu ions strongly influenced the magnetocrystalline anisotropy.
Smirnovskaya et al. (1978) measured the magnetic anisotropy of
Ba2COl.ysZn0.2sFe12022, (Col.75Zn0.25)Y, single crystals in the temperature range
78-290 K. The measurements carried out in various magnetic fields up to 20 kOe for
three crystallographic planes; the basal plane, the plane (0110) and (1210). Figure 17
shows the torque curves of (Col.75Zn0.25)Yin the (0110) plane and the basal plane. The
torque curve measured in the basal plane at 78 K in a field of 20 kOe is shown in fig.
17(a). It follows from this curve that the easy magnetization directions in the basal
plane meet at angles of 60° and that they lie along the crystallographic direction
[1010], [1100] and [0110]. Figure 17(b) gives the torque curves obtained at 78, 140 and
175 K in magnetic fields of 5, 10 and 20 kOe in the (0110) plane. Furthermore, there
are many experiments on the torque anisotropy measurements for Co2Y and Co2Z by

K;...___
f
E
g
o
0

g- I
E -1
o

to
o -2

.,e"
-3 /
200 40O 6O0
T (K)
Fig. 15. Temperature dependence of anisotropy constants K~, K 2 a n d K3for (ZnCo)W(Asti et al.
1978).
414 M. SUGIMOTO

25
(Z 2-2X C u 2 x ) Y

, X=l

20
\

15

x
X=7

"10

×=0
1-

0
0 ioo 20o 3o0
T(K)
Fig. 16. Temperature dependence of anisotropy field HA for various x values (Albanese et al. 1978).

Bickford (1960a) as well as the crystalline anisotropy field for Zn2Y, C02Y, Mg2Y and
Ni2Y by Smit and Wijn (1959).
The temperature dependence of the first-order anisotropy constant/£1 and the
second-order constant K: for Ba3Co2Fe24041, CozZ compound is shown in fig. 18.
According to Rinaldi and Asti (1976), a positive cubic first-order anisotropy due
to Co 2+ ions in octahedral sites probably gives a negative contribution to K1 and a
positive contribution to K2. It seems reasonable to assume that K2 is essentially
due to cobalt ions, because K2 is nearly zero for all the hexagonal ferrites not
containing cobalt. The transition from planar to axial configurations can be
interpreted by taking into account the various sublattices whose magnetizations
have different temperature dependences for K~, At low temperatures the prevail-
ing contribution to K1 is probably from Co 2+ ions occupying the 2dv sublattice.
Figure 19 shows the complete magnetization loop and a section of it with some
inner curves of polycrystalline Ba3Co~.75Zn0.2sFe24041, (CoxZn2_x)Z measured at a
temperature of 4.2 K. The measurements were carried out in several cycles with
increasing field amplitude in the positive field range. The recurrent curves show
that irreversible processes occur when the sample is magnetized up to above 13'.
Moreover, fig. 20 gives magnetization curves of a grain-oriented sample of the
same composition. This constricted Perminvar-type hysteresis loop may be inter-
preted by a reversible rectifier characteristic induced in single domain particles
with conar anisotropy.
13a2Zn2Fe28046(Zn2X) exhibits a uniaxial anisotropy at all temperature and
P R O P E R T I E S OF F E R R O X P L A N A - T Y P E H E X A G O N A L F E R R I T E S 415

[10~01 111201 [01101 [1210l 111001


I I ] I L ] I I I ] I ] I
(a)
0 10 20 30 /.0 50 60 70 80 90100110120
~P ( d e g r e e )

0 ~- -

~ -¢.0

o
-ao

120
80 , , ~
D E \.F (b)

J /.,0
0
-40
-80
-120 t I L I I I I

8090100110 80 90100110 8090 100110


-0" (degree)
Fig. 17. Torque curves in the (0iI0) plane for Col.75Y. (A) 20 kOe, (B) 10 kOe, (C) 5 kOe, all at
T = 78 K, (D) 140 K, (E) 160 K, (F) 175 K, all in H = 20 kOe (Smirnovskaya et al. 1978).

3
%

-1
200 z,O0 600
T (K)
Fig. 18. Temperature dependence of crystalline anisotropy constants K1 and K2 for Co2Z (Albanese et al.
1976a).
416 M. S U G I M O T O

revemib

,4-" ~ /" / / / PB'/'~ T=4.2 K /

" 250 - ~ . . . .
/ < ~eversible
o %3 S
"; / ~2-
1-
200

150
0 1 20
H (kOe [,,t061(z, TT) Alrn]
Fig. 19. Magnetization curve of isotropic polycrystalline sample (Col.75Zn0.25)Y, T = 4.2 K (Oerling
1970).

Ea=K~sin20. Figure 21 shows the temperature dependence of the mag-


netocrystalline anisotropy constant K~ for Zn2X or C02X which was obtained by
evaluating HAMs~2 (Tauber et al. 1970). At room temperature the c-axis of
Ba2Co2Fe28046(Co2X) is the hard direction and (0001) the easy. The anisotropy
energy is given as follows:

EA = K~ sin 2 0 + K2 s i n 4 0 + K~ s i n 3 0 c o s 0 c o s 3q~ 2 . (2)

At about 416 K, however, E A = K 1 sin20 and the c-axis becomes the easy direc-
tion and [0001] the hard.
Silber and Tsantes (1969) made measurement of the magnetocrystalline aniso-
tropy of single crystals of BazZn2Fe2,O46(ZnzX) and BazCo2FezsO46(CozX) by the
ferromagnetic resonance method at room temperature. At room temperature, the
one-ion Slonczewski theory (1958) accounted rather well for the change in
anisotropy on substitution of cobalt for zinc in X-type compounds. The higher
order anisotropy constants observed in the cobalt-bearing samples can be unam-
biguously attributed to the cobalt ions. In fig. 22 the dependence of the anisotropy
constants on temperature for CoZnX is shown.
The temperature dependence of the crystalline anisotropy constant/£1 and the
anisotropy field HA for Ba4Zn2Fe36Or0(Zn2U) crystal is shown in fig. 23. KI was
calculated from H A and Ms at different temperature. A rapid variation of/£1 with
increasing temperature is seen. The constancy of H A for a wide range of
PROPERTIES OF FERROXPLANA-TYPE H E X A G O N A L FERRITES 417

o 2
X

(a)
L t [ i I
2'0
~ -1 H (kOe)

H//c

~ 3
%
~ 2

IE 1 (b)
I I
-26 -,'o 10 15 20
-1 H (kOe)

-2
HLc

'~ 3
%
2
S
:E I (c)
I I
-20 -1~5 - 1'0 - 5 1(3 15 20
-I H (kOe)

-2
HAc
J -3

Fig. 20. Magnetization curves of grain oriented (Co1.75Zn0.25)Y, T = 4.2 K. (a) a = 0°, (b) a = 45 °, (c)
a = 90°. a is the angle between the c-axis and the direction of applied field (Gerling 1970).
418 M. SUGIMOTO

100 I ' I ' r r 25

~ 8O 20

o
1 (C02X) <

~ 6o
i.I

2 ~0
o

~ 20 5

o 0
0 200 400 600
T (K)
Fig. 21. Temperature dependence of crystalline anisotropy constant K1 and anisotropy field HA of
Zn2X and Coax (Tauber et al. 1970).

1.2

~E 1.0
~n~ 0.2

0.6 K1

:~" "~ 0.1 2 -0.4

0.0 I
o . o = ~ ~ ~ -o.6F
100 "200 300 L00 100 200 S00 L00 100 200 300 ,'-00
T(K) T (K) T (K)
Fig. 22. Temperature dependence of crystalline anisotropy constants for CoZnX (Silber and Tsantes
1970).
P R O P E R T I E S OF F E R R O X P L A N A - T Y P E H E X A G O N A L FERRITES 419

2 2
E -g
u~
O
× -.i"

E %
oi x

%
O
x HA
v

200 400 600

T (K)
Fig. 23. Temperature dependence of crystalline anisotropy constant K1 and anisotropy field HA for
Zn2U (Kerecman et al. 1968).

i I i I I
4 12
K2

E 0 10
u E

~o -0.5 z~ ~ 8

-~.0 6 "~

\ - 1.5 z, o
HA

-~ -zo 2
:Z"

i I l L
0 100 200 300
T K)
Fig. 24. Temperature dependence of crystalline amsotropy constants K1 and/<2, and anisotropy field
HA for Zn2-18H (Tauber et al. 1971).
420 M. SUGIMOTO

temperatures may be due to the proportional change in K1 and Ms.


BasZn2Ti3FelzO31(Znz-18H), B asMg2Ti3Fe12031(Mg2-18H), B aSll(Nil.lCUo.4)Ti2.7-
Fe]z3Mn0.4031(NiCu-18H) and Bas.a(Mgl.3Zn0.7)Ti2.9Fen.7031(MgZn-18H) exhibit
an easy plane of magnetization at all temperatures except for N i C u - 1 8 H and
their anisotropy energy is EA = K1 sin 2 0 + K2 sin 4 0. T h e anisotropy constants K1
i i I r i 4

2 ,10

K~ 7
"-. f \
<
go_ 8~
x

~-2 6
o x

-4 /-. o

-'r

,x-," - 6 2

I r I i I
100 200 300 400
T(K)
Fig. 25. Temperature dependence of crystalline anisotropy constants K1 and/(2, and anisotropy field
HA for Mg2-18H Tauber et al. 1971).

r I I i I

1 - [] K2 6

-
g
o s
g<
y...
-1 ~ ~o~

\ -2 3
o

"~ -/, 1

I 1 I I I I 0
0 100 200 300
T (K)
Fig. 26. Temperature dependence of crystalline anisotropy constants K1 and K2, and anisotropy field
HA for MgZn-18H (Savage et al. 1974).
PROPERTIES OF FERROXPLANA-TYPE HEXAGONAL FERRITES 421

and K2 for Zn2-18H, Mg2-18H, M g Z n - 1 8 H and CuNi-18H, respectively, are


shown in figs. 24-27. The crystalline anisotropy constants KI and K2 were extracted
from the following equation:

I--I~/I = (-2K1 + 4K2)/Ies + (4K212/I4), (3)

where Hi is the internal field, Is is the spontaneous magnetization. The anisotropy


field HA was computed from

HA = -- 2(K1 + 2K2/Is). (4)

The magnetization was invariant to rotation about [0001] for all temperatures at
which the basal plane is the easy plane of magnetization.

3 i I J I i I ~ r 8

1 6 <
E
u
"" 0 5--

¢,1 .ac

.e
- -3 2 ~

-4 1

-5 ~ 1 i I i I , I 0
100 200 300 4,00 500

T(K)

Fig. 27. Temperature dependence of crystalline anisotropy constants K 1 and K2, and anisotropy field
H A for CuNi-18H (Savage et al. 1974).

5. Miissbauer effect

Many studies on the M6ssbauer measurements for both single crystals and
polycrystalline samples have been carried out to determine the magnetic
behaviour of the cations in various lattice sites. Figures 28 (a) and (b) show the
spectra of the polycrystalline Mg2W measured at 85 and 300 K, respectively. The
spectra have been resolved into different Z e e m a n sextets due to the Fe 57 nuclei in
the various sublattices. In discussing the connection between the sextets and the
sublattices, it is possible to use the analogy with M-type compounds, because the
422 M. S U G 1 M O T O

i l I l I r [ l 1 [ 1

1.00

O.98

0.96

•, 0.9 L
0
~- 0.92
I I I I II I
II , , , , ,
c- III L.i . . . . . . L. . . . . L . . . . . l . . . . . . . . . . . J

t-
tOO
o
u 0.99

~, 0.98
>
.- 0.97
0.96
O
n.. 0.95

I L E e I J
II , L , ,
III j , , r ,
IV , I , l
V L-L . . . . . t . - _t . . . . l ........ 1

J I I I r I I I I
-I0 -8 -6 _L. -2 0 2 ~. 6 10

Velocity (mmlsec)

Fig. 28. MOssbauer s p e c t r u m for polycrystalline MgzW c o m p o u n d : (a) T = 85 K. I - sublattices a, fw,


fvl, I I - sublattice K, I I I - s u b l a n i c e b; (b) T = 300 K. I - s u b l a t t i c e K, I I - sublattice ftv, I I I - s u b l a t t i c e
a, I V - sublattice fvi, V - sublattice b ( A l b a n e s e and Asti 1970).

W-type structure is closely related to that of the M-type compound. For example,
the sextets attributed to Fe in the trigonal position b (fivefold coordination) are
denoted in fig. 28 by a dotted line for 1 site out of 18 in the unitary cell.
In fig. 29, the temperature dependence of the hyperfine magnetic fields for the
observed sextets is shown. Albanese et al. (1976a) also measured the M6ssbauer
absorption in Zn2W to determine its spin orientation. Kimich et al. (1970) studied
the Fe2W compound. The preferred sites of Fe 2+ in the lattice of Fe2W were
determined by Fayek et al. (1980) from a M6ssbauer study.
The M6ssbauer absorption spectra of Fe 57 14.4keV gamma rays for poly-
crystalline samples of Co2Y at different temperatures a r e shown in fig. 30. The
spectrum measured at 78 K shows only one Zeeman sextet with a mean hyperfine
magnetic field of 508kOe. With increasing temperature the spectra can be
interpreted as the superposition of three different sextets. Albanese et al. (1975b)
established the correspondence between the three observed sextets and the iron
sublattices of CozY structure as follows: Sextet II, being the only one which is
contracted in the presence of the external field, is due to all the Fe 3+ ions with
spin up, i.e., the sublattices 18Hvi, 3avl, 3bvi. Sextet III, the one having the highest
hyperfine field, can be assigned to the sublattice 6C~v on the grounds of the
PROPERTIES OF FERROXPLANA-TYPE HEXAGONAL FERRITES 423

I I ] I I I I
600

K
500 -.~--&-.-&.-~.::... b
--
__
- - - -

L,oo
\L.IX
x 300

200

-r I00

I I I I 1 I I I
100 200 300 t-O0 500 600 700 800
T(~)
Fig. 29. Hyperfine magnetic fields at iron nuclei of various sublattices ol Mg2W c o m p o u n d as a
function of temperature (Albanese and Asti 1970).

0,l 909B
~ 0,:4' 9, ~:""0.. 9i .....
6 1 ;. 0.;: 0 . 2,.. /. .. .A~

0.90
100 ':~,,: .~. ~ ~ ~",,. = .;s'~c

0.98
0
036

.E 0.94 b)
E I L I I I I
21 II k , , i __,
O
U
~. 1.00
0.98
-~ 0.96
er 0.9z, c)

Ill ~ , , _ i ,

1.00
0.9B
0.96
0.94
092
090 (d)
i I i I I A_ I I I i i

-10-8-6-4-2 0 2 4 6 8 10
Velocity ( m m / s e c )
Fig. 30. Mbssbauer spectra for CozY compound: (a) T = 78 K, (b) T = 296K, (c) T = 403 K, (d)
T = 673 K (Albanese and Asti 1970).
424 M. S U G I M O T O

I I I ¢ I I i 1 I I I I I I h I I I P i

1.00

¢-
0.98
¢-

~ 0.96

] i J I -- i L J
[[ I i i i l J

III u_ , ~ i ~ ,

I I [ I I J I I I I I I I I ! i r i )
-10 -8 -6 -4 -2 2 L 6 8 10

Velocity (mm/sec)

Fig. 31. M6ssbauer spectrum for CozZ polycrystalline compound at room temperature (Albanese and
Asti 1970).

[ ~ I ~ I I I ' t r I f I i I r I i I i [

0.980.961'00~ ~ ~ . . , ., . . . . .:~:~-•~F

O~ I I I
it I I I
.~- i ~ l
'~ ] , J I [ I I , P E i ~ I , ]
ou -10 -B -6 -/, -2 0 2 z, 6 8 10
> -- I I ] i I i i I ~ { ~ P ' i ~ i i I

1.00 4:.......... i"~ . ,.""'" .~ ".

0.98

0.96

0.9/.,
l I I i I
[ I I l I
I I I I I I

-10 -8 -6 -4 -2 0 2 4 6 B
Velocity (mm/sec)
Fig. 32. M6ssbauer spectra for Co2Z single-crystal compound at room temperature: (a) Hext = 0, (b)
/-/~×t = 25 kOe (Albanese and Asti 1970).
PROPERTIES OF FERROXPLANA-TYPE HEXAGONAL FERRITES 425

spectrum taken in an external field. Finally, sextet I must be assigned to the


remaining spin-down sublattices, i.e., to 6Cry and to the fraction of 6CvI sublattice
occupied by iron ions. M6ssbauer measurements on the Mg2Y compound (A1-
banese et al. 1975b), on (Znl-xCux)2Y (Albanese et al. 1978), on Zn2Y and
(Mn0.625Zn)2Y (Albanese et al. 1968) have been reported.
Figure 31 shows the M6ssbauer spectra for a polycrystalline sample of CozZ at
room temperature. T h e spectra can be interpreted as the superposition of at least
four sextets. In order to determine the spin orientation of the various iron
sublattices, one often uses the M6ssbauer spectrum for a single crystal measured
in an external magnetic field perpendicular to the c-axis and to the gamma ray
direction. The spectra for a CozZ single crystal in the absence and in the presence
of the external magnetic field of 25 kOe, respectively, are shown in fig. 32. In the
presence of the external field sextets II and III collapse into a single sextet with
broad lines proving that sextet III is due to iron sublattices with spin-up and sextet
II to iron sublattices with spin-down. In a similar way, by comparing the Hhf
values both in the presence and in the absence of external field, sextet IV is
assigned to be the spin-down sublattice. As regards sextet I, from the spectrum in
an external field, it may b e deduced that part of the contributing sublattices has
spin-up, and part has spin-down.

6. Magnetostriction and NMR

6.1. Magnetostriction

The anisotropic part of the magnetostriction of the ferroxplana-type compounds is


expressed by a phenomenological formula with four constants, which is the same
formula as that of W-type compounds described in ch. 5. The constants AA, AB, Ac
and AD are also determined experimentally in the same way. The values of these
constants for FeeW single crystals were determined as AA = 13× 10-6, As =
3 × 10-6, Ac = - 2 3 × 10-6 and AD = 3 × 10-6 using wire strain gauges (Fonton and
Zalesskii 1965). In these measurements, two samples in the form of disks having
different cuts were used: parallel to the basal plane and parallel to the plane with
the c-axis.
Figure 33 shows the magnetostriction of a single crystal of Fe2W cut parallel to the
basal plane. In fig. 33(a), c u r v e s /~A and AB illustrate, respectively, the dependence
of the longitudinal and transverse magnetostriction in the basal plane on the
external field. The magnetostriction A, reaches saturation in fields of about
18000 Oe, corresponding to the saturation magnetization of the sample at fight
angles to the c-axis. Curve C represents the change in the magnetostriction in the
basal plane when the sample is magnetized along the c-axis (at right angles to the
plane of the disk).
Figure 33(b) shows the dependence of the magnetostriction in t h e basal plane
on the angle q~, obtained by rotating the crystal in a magnetic field of 21000 Oe.
Here ~ is the angle between the direction of an external magnetic field and the
426 M, S U G I M O T O

18 18
16 (a) 16
1/, XA I~
12 12
10 10
8 8
6 6
4 /.
"~ 2 2
% r q r I r f
0
18 21 2427 30 60 90 120150180
-2 ~ 2

-4 e) -4 (degree)
-6 -6
-8 o -8
-10 -10
Fig. 33. Magnetostriction of a disk, cut parallel to the basal plane of Fe=W single crystal: (a)
dependence of the magnetostriction in the basal plane on the external intensity, (b) dependence of the
magnetostriction along the direction x on the orientation of the magnetization vector in the basal
plane (Fonton and Zalesskii 1965).

c-axis. The value of the magnetostriction for q~ = 0 represents the transverse effect
in the basal plane and the longitudinal effect is given by ~0 = 90 °.
Figure 34 shows the dependence of the magnetostriction on the external field
intensity for a single crystal of Fe2W with disk shape cut parallel to the plane with

15

10

,-<
~o" -5
O

-10

-15

-20

Fig. 34. Dependence of the magnetostriction on the external field intensity for a disk, cut parallel to
the plane with the c-axis of Fe2W single crystal (Fonton and Zalesskii 1965).
PROPERTIES OF FERROXPLANA-TYPE HEXAGONAL FERRITES 427

the c-axis. T h e ,hA curve in fig. 34 corresponds to the ,hA curve in fig. 33. The
difference in the saturation magnetostriction in these two cases seems to be due to
the different initial states of the samples. Since, as seen in the figure, the
magnetostriction constants are opposite in sign and analogous to that o f a cobalt
crystal, Fe2W exhibits first an elongation and then contraction of the linear
dimensions. Kuntsevich and Palekhin (1978) measured the t e m p e r a t u r e depen-
dences of the longitudinal (All) and transverse (hl) saturation magnetostrictions in
the basal plane for BaCo1.sFe0.2Fe16027(Col.8Fe0.E)W,
2+ 3+ and observed their abrupt
changes in the t e m p e r a t u r e range 430-460K. In addition, the t e m p e r a t u r e
dependence of the longitudinal magnetostriction (`hn) for Cu2Y in the t e m p e r a t u r e
range 200-350 K was measured by Belov et al. (1980).

6.2. N M R

T h e nuclear magnetic resonance (NMR) of 55Mn has been studied in single


crystals of Ba2Zn2_xFe12_yMnx+yO22(Zn2Y type) for values of (x + y) of about 0.5
by Streever et al. (1971). In fig. 35 (a) and (b), the zero field N M R spectra of
Mn0.sZnl.sY single crystal measured at 4.2 K and 77 K are shown. The spectrum at
4.2 K was interpreted on the basis of the MnFeaO4 results that the broad line

1.0 i i i I I

0.8
4.2°K

0.6
(a)
0.4

~. 0.2
E

>

"~ 1.0 ~ i I i I
ct"

0.8
7 7*K
0.6
Cb)
0.t.

0.2

'1
300 340 380 420 460 500 5/.0 580 620
Frequency (MHz}
Fig. 35. The zero field NMR spectrum of Mn0.sZnY measured at (a) 4.2 K, (b) 77 K (Streever et al.
1971).
428 M. SUGIMOTO

extending from about 300 to about 470 MHz is due to the Mn 3÷ ions (replacing
iron) on the octahedral sites and two relatively narrow lines at 555 and 585 MHz
are due to the Mn 2+ ions (replacing zinc) on the two types of tetrahedral sites with
roughly the same degree of preference. Both Mn 2+ lines were observed to shift to
higher frequency with externally applied fields which is consistent with this
interpretation. The hyperfine fields and the temperature dependences of the
resonance frequencies for the Mn 2÷ ions on the two types of tetragonal sites can
be consistently explained in terms of the different electronic environments of the
two sites.

7. High frequency magnetic properties

Ferromagnetic resonance in the high frequency range is a well-known


phenomenon. It involves the excitation of magnetic dipoles in a material under
the driving force of a high frequency magnetic field. As a material for high
frequency or microwave devices, considerable interest has previously been shown
in the ferroxplana-type compounds having large magnetic anisotropy and lower
dispersion frequency.
In compounds with planar anisotropy, ferromagnetic resonance occurs at a
frequency which depends on both the applied magnetic field HAl and the
anisotropy field HA2 as illustrated in fig. 36. The resonance frequency is given by
the relation (Bady 1961):

(/"/7) 2 = H&(HA, + HA2),

where F is the resonance frequency (GHz) and y is the gyromagnetic ratio


(assumed to be 2.8). Since/-/12 usually can take very much higher values than HA1,
the resonance frequency reaches higher values. Moreover, the static initial per-
meability is given by (Smit and Wijn 1959):

#Zo- 1 = ~4~'o'JHA, ,

where/z0 is the static initial permeability and o-s is the saturation magnetization.
As seen in this equation, it is possible to obtain a compound with higher initial
permeability up to the high frequency range by controlling the value of HA1.
Figure 37 shows the frequency dependence of the initial permeability for Co2Z.
For comparison, this figure also gives the property of the spinel NiFe204. Although
the permeability at low frequency is approximately equal for both compounds, the
dispersion frequency of about 1000 MHz for Co2Z is substantially higher than that of
about 200 MHz for the spinel. Such higher-frequency dispersion characteristics are
found with many other planar compounds having large HA,. The dispersion may be
due to the ferromagnetic resonance.
The experiments on the ferromagnetic resonance linewidth can be roughly
divided into three classes according to the measuring method: The low power
PROPERTIES OF FERROXPLANA-TYPE H E X A G O N A L FERRITES 429

14
I ! I
12

10

2 --

] 4
0
10 2 5 100 2 5 1000 2 5
Frequency (MHz)

Fig. 36. Dependence of ferromagnetic resonance frequency of a spherical sample of a ferroxplana


compound on the applied magnetic field and the anisotropy field (Braden et al. 1966).

transverse pumping method, the higher power transverse pumping method and
the high power parallel-pumping method. The theoretical discussion of
Schl6mann et al. (1962, 1963) has led to the conclusion that a large planar
anisotropy favours the excitation of spin waves and reduces the threshold for the
onset of instability. Mita (1965) has analyzed the first and the second order
non-linear behaviours observed in Zn2Y on the basis of this theory, and he also
analyzed the linewidth induced by surface imperfections in Zn2Y on the basis of
the two-magnon mechanism (1968). The theory of parallel pumping phenomenon
has been worked out by Bady et al. (1962). Douthett et al. (1962) found that their
phenomenological theory closely predicts the response in Zn2Y single crystal
through frequency doubling experiments at low levels. Furthermore, Hwa and
Silver (1977) have experimentally compared the two theoretical models of Gure-
vich (1969) and Rado (1972) for ferromagnetic resonance, and found that both
models give good agreement with the experimental results of Mg2Y at 9.3 GHz.
The microwave linewidth AH of Y- and Z-type compounds respectively are
shown in figs. 38 (a) and (b) (Braden et al. 1966). In Y-type compounds, zXH
varied between 100 Oe in (Cu0.7Ni0.3Znl.0)Y and 1430 Oe in Co2Y. In general, the
Y-type compounds were found to have the most useful microwave properties
compared with W- and Z-type compounds. The temperature dependence of the
linewidth of Co2Y is shown in fig. 39. The linewidth increases linearly as the
temperature is decreased to 220 K, and below 220 K it remains fairly constant
down to 77 K. The linewidth of C o 2 Y behaves similarly to that of Zn2Y. The field
for resonance (H0) decreases to a sharp minimum at 220 K, and then increases
again as the temperature is lowered. Grant et al. (1974) observed the lowest
resonance linewidth of 38 Oe for a single-crystal sphere of Zn2Z at 35 GHz at
430 M. SUGIMOTO

40

35

30

"~ 25
1-"

c 20

IJ"
L.
m 15

10

0 2 ~ 6 8 10 12
HAl(kOe) ix106 /(47F)AIm]
Fig. 37. Magnetic spectrum of a polycrystalline specimen of C o 2 Z . For comparison, the spectrum of
NiFezOa with approximately the same low frequency permeability is given (Smit and Wijn 1959).

r o o m temperature, while Braden et al. (1960) obtained the linewidth of 1115 O e


for polycrystalline Zn2Z. K e r e c m a n et al. (1969) measured the linewidths of ZnzU
and Mn-substituted ZnzU and obtained the value of 120 and 1 8 . 5 0 e at 26.5 G H z
and at 300 K. Savage et al. (1965) obtained a minimum linewidth of 3 . 8 0 e at
9.0 G H z at r o o m temperature for Mn-substituted ZnzY single crystal. Weiner and

C02 Y 1260 655 Zn2 Y 655 1260 C02 Y


1430 • 29~0 " - "

N i2 Y 300 870 Cu Y 87o 30o Ni2 Y


Fig. 38(a). The microwave linewidtb of Y-type compounds (Braden et al. 1966).
PROPERTIES OF FERROXPLANA-TYPE HEXAGONAL FERRITES 431

Ni2 Z Ni2 Z

(b)

Zn 2 V Cu 2 7
Fig. 38(b). T h e microwave linewidth of Z-type compounds. T h e dotted line indicates the region of
zero anisotropy (Braden et al. 1966).

, , , i , , '- 4000
1800 f =23 k M c
o_

"-{ 1600
< - 3000 <
1~o0 ~-
,~ AH d

G 1200 ~,
2000
~. 1000 o

o o
800 "~

z~q 1000 :~
600

/.00 ~s
I I I I Y r I I I A 0
1oo 150 200 250 300 3so
T(K)
Fig. 39. Field f o r resonance, Hr, and l i n e w i d t h , / - / , of Co2Y as a function o f t e m p e r a t u r e ( B u f f l e [ ]962).

Dixon (1970) have theoretically and experimentally investigated subsidiary-


resonance phenomena in an Mn-substituted ZneY single crystal and observed a
linewidth of 5.0 Oe at 17.2 GHz. Investigations on the effect of surface finish on
the linewidth were carried out by Dixon (1963) for Zn2Y and by Kerecman et al.
(1969) for Mn substituted Zn2U.
The variation of the magnetic susceptibility with microwave magnetic field for
Zn2Y single crystal is shown in fig. 40. Measurements were made on an unpolished
sample with a diameter of 0.02 inches. It should be noted that the susceptibility,
the field required for resonance, and the shape of the resonance curve vary as a
function of RF magnetic field intensity (Dixon 1962). Green and Healy (1963)
measured the imaginary part of the parallel susceptibility for Zn2Y single crystal
with spherical and flake shapes and calculated the spin wave linewidth of 8 . 0 0 e
432 M. S U G I M O T O

~~
I J I J I I I
1.0 -
X
(0.140e)
X

__>, 0.8

Q.

8 ",/~~~(3650e)
0.6

.,$
i-
O'1
0.4
F

N 0.2
E
o
z
0 1 r I ? o i I
400 600 800 1000 1200 1/.,00 1600 1800
HA[0e ) [ x 103/(/.,/'/1A/rn ]
Fig. 40. Variation of the normalized susceptibility and the field required for resonance with the
incident rf power as a p a r a m e t e r for Zn2Y compound: frequency = 8957 Mlqz, A H = 85 O e (Dixon 1962),

for a sphere of Zn2Y and 9.24 Oe for flake Zn2Y, respectively, from the minimum
RF critical field. Helszajn et al. (1969) measured the decrease of the main-
resonance susceptibility at 17.03 GHz under second-order perpendicular pumping
for Mn-Zn2Y.
The anisotropy field for W-, Y- and Z-type compounds are shown in figs. 41
(a), (b) and (c). Table 7 shows the high frequency properties for some ferroxplana-
type compounds.

Ni2W Co2 W Ni2W

(a)

Zn2 W Cu2 W
Fig. 41(a). T h e approximate location of zero anisotropy for W-type compounds. T h e data show values
of linewidth and anisotropy field for s o m e c o m p o u n d (Braden et al. 1966),
PROPERTIES OF FERROXPLANA-TYPE H E X A G O N A L FERR1TES 433

Co2Y Zn2Y Co2Y


15 21

N; 2 ¥ Cu2¥ Ni2 ¥
Fig. 41(b). The anisotropy field (in kOe) of Y-type compounds which have plana anisotropy (Braden et
an. 1966).

Ni2 Z -t9 - 2.8 - 9.9 C02 Z_9.9 -2.8 -1.9 Ni2 Z

, \/-l~\/I-,OSjr-jj/.~",/,~__~

(c)

Zn2 Z Cu2 Z
Fig. 41(c). The anisotropy field (in kOe) of Z-type compounds. Planar anisotropy is indicated by a
negative sign, and uniaxial anisotropy is indicated by a positive sign (Braden et al. 1966).

TABLE 7
The microwave linewidth, the measuring frequency and the anisotropy field for fer-
roxplana-type compounds at room temperature.

Compounds AH f HA
symbol (Oe) (GHz) (kOe) Ref.

Zn2Y 13 9 9.5 Verweel (1967b)


(Zno.9sFeo2~os)2Y 14 9 12.0 Verweel (1967b)
Zn2Z 38 35 4.8 Smit and Wijn (1959)
Zn2X 45 54 13.7 Silber and Tsantes (1969)
ZnzU 155 26-40 10.0 Kerecman et al. (1969)
434 M. S U G I M O T O

8. Electric properties and other effects

8.1. Conductivity

In high frequency applications, the low electrical conductivity of materials is very


important. As was described in ch. 5 concerning W compounds, the ferroxplana-
type compounds with hexagonal structure show a similar anisotropy of the
electrical conductivity. The conductivity in the basal plane is higher than that
along the hexagonal axis. Figure 42 shows the temperature dependence of the
electrical conductivity, o-ii (parallel to the c-axis), and o-± (perpendicular to the
c-axis) for single crystals of Fe2W and Co2W (Simga et al. 1966). As seen in this
figure, the values of o71 are 2 to 3 orders higher than ~r±. The possible origin of the
anisotropy in the Conductivity may be the alternating layers in the direction of the
c-axis and thus the electrical resistance in this direction is greater than in the basal
plane. In table 8, values of the activation energies for Fe2W and Co2W are given.
Kasami et al. (1966) measured the temperature dependence of the electrical
resistivity for Zn2Y at 77-500 K, and observed a strong anisotropy in the electrical

10 I I i i i i

10- I

10-2

S 10-3
E
U
i
C~ 10-4

"t::,
10-5

lff 6

10-7

10 6
4 6 8 10 12
lIT (xlO-3K -1)
Fig. 42. Temperature dependence of the electrical conductivity, oql, (parallel to the c-axis) and o-±,
(perpendicular to the c-axis) for single crystals of Fe2W, Co2W and Zn2Y (Sim~a et al. 1966, Kasami et al.
1966).
PROPERTIES OF IZlERROXPLANA-TYPE HEXAGONAL FERRITES 435

TABLE 8
Activation energies, q, in (eV) for the conductivity of the samples Fe2W and
Co2W (Simga et al. 1966).

Direction Temperature range (K)


Crystal of current 7%100 100--170 1 7 0 - 3 0 0 300-670

Fe2W 11c-axis 0.025 0.029 0.041 -


L c-axis 0.029 0.033 0.039 -
Co~W IIc-axis 0.059 0.079 0.095 0.095
2 c-axis 0.064 0.075 0.083 0.083

resistivity as well as a l o w e r v a l u e of a.c. resistivity c o m p a r e d to that of t h e d.c.


resistivity. B u n g e t a n d R o s e n b e r g (1967) m e a s u r e d t h e e l e c t r i c a l resistivity of Co2Z
at 200-500 K. K a s a m i a n d K o i d e (1966) m e a s u r e d t h e electrical resistivity of Zn2U at
room temperature.

8.2. Dielectric constant

T h e d i e l e c t r i c c o n s t a n t s of p o l y c r y s t a l l i n e BaNi2AlxFe16_xO27, (NiA1) W , as a
f u n c t i o n of a l u m i n i u m c o n c e n t r a t i o n , x, has b e e n m e a s u r e d at 9.5 G H z a n d at
r o o m t e m p e r a t u r e (Taft 1964); See t a b l e 9.

TABLE 9
Dielectric constant, e of polycrystalline
isotropic compounds of BaNi2AlxFe16_xO27
(Taft 1964).

x 0.60 0,73 0.86 1.00

15.4 14.9 15.1 14.5

8.3. Jahn-Teller effect

C u F e 2 0 4 is a w e l l - k n o w n f e r r i t e which exhibits t h e J a h n - T e l l e r p h e n o m e n o n d u e
to C u 2+ ions l o c a t e d at o c t a h e d r a l sites. B e l o v et al. (1980) o b s e r v e d a s t e e p rise of
p e r m e a b i l i t y at 300 K for Cu2Y as i l l u s t r a t e d in fig. 43. Cu2Y, which has Tc--
650°C, e x h i b i t e d also a m a r k e d c h a n g e of t h e m a g n e t o s t r i c t i o n as well as of t h e
m a g n e t o r e s i s t a n c e at t h e s a m e t e m p e r a t u r e . T h e m a g n e t i c a n i s o t r o p y c o n s t a n t
was K1 < 0 b e l o w 300 K a n d K1 > 0 a b o v e this t e m p e r a t u r e . T h i s p h e n o m e n o n is
i n t e r p r e t e d as t h e J a h n - T e l l e r effect d u e to C u 2+ ions l o c a t e d at o c t a h e d r a l sites in
spinel blocks.
436 M. SUGIMOTO

20

15

~10

200 300 400


T(K)
Fig. 43. Temperature dependences of the permeability,/~, of Cu2Y (Belov 1980).

8.4. Magneto-optical effect

Figures 44 (b), (c) and (d) give the spectral dependence of the transverse Kerr
effect for Co2W, Co2Y and Co2Z single crystals, respectively. For comparison, the
spectrum of the W compound (a) is also presented. As seen in figs. 44 (b), (c) and
(d), all ferroxplana-type compounds exhibited a positive peak of rotation angle at
energies of 1.9-2.3 eV, respectively. Such a marked peak, however, could not be
seen in the W compound as shown in fig. 44(a). The peak at 2.5-5 eV was also
observed in the W compound.

8.5. Domain observation and chemical analysis

Verweel (1967a,b) reported that the Boch walls of the ferroxplana-type com-
pounds can be displayed by the well-known Bitter technique. Fano and Licci (1975)
reported an analytical method based on automatic potentiometric titration for the
analysis of Zn2Y.
PROPERTIES OF F E R R O X P L A N A - T Y P E H E X A G O N A L FERRITES 437

LIJ

x© -I
.= 0

= 0
o~
E~
(PoJ£OLX) Q'eI~UD IJO!~.D)~OJ J l e ~ l eSJ~)ASUDJL o~
0
0

hi

~o ~

• =

I J I I r

(PDJ£OLX) ~ el6Ub UO!;D~OJ JJa~l esJa^suoJ/


L~
438 M. S U G I M O T O

M -- 0

© 0

= E

i i [ I
i
E~
(FoJcO|x)~) ' el BuD uo!~D~oj JJa~l eSJeASUOJ1

> d"

I I I ) I _ /
~0 ",-4" ~ 0 ~ .~t
i • =
(pD.~Ol.X) 9 'eIBuD UO!ID],OJ JJ'~l a':JJeASUDJI
PROPERTIES OF FERROXPLANA-TYPE H E X A G O N A L FERRITES 439

References

Acquarone, M., 1979, J. Phys. C. Solid State Enz, U., 1961, J. Appl. Phys. 32, 22S.
Phys. 12, 1373. Fagg, L.W. and S.S. Hanna, 1959, Rev. Mod.
Albanese, G. and G. Asti, 1970, IEEE Trans. Phys. 31,711.
MAG-6, 158. Fano, V. and F. Licci, 1975, Analyst. 100, 507.
Albanese, G. and S. Rinaldi, 1974, J. Appl. Fayer, M.K. and A.A. Bahgat, 1980, Indian J.
Phys. 45, 3400. Pure Appl. Phys. 18, 945.
Albanese, G., G. Asti and C. Lamborizio, 1968, Fonton, S.S. and A.V. Zalesskii, 1965, Soviet
J. Appl. Phys. 39, 1198. Phys. JETP, 20, 1138.
Albanese, G., M. Carbucicchio and G. Asti, Gerling, W.H., 1970, IEEE Trans. MAG-6, 737.
1975a, Nuovo Cimento, 14, 207. Glass, H.L. and J.H.W. Liaw, 1978, J. Appl.
Albanese, G., M. Carbucicchio and A. Deriu, Phys. 49, 1578.
1975b, Appl. Phys. 7, 227. Gordon, J., R.L. Harvey and R.A. Braden,
Albanese, G., A. Deriu and S. Rinaldi, 1976a, 1962, J. Am. Ceram. Soc. 45, 297.
J. Phys. C. Solid State Phys. 9, 1313. Grant, R.W., M.D. Lind, G.P. Espinosa and
Albanese, G., M. Carbucicchio and G. Asti, I.B. Goldberg, 1974, AIP Conf. Proc. 24,
1976b, Appl. Phys. 11, 81. 493.
Albanese, G., M. Carbucicchio, F. Bolzoni, S. Green, J.J. and B.J. Healy, 1963, J. Appl. Phys.
Rinaldi, G. Sloccari and E. Lucchini, 1977, 34, 1285.
Physica, 86--88B, 941. Gundlach, R., 1968, Electronics, 41, 104.
Albanese, G., A. Deriu, F. Licci and S. Rinaldi, Harvey, R.L., I. Gordon and R.A. Braden,
1978, IEEE Trans. MAG-14, 710. 1961, RCA Review, 648.
Asti, G., F. Bolzoni, F. Licci and M. Canali, Helszojn, J. and J. Mestay, 1969, Electronics, 5,
1978, IEEE Trans. MAG-14, 883. 525.
Auld, B.A., R.E. Tokhein and D.K. Winslow, Hodges, L.R. and G.R. Harrison, 1964, J. Am.
1963, J. Appl. Phys. 34, 2281. Ceram. Soc. 47, 601.
Bady, I., 1961, IRE Trans. MTT-9, 60. Huijser-Gerits, E.M.C. and G.D. Rieck, 1974,
Banks, E., M. Robbins and A. Tauber, 1962, J. J. Appl. Cryst. 7, 474.
Phys. Soc. Japan, 17, 196. Huijser-Gerits, E.M.C. and G.D. Rieck, 1976,
Belov, K.P., A.N. Goryaga, L.G. Antoshina J. Appl. Cryst. 9, 18.
and M.M. Lukina, 1980, Sov. Phys. Solid Huijser-Gerits, E.M.C., G.D. Rieck and D.L.
State, 22, 2013. Vogel, 1970, J. Appl. Cryst. 3, B243.
Bickford, L.R. Jr., 1960a, Phys. Rev. 119, 1000. Hwa, C. and L.M. Silber, 1977, Physica, 86-
Bickford, L.R. Jr., 1960b, J. Appl. Phys. 31, 259S. 88B, 1239.
Blocker, T.G. and A.J. Heeger, 1967, J. Appl. Jonker, G.H., H.P.J. Wijn and P.B. Braun,
Phys. 38, 1111. 1956/57, Philips Techn. Rev. 18, 145.
Braden, R.A., J. Gordon and R.L. Harvey, Kasami, A. and S. Koide, 1966, J. Phys. Soc.
1966, IEEE Trans. MAG-2, 43. Japan, 21, 552.
Braun, P.B., 1957, Philips Res. Rept. 12, 491. Kasuya, T. and R.C. LeCraw, 1961, Phys. Rev.
Burlier, C.R., 1962, J. Appl. Phys. 33, 1360. 6, 223.
Bunget, I. and M. Rosenberg, 1967, Phys. Sta- Kerecman, A.J., A. Tauber, T.R. AuCoin and
tus Solidi, 20, K163. R.O. Savage, 1968, J. Appl. Phys. 39, 726.
Castelliz, L.M., K.M. Kim and P.S. Boucher, Kerecman, A.J. and T.R. AuCoin, 1969, J.
1969, J. Canadian Ceram. Soc. 38, 57. Appl. Phys. 40, 1416.
Cook, C.F. Jr., 1967, J. Appl. Phys. 38, 2488. Kimich, T.A., V.F. Belov, M.N. Shipko and
Dixon, S. Jr., 1962, J. Appl. Phys. 33, 1368. E.V. Korneev, 1970, Soviet Phys. Solid State,
Dixon, S. Jr., 1963, J. Appl. Phys. 34, 3441. 11, 1960.
Dixon, S. Jr., A. Tauber and R.O. Savage Jr., Kohn, J.A. and D.W. Eckart, 1964a, J. Appl.
1965, J. Appl. Phys. 36, 1018. Phys. 35, 968.
Douthett, D.D., I. Kaufman and A.S. Risley, Kohn, J.A. and D.W. Eckart, 1964b, Z. Kris-
1962, J. Appl. Phys. 33, 1395. tallogr. 119, 454.
Drobek, J., W.C. Bigelow and R.G. Wells, Kohn, J.A. and D.W. Eckart, 1971, Mat. Res.
1961, J. Am. Ceram. Soc. 44, 262. Bull, 6, 743.
440 M. SUGIMOTO

Kuntsevich, S.P. and V.P. Palekhin, 1978, Silber, L.M: and E. Tsantes, 1970, Proceedings
Soviet Phys. Solid State, 20, 1661. of ICF, 40.
Landuyt, van J., S. Amelinckx, J.A. Kohn and Simga, Z., A.V. Zalesskij and K. Zaveta, 1966,
D.W. Eckart, 1974, J. Solid State Chem. 9, Phys. Status Solidi, 14, 485.
103. Slonczewski, J.C., 1958, Phys. Rev. 110, 1341.
Licci, F. and G. Asti, 1979, IEEE Trans. MAG- Smirnovskaya, E.M., T.M. Perekalina and S.A.
15, 1867. Cherkezyan, 1978, Soy. Phys. Solid State, 20,
Lotgering, F.K., 1958/59, Philips Tech. Rev. 20, 1953.
354. Smit, J. and H.P.J. Wijn, 1959, Ferrites (Philips
Lotgering, F.K., 1959, J. Inorg. Nucl. Chem. 9, Technical Library, Eindhoven).
113. Stearns, F.S. and H.L. Glass, 1975, Mat. Res.
Lotgering, F.K., P.H.G.M. Vromans and Bull. 10, 1255.
M.A.H. Huyberts, 1980, J. Appl. Phys. 51, Streever, R.L., T.R. AuCoin and P.J. Caplan,
5913. 1971, J. Phys. Chem. Solids, 32, 519.
Mita, M., 1965, J. Phys. Soc. Japan, 20, 1599. Suemune, Y., 1972, J. Phys. Soc. Japan, 33, 279.
Mita, M., 1968, J. Phys. Soc. Japan, 24, 725. Suhl, H., 1956, Proc. IRE. 44, 1270.
Neckenbfirger, E., 1966, IEEE Trans. MAG-2, Tauber, A., S. Dixon Jr. and R.O. Savage, Jr.,
473. 1964, J. Appl. Phys. 35, 1008.
Neckenbtirger, E., H. Severin, J.K. Vogel and Tauber, A., J.S. Megill and J.R. Shappirio,
G. Winkler, 1964, Z. Angew. Phys. 18, 65. 1970, J. Appl. Phys. 41, 1353.
Neumann, H. and H.P.J. Wijn, 1968, J. Am. Tauber, A., R.O. Savage, Jr. and M.D. Gre-
Ceram. Soc. 51, 536. benau, 1971, J. Appl. Phys. 42, 1738.
Okazaki, K. and H. Igarashi, 1970, Proceedings Uitert, L.G. and F.W. Swanekamp, 1957, J.
of ICF, 131. Appl. Phys. 28, 482.
Perekalina, T.M., D.G. Sannikov and E.M. Verweel, J.,:1967a, Z. Angew Phys. 23, 200.
Smirmovskaya, 1979, Sov. Phys. Solid State, Verweel, J., 1967b, J. Appl. Phys. 38, 1111.
21, 1579. Vinnik, M.A., A.P. Erastova and Yu.G. Sak-
Reisch, F.E., R.W. Grant, M.D. Lind, G.P. sonov, 1968, Fiz. Tverd. Tela, 8, 269; 1966,
Espinosa and I.B. Goldberg, 1975, IEEE Sov. Phys. Solid State, 8, 219.
Trans. MAG-11, 1256. Voekov, D.V. and A.A. Zheltukhin, 1980, Izv.
Robbins, M., S. Lerner and E. Banks, 1963, Akad. Nauk. SSSR, 44, 1480.
Phys. Chem. Solids, 24, 759. Wijn, H.P.J., 1952, Nature, 170, 707.
Savage, R.O. and A. Tauber, 1964, J. Am. Weiner, M. and S. Dixon, Jr., 1970, IEEE
Ceram. Soc. 47, 13. Trans. MAG-6, 397.
Savage, R.O. Jr., S. Dixon Jr. and A. Tauber, Yamzin, 1.I., R.A. Sizov, I.S. Zheludov, T.M.
1964, J. Appl. Phys. 36, 873. Perekalina and A.V. Zalesskii, 1966, Ah.
Savage~ R.O., A. Tauber and J.R. Shappirio, Eksperim. i Teor. Fiz. 50, 595; 1966, Sov.
1974, AIP Conf. Proc. 34, 491. Phys. JETP, 23, 395.
Schl6mann, E., R.I. Joseph and I. Bady, 1963, Zalesskii, A.V. and T.M. Perekalina, 1965, Zh.
J. Appl. Phys. 34, 672. Eksperim, i Teor. Fiz. 48, 48; 1965, Soy.
Silber, L.M. and E. Tsantes, 1969, IEEE Trans. Phys. JETP, 21, 64.
MAG-5, 6O0.
chapter 7

HARD FERRITES AND


PLASTOFERRITES

H. STABLEIN
Fried. Krupp GmbH
Krupp WlDIA
Krupp Forschungsinstitut D 4300 Essen
F.R.G.

Ferromagnetic Materials, Vol. 3


Edited by E.P. Wohlfarth
© North-Holland Publishing Company, 1982

441
CONTENTS

1. G e n e r a l . . . . . . . . . . . . . . . . . . . . . . . . . . . 443
1.1. H i s t o r i c a l d e v e l o p m e n t . . . . . . . . . . . . . . . . . . . . 443
1.2. A s p e c t s for a p p l i c a t i o n s as p e r m a n e n t m a g n e t m a t e r i a l . . . . . . . . . 444
1.3. C o m p o s i t i o n s a n d p h a s e d i a g r a m s . . . . . . . . . . . . . . . . 449
1.3.1. B a O - F e 2 0 3 s y s t e m . . . . . . . . . . . . . . . . . . . 451
1.3.2. B a O - F e 2 0 3 - b a s e d s y s t e m s . . . . . . . . . . . . . . . . . 454
1.3.3. S r O - F e 2 0 3 s y s t e m . . . . . . . . . . . . . . . . . . . 457
1.3.4. S r O - F e 2 0 3 - b a s e d s y s t e m s . . . . . . . . . . . . . . . . . 458
1.3.5. P b O - F e 2 0 3 s y s t e m . . . . . . . . . . . . . . . . . . . 459
1.3.6. B a O - S r O - P b O - F e 2 0 3 - m i x e d s y s t e m s . . . . . . . . . . . . . 461
1.3.7. C a O - F e 2 0 3 - b a s e d s y s t e m . . . . . . . . . . . . . . . . . 462
2. M a n u f a c t u r i n g t e c h n o l o g i e s of h a r d ferrites . . . . . . . . . . . . . . 462
2.1. U s u a l t e c h n o l o g y . . . . . . . . . . . . . . . . . . . . . . 462
2.1.1. R a w m a t e r i a l s ; m a i n c o m p o n e n t s a n d a d d i t i v e s . . . . . . . . . 464
2.1.2. M i x i n g ; g r a n u l a t i o n . . . . . . . . . . . . . . . . . . . 468
2.1.3. R e a c t i o n s i n t e r i n g ; i n t e r m e d i a t e p r o d u c t s . . . . . . . . . . . 472
2.1.4. P r e p a r a t i o n of m o u l d a b l e p o w d e r . . . . . . . . . . . . . . 480
2.1.5. D i e p r e s s i n g ; o r i e n t i n g field . . . . . . . . . . . . . . . . 489
2.1.6. Final s i n t e r i n g ; s h r i n k a g e ; grain g r o w t h . . . . . . . . . . . . 497
2.1.7. M a c h i n i n g (grinding etc.) . . . . . . . . . . . . . . . . . 508
2.1.8. M a g n e t i z i n g a n d d e m a g n e t i z i n g . . . . . . . . ~ . . . . . . 510
2.2. Special t e c h n o l o g i e s . . . . . . . . . . . . . . . . . . . . . 513
2.2.1. Single s i n t e r i n g t e c h n i q u e . . . . . . . . . . . . . . . . . 514
2.2.2. P r e c i p i t a t i o n t e c h n i q u e s . . . . . . . . . . . . . . . . . 517
2.2.3. M e l t i n g t e c h n i q u e s . . . . . . . . . . . . . . . . . . . 519
2.2.4. F l u i d b e d a n d s p r a y t e c h n i q u e s . . . . . . . . . . . . . . . 525
2.2.5. H o t p r e s s i n g a n d hot d e f o r m a t i o n t e c h n i q u e s . . . . . . . . . . 526
2.2.6. R o l l i n g a n d e x t r u s i o n t e c h n i q u e s . . . . . . . . . . . . . . 527
2.2.7. P r e p a r a t i o n of thin l a y e r s . . . . . . . . . . . . . . . . . 535
3. T e c h n i c a l p r o p e r t i e s of h a r d ferrites . . . . . . . . . . . . . . . . . 535
3.1. M a g n e t i c c h a r a c t e r i s t i c s at r o o m t e m p e r a t u r e ; s t a n d a r d i z a t i o n . . . . . . 535
3.2. I n f l u e n c e of t e m p e r a t u r e o n m a g n e t i c p r o p e r t i e s . . . . . . . . . . . 550
3.3. I n f l u e n c e of d e v i a t i o n from the p r e f e r r e d axis, of m e c h a n i c a l stress, a n d of n e u t r o n
i r r a d i a t i o n on m a g n e t i c p r o p e r t i e s . . . . . . . . . . . . . . . . 561
3.4. V a r i o u s physical a n d c h e m i c a l p r o p e r t i e s . . . . . . . . . . . . . . 567
3.5. C o m p a r i s o n w i t h o t h e r p e r m a n e n t m a g n e t m a t e r i a l s ; a p p l i c a t i o n s . . . . . 577
4. B o n d e d h a r d ferrites, p l a s t o f e r r i t e s . . . . . . . . . . . . . . . . . 582
4.1. M a n u f a c t u r i n g t e c h n o l o g i e s for p l a s t o f e r r i t e s . . . . . . . . . . . . 583
4.2. T e c h n i c a l p r o p e r t i e s a n d a p p l i c a t i o n s of p l a s t o f e r r i t e s . . . . . . . . . 585
References . . . . . . . . . . . . . . . . . . . . . . . . . . . 592

442
1. General

' H a r d ferrites' are permanent magnet materials on the basis of the phases
BaFe12019, SrFe12019 or PbFe12019 as well as their solid solutions, e.g.
Bal-xSrxFe~zO~9. After appropriate treatment in production they possess con-
siderable magnetic 'hardness', i.e., coercivity, in an order of magnitude of several
100 k A / m (several kOe)*. Being oxidic the materials are mechanically hard too,
but this is not what is meant by the term 'hard ferrites'. Chemically and
crystallographically the hard ferrites belong to the large group of hexagonal
ferrites which all exhibit hexagonal crystal structures of considerable similarity
and among which there are both hard as well as soft magnetic substances. Of
these only the hard ferrites on the basis of BaFea2019 and SrFelzO19 are of
considerable economic importance. Tonnage-wise they are the main permanent
magnet materials in use today, see fig. 77. In contrast, hard ferrites on the basis of
PbFe12019 are not in c o m m o n use for magnetic and, in particular, toxicological
reasons. 'Plastoferrites' are composite materials composed of hard ferrite powder
and plastics. The pulverized particles are embedded in the plastic and this makes
it possible to use the advantageous forming methods known in plastics technology
for hard ferrite materials as well.
This chapter mainly covers the production processes for, and properties of, hard
ferrites and plastoferrites and deals with them in detail in sections 2 to 4. Section 1
contains some remarks on their historical development and mentions com-
positional, crystallographic and magnetic aspects as far as this is necessary for a
direct understanding of sections 2 to 4. The fundamentals are covered in detail in
chs. 5 and 6 of Kojima and Sugimoto in this Volume.

1.1. Historical development

As mentioned in the introduction, the phases BaFe12019, SrFe12019 and PbFe12019


form the basis of the hard ferrite magnets. These can also be written in the form

*In this chapter SI units are used first, followed by CGS units, l kA]m ~ 12.57Oe~
12.50e = 100/80e; 1 mT ~ 10 G; 1 kJ/rn3~-0.1257MGOe ~ 0.125 MGOe = 1/8 MGOe; 100 kA/m ~-
1.257 kOe ~ 1.25kOe = 10/8kOe; 1 T ~ 10 kG.

443
444 H. ST]kBLEIN

MO.6Fe203 with M = Ba, Sr, Pb. Adelsk61d (1938) was the first to correctly
describe the composition and crystal structure of these phases. H e thus
established a link with the results of earlier work on the mineral magnetoplumbite
(Aminoff 1925, who also proposed the name; Blix 1937), the structure of which is
isomorphous with those of the compounds mentioned above and whose com-
position was described by Berry (1951) as PbFeT.sMn3.sA10.sTi0.5019. Some of the Fe
ions are substituted by manganese, aluminium and titanium. Aspects relating to
the structure of these substances and their relation to similar lattice structures of
the/3-A1203 type (Na20"llA1203) or the aluminates MO.6A1203 (M = Ca, St, Ba)
were of particular interest at the time, while the magnetic properties hardly
aroused attention. Only the high magnetizability of magnetoplumbite was men-
tioned (Aminoff 1925, Flink 1924) and this characteristic was utilized for separat-
ing magnetoplumbite from non-magnetic admixtures (Blix 1937). In this con-
nection it should be mentioned that Hausknecht (1913) had already experimented
with mixtures of iron oxides and barium oxide and discovered an abnormally high
magnetism in annealed specimens roughly composed of BaO + 5Fe203 but did not
look further into the question of structure. Further details on the manufacture of
BaFe12019 from mixtures of Fe203 and BaCO3 were given by Erchak et al. (1946)
to the effect that t h e compound forms during annealing at a temperature above
around 750°C in an oxygen atmosphere. The occurrence of the compound was
proven with atomic ratios Fe/Ba ~ 4 with emphasis on Fe/Ba ~ 12.
Initially merely of an academic nature, interest in this compound changed
rapidly when from about 1950 its magnetic properties were investigated more
closely. At that time the soft magnetic (cubic) spinel ferrites developed at Philips
were already known (Snoek 1947). In the further course of this work the
considerable crystal anisotropy of BaFe~2Oi9 was discovered and a technology
developed which permitted it to be utilized as a permanent magnet material
(Went et al. 1952, Rathenau et al. 1952). Independently of this and almost at the
same time, researchers at Krupp discovered the hard magnetic properties of this
material and put them to use (Fahlenbrach 1953, Fahlenbrach et al. 1953). Since
then hard ferrites have been increasingly used as permanent magnets throughout
the world, not only in electrical engineering. They proved particularly suitable in
the area of magnet mechanics where the forces set up by magnetic fields are
utilized for adhesion, attraction, repulsion, clamping, rotating, etc. On the work
done in parallel by researchers, manufacturers and users in the field of hard
ferrites there is a vast amount of literature. Some of this is referred to in the
pertinent sections below. Further information is found in chapters 5 and 6 by
Kojima and Sugimoto in this Volume.

1.2. Aspects for applications as permanent magnet material

The potential suitability and actual use of any material for a specific purpose
depend on a number of physical and economic factors. For permanent magnet
applications in particular the material is expected to exhibit the following proper-
ties:
HARD FERRITES AND PLASTOFERRITES 445

(a) Magnetic saturation polarization* Js should, at least up to r o o m t e m p e r a t u r e


and preferably appreciably above r o o m temperature, be as large as possible.
(b) Coercivity must be sufficient for the same t e m p e r a t u r e range. T o be able to
achieve this, there must be preferred directions for the spins which correspond to
a sufficiently large minimum of free energy so that the spins are coupled to these
directions.
(c) For the same t e m p e r a t u r e range the material should be stable both struc-
turally and chemically.
(d) T h e cost of the material and of production should be low in relation to the
characteristics attainable. This means in particular that starting materials needed are
available in adequate quantities, are of sufficient purity and easy to handle, that the
composition of the material in terms of main and accompanying constituents and the
individual stages of production in terms of level and variations of temperature, time,
pressure and atmosphere, the required aids etc. are sufficiently inexpensive and
non-critical and that suitable forming facilities are available.
For none of these aspects is there a clear-cut, generally applicable limit governing
suitability or non-suitability as a p e r m a n e n t magnet material. Which material is best
suited to any particular application depends on how all the attendant circumstances
are taken into account.
The question that concerns us here is how hard ferrites m e e t requirements (a)
to (d).

(a) Saturation polarization of the c o m p o u n d BaFe12019 at r o o m t e m p e r a t u r e is


described by Stuijts et al. (1954) as Js = 0.475 T (4.75 kG). It is thus considerably
below that of iron (2.15T) or that of AlNiCo alloys (up to about 1.4T). In
commercial hard ferrites the value is up to about 10% lower owing to the porosity
of the samples, the presence of non-magnetic phases and, perhaps, partial
substitution of iron ions by, for instance, aluminium. T h e r e has thus been no lack
of attempts at finding substitutes to increase the saturation polarization but as yet
without appreciable success (Schieber 1967, Asti 1976). An exact comparison of
the values given in literature as saturation polarization is often problematical,
however, if the composition and homogeneity of the specimens are not exactly
defined and an adequate field strength was not allowed for in measuring.
In fig. 1 the saturation polarization of BaFe12019 is shown as a function of
t e m p e r a t u r e (Smit et al. 1959). The Curie t e m p e r a t u r e of about 450°C is still
relatively high but at about -0.2% /K the slope of the curve is unusually steep.
The resultant change in the magnetic flux can be very undesirable in certain
applications. Suitable substitutes were therefore tried to improve t e m p e r a t u r e
variation (Heimke 1960, Esper et al. 1972, 1975, H a n e d a et al. 1975). This aim has

* In this chapter the term "magnetic polarization" or, shortly, "polarization" J is used as defined by
the International Electrotechnical Commission in IEC Publication 50(901), Advance edition of
International Electrotechnical Vocabulary, 1st ed., 1973 (term no. 901-01-13), It is connected to
"magnetization" M (term no. 901-01-07) and "magnetic flux density" B (= "magnetic induction"; term
no. 901-01-03) by J =/x0M = B -/*oH, where magnetic constant/z0 = 4~- × 10 vH/m (term no. 901-01-
11),
446 H. ST~BLEIN

150
mT.crn3
g --".,,.

4
7 5o

, i I
\
-273-200 o 200 °C ~oo

Fig. 1. Saturation polarization Yjsample density p of BaFe12019as a function of temperature (Smit et al.
1959).

been essentially achieved but only by way of incurring other inevitable drawbacks
that have so far prevented practical use. The magnitude and temperature varia-
tion in the saturation polarization of hard ferrites therefore do not compare very
favourably with other permanent magnet materials.

(b) The hard ferrites have a relatively high crystal anisotropy energy Ek which for
these hexagonally crystallizing substances can be simply expressed as

Ek = K sin 2 0.

In the Weiss domain considered, 0 is the angle between c-axis and polarization Js
(see fig. 39) and K is the anisotropy constant. According to Rathenau et al. (1952),
at room temperature it has roughly the value K = + 3 × 1 0 5 J / m 3 ( + 3 ×
106 erg/cm3), which means that the c-axis is the preferred direction for the spins
and re-magnetization of a domain should take place at a reverse field (anisotropy
field HA) of 2K/Ys = 1260 kA/m (15.8 kOe), provided that re-magnetization occurs
in a single-domain particle by coherent rotation of the spins. Assuming the same
conditions in a texture-free specimen, polarization should disappear wher~ there is
a counter field of 0.96K/Js = 600 kA/m (7.6 kOe). Coercivities of up to 80% of
this value have in fact been found; however, only in very fine particles of around
0.1 p~m (Mee et al. 1963, Haneda et al. 1973a, Gordes 1973). In commercial
specimens the crystallites are present in sizes of around 1 ~m, and only smaller
jHc values are attained. In fig. 2, K, and in fig. 3, 2K/Ys and 0.96K/Js are plotted as
a function of temperature (Rathenau et al. 1952). From the latter the coercivity
might be expected to be largely independent of temperature. In fig. 3, however, it
can be seen from the jH~ curve of a specimen having crystallites of around 1 p~m
that this is not the case, the temperature coefficient of this specimen being about
HARD FERRITES AND PLASTOFERRITES 447

3
m3

i i i ~ - -
-200 0 200 °C400
T
Fig. 2. Crystal anisotropy constant K of BaFe12019 as a function of temperature (Rathenau et al.
1952).

1600 20
k__~A
I
m H = 2___~K kOe

1200 15

800 10
HA

~00 5

'o c 6000
T
Fig. 3. Anisotropy field HA = 2K/Js for anisotropic and HA = 0.96K/Js for isotropic BaFe12019 and
coercivity jHc of fine-grained sintered specimens (~1 p,m) as a function of temperature (Rathenau et al.
1952).

+0.4% /K at room temperature and thus fairly high. The temperature dependence
of jHc in specimens with crystallites of varying sizes was also measured by Sixtus
et al. (1956). The deviation of the quantities 0.96K/Js and ]/arc from one another is
explained by the formation and shift of domain walls during re-magnetization
(Goto et al. 1980).
Although the coercivity jHc of commercial hard ferrites is thus appreciably
below the theoretical value, it is attractively high enough for a large number of
448 H. ST,~d3LEIN

applications. It was several times higher than that of the AlNiCo materials then
commonly used. The resultant stability of the magnetic flux meant that the hard
ferrites could be rapidly introduced into dynamic permanent magnet applications
with very beneficial consequences. A more detailed description of Js and K of the
various hexaferrites is given in sections 3.1 and 3.2.

(c) A stable structure can by no means be taken for granted in permanent magnet
materials. In a number of materials the structure at room temperature is in a
'frozen', metastable state which by diffusion or allotropic transformation can
change towards equilibrium as a result of a temperature rise, for instance, this
having an adverse effect on the magnetic values, especially the coercivity. This
phenomenon is termed 'structural ageing' which in most materials occurs far
below the Curie temperature and thus imposes restrictions on the actual use of
the magnets. Hard ferrites are an exception in this respect. Their structure remains
stable far beyond the Curie t e m p e r a t u r e - i n Ba- and Sr-based hard ferrites up to
temperatures of more than 1400°C (in air) before oxygen is released and phase
transformations occur; cf sections 1.3.1 and 1.3.3.
Chemical stability prevails when the material does not react with the ambient
medium. Naturally, the type of reaction and its extent depend on the medium. As
the cations of the BaFe12Oi9 and SrFe120~9 lattices are in the highest state of
oxidation, these materials are stable and not liable to oxidize which is an
advantage over metallic materials. Data on the stability of the hard ferrites in
different media are given in section 3.4.

(d) With hard ferrites the costs of materials and production are lower than in
other grades of permanent magnets if related to the same energy content. A
number of favourable circumstances contributes to this, some of which may be
briefly mentioned here. Detailed technological information is given in section 2.1.
The raw materials required are inexpensive, abundantly available and easy to
handle. Extreme purity is not necessary. As a rule the iron oxides used have 0.5 to
1% by weight of impurities, and natural hematite (a-Fe203) or iron oxides from
wastes can also be used. Powder metallurgical treatment is effected by the
time-tested methods used in the ceramics industry and, in the case of the
plastoferrites, in the plastics industry. The temperatures necessary for the raw
materials to react to form ferrite ('calcination') and for sintering are normally
around 1200°C to 1300°C. Both processes take place in air, while with metal
powders a protective atmosphere or vacuum is always required. For the high-
grade ferrites a powder consisting of single-domain particles in sizes of 1 ~zm or
less is needed, a fineness which can easily be adjusted by grinding in water. With
metal powders, however, water would not be suitable as a grinding fluid so that as
a rule organic liquids have to be used which, in view of pollution control
requirements and for health reasons, are more difficult to handle. The powder is
generally shaped by one-sided or two-sided pressing in a die and the plastoferrites
by injection moulding, calendering, etc., i.e., by well-known processes.
General information and special aspects of the requirements (a) to (d) are
HARD FERRITES AND PLASTOFERRITES 449

described, for example, in the books and papers of Von Aulock (1965), Becker (1962),
Gmelin (1959), Heck (1967), Heimke (1976), Landolt-B6rnstein (1962, 1970, 1981)
and Smit et al. (1959) as well as in the books mentioned at the end of section 3.5.

1.3. Compositionsandphasediagrams
A number of relevant phase diagrams is discussed in this section. The in-
formation from the diagrams is, among other things, useful in the following
questions and problems:
(1) The diagrams furnish information on the existence ranges of the hexaferrite
phases MO.6FezO3 (M, e.g. Ba, Sr, Pb) in terms of composition, temperature and
oxygen partial pressure.
(2) They furnish information on further phases MO.nFe203 (n = molar ratio of
Fe203 and MO) which may form as intermediates ('precursor phases') in the
reaction of the raw materials or which are present after the reaction in equili-
brium with the hexaferrite phase. Some of these phases play an important role for
production and properties.
The emphasis is placed on the quasi-binary systems MO-Fe203 because these
are relatively easy to discuss and represent good approximations even for more
complex conditions. Such conditions obtain in practice because apart from the
main constituents MO and Fe203 the magnet specimens contain further sub-
stances introduced intentionally or otherwise which are important and sometimes
decisive for the production process and properties. The relevant quasi-ternary and

1600
oC
Liq!id ~ l@" !N ' ..
1565 °

l"°°,s .'- 7
T
1200 l BF

I"e " i 1'.BF"+,,M"

1000 11 II
8°°2 '1 60 ' 80 LI ~ 1o0
|
B2F BF PI %5
rnol %
Fig. 4. Phase diagram of BaO-Fe203 (Ooto et al. 1960). Atmosphere: 02 above and air below
eutectic temperature.
450 H. STJid3LE1N

higher phase diagrams are therefore also mentioned. This section concentrates on
the compositions that are significant for magnet engineering and magnet ap-
plications, i.e., on the side of the systems that is rich in Fe=O3.
In connection with the phase diagrams there is of course the question as to the
possible substitutes in the anion and cation sublattices of the hexaferrites. Here,
too, only the most important are mentioned; more detailed information is given in
chapters 5 and 6 of Kojima and Sugimoto in this Volume.
The following abbreviations are used in the phase diagrams and in the text:
B - = B a O ; S-= SrO; P - = P b O ; F-~Fe203; M--= (Ba, Sr, Pb)O.6Fe203 (mag-
netoplumbite), if need be by BaM, SrM and PbM; as to X and W see table 2.

1600
oC

l~lot5 o ,..¢'_ _7%50°


I
/
I

I
I c~-B F+ M
I
1200 I
I
I1~5oz5o
I i
I~-BF .BF~

I000 Ip_BF+BE

'r--+:
800 _ _ ~ _ - B F * B ~
IF _F i

~o I 0•
60 80 100
B~ B5 M ~2o3
too/% =
Fig. 5. Phase diagram of BaO-Fe203 (Batti 1960). Atmosphere: 02 for liquids, otherwise air.

1600 ~-

3
,26o~ ~BF-+L~ 1338
° M+j

,2o, ' W-I / ' ,oo

me/ %
Fig. 6. Phase diagram of BaO--Fe203 (Ziolowski 1962). Atmosphere: air.
HARD FERRITES AND PLASTOFERRITES 451

Phase designations in inverted commas mean that owing to an appreciable


homogeneity range the actual composition may deviate from the one indicated in
the formula.

1.3.1. B a O - F e 2 0 3 system
T h e systems and subsystems mentioned by G o t o et al. (1960), Batti (1960),
Ziolowski (1962), Van H o o k (1964), Sloccari (1973) and A p p e n d i n o et al. (1973)
are given in figs. 4-12. They do not correspond in all details. T h e following may
be stated:

H o m o g e n e i t y range. According to Van H o o k (1964) BaFe12019 is in terms


of oxygen content and cation ratio stoichiometrically at least within

1600
oC
Liquid
1495° /
B2F L+B~. .~. 14550/.. t ..xl,. l-x÷%q
1400 --/L +M
..... 9F÷L ~,Z/ 1315+..5°

B2F+ BF BF+M
1200

20 40 60 ao II too
B2F BF Mxw %03
rnol %
Fig. 7. Phase diagram of BaO-Fe203 (Van Hook 1964), isobaric projection with 02 atmosphere of i bar
(the compounds Fe304, X and W also contain Fe2+).

1600 , Liquid ~ """


0C . s . ~ ~''~

BF+ L L .~ "" "" L+Fe304

I ~00
M÷Fe3q

sF÷w iT - w÷%£
8F÷ X
1200 - -
M'FsO~
M
BF+ M F ÷

I000
60 80 I00
BF MXW Fe203
tool %
Fig. 8. As fig. 7, but with COz/CO atmosphere (1OO2 ~ 0.01 bar).
452 H. ST]/d3LEIN

Baa+0.04Fe12019_+0.03 between room temperature and 1450°C at Po2 = i bar. Other


authors who examined specimens made in air or oxygen arrived at the conclusion too
that there is only a very small homogeneity range (Batti 1960, Ziolowski 1962,
St~blein et al. 1969, Reed et al. 1973, Lacour et al. 1975a and Sloccari et al. 1977a).
According to Goto et al. (1960) however, the homogeneity range extends from
BaO-6Fe203 to BaO-4.5Fe203 (1350°C) and to BaO.5Fe203 (800°C) respectively.
The cause of this discrepancy has not yet been reliably clarified. It is conceivable,
however, that the specimens deviating from the stoichiometric composition were of
the two-phase type and that the BaO.Fe203 phase was overlooked. This is possible
because its presence, particularly in small amounts, is difficult to verify or not at all
verifiable radiographically and because it easily dissolves in concentrated muriatic
acid so that it cannot be detected microscopically in specimens thus prepared.

1600
I
oC z/L +
Liquid , / Fe203
f~
I

1~O0
" F'~,~ //.L+M
I L+BF~ ~ .
M
BF+M
Fe203
1200

i i

~0 60 80 I00
BF M Fe203
tool %

103
riK]
O.52

0.5~
Fig. 9. As fig. 7, but with 02 atmosphere of 40 bar.

l°c{
1650
i ~6oo
1550
0.56

Q58
M=
X+L~ 1500 E
1450

(2.60 1400
i i

OOt o7 7 lb bar 76g


Oxygen pressure
Fig. 10. Decomposition of barium hexaferrite M as a function of oxygen pressure and temperature
compared with that of Fe203 (Van Hook 1964).
H A R D FERRITES AND PLASTOFERRITES 453

,BF+ M
1200
oC ii150o+10o

tO00
BF
+ BF,M
55 2
800

dO 6,0 80 100

rnol % - - ~
Fig. 11. Phase diagram of BaO-Fe203 (Sloccari 1973). Atmosphere: air.

1200 r - -
oC ' 1 j--' !
11100 c~-BF+M I

1000
0- Fy3 i 0ol
800 I
~o I ~ 80 I00
M Fe 0
BF z %5 23
tool %

Fig. 12. Phase diagram of BaO--Fe203 (Appendino et al. 1973). Atmosphere: air.

Recently, the existence of a metastable compound BaO.nFezO3 with n =


4.0 to 5.8 at 600°C and n ~-- 16/3 at 900°C was reported (Durant et al. 1980, 1981),
decomposing at 950°C and having nearly the same lattice parameters as
BaO.6Fe203 (see table 31).
BaFe12Oa9 is unstable at elevated temperatures. At a pressure of Po2 ~ 0.01 bar
the hexaferrite, as can be seen in fig. 8, releases 02 and disintegrates into the
phases BaFelsO23 ('X-phase'; BaO-FeO.7Fe203) and BaFe204; at Po2 = 1 bar there
is, as can be seen in fig. 7, a peritectic reaction attended by the formation of the
X-phase, and it is only at Po2 = 40 bar, as can be seen in fig. 9, that (almost)
congruent melting takes place at about 1500°C (Van H o o k 1964, 1976)o This latter
fact deserves attention in making BaFe12019 monocrystals from the melt, cf.
Menashi et al. (1973). Figure 10 gives the decomposition temperature as a function
of pressure Po2. In air it is around 1430°C, a somewhat smaller value than
according to the phase diagrams in figs. 5 and 6. Towards lower temperatures the
hexaferrite lattice is completely stable as all experience gathered so far has shown.
454 H. STJLBLEIN

Neighbouring phases. All phase diagrams correspond in that on the side richer in
Fe203 no further B a - F e - o x i d e phases occur below 1250°C, i.e., the two-phase field
(BaFe12019 + c~-Fe203) occurs there, cf. figs. 4, 5 and 7 to 9. Conditions on the side
of BaFe12019 richer in B a O are not so clear, cf. table 1. BF, BzF3 and BF2 are
mentioned as neighbouring phases. As can be seen from figs. 6 and 12, BF occurs
in several modifications (Ziolowski 1962, Meriani 1972, Appendino et al. 1973,
Haberey et al. 1974). BaFe12019 and BaFe204 (BF) form an eutecticum for which
eutectic temperatures between 1315 and 1370°C are given. The two phases BzF3
and BF2 occur only up to 1150°C maximum and the adjustment to equilibrium
proceeds very slowly (Ziolowski 1962, Sloccari 1973, Appendino et al. 1973) which
perhaps explains why these phases were not found by other workers.

1.3.2. BaO-Fe20~-based systems


As mentioned above, the hexaferrite decomposes at elevated temperatures and
oxygen partial pressures that are not excessively high to form the 'X-phase'
BaO.FeO.7Fe203, cf. figs. 7, 8 and 10, at which one of every 15 Fe ions is in a
two-valued state. T o represent this compound in the quasi-binary system B a O -
Fe203 is therefore incorrect. Figure 13 shows the Fe203-rich portion of the
B a O - F e 2 0 3 - M e O system, with Me meaning a divalent cation, e.g., a 3d-element,
i.e., particularly Fe 2+. The letters correspond to certain stoichiometric com-
positions explained in detail in table 2. Compounds of the X-, W-, Z- and Y-type
were described by Braun (1957), cf. Smit et al. (1959). Later, further compounds
were found which range between M and X or between U and Y (Kohn et al.
1971). In total the structurally allied compounds lying on the straight lines M - W
and M - Y are identical with the designation 'hexagonal ferrites'. Also stackings of
mixed hexagonal ferrites are known and were investigated by means of high
resolution electron microscopy and diffraction (Hirotsu et al. 1978, Van Tendeloo
et al. 1979). Investigations into the B a O - F e 2 0 3 - F e O system were carried out by
Batti (1961a). In the system BaO-FezO3-ZnO, Slokar et al. (1978a) found
the corresponding X-, W-, U-, Z- and Y-phases at 1200°C in air. An outline
of various systems is given by Batti (1976). A material with composition
3BaO-4FeO-14Fe203 was reported by Brady (1973) and Durant et al. (1981).
G o t t o (1972) and Mansour et al. (1975) compiled further B a - F e - o x i d e com-
pounds containing Fe 4+ or Fe 6+ in addition to Fe 3+. A comprehensive review on
the BaO-Fe~O3-FeO system and on crystallographic and magnetic data of the
compounds involved was given by Sch6ps (1979). Batti (1961b) carried out
investigations into the BaO-FezO3-AI203 system at i bar 02 and 1300°C. It was
found that iron can be substituted in the compound BaFea2019 up to the com-
position BaFe4.zA17.8019.
Only a few investigations were made to find out whether Ba can be substituted
by Ca. According to Van Uitert (1957) up to 70 at.% of Ba can be replaced
without changing the lattice type. According to investigations by Sloccari et al.
(1977a) into the B a O - F e 2 0 3 - C a O system, there is a solid solution between
BaFe12019 and a hypothetical CaFesO13 at 1100°C in air up to a molar ratio of
about 1.75 : 1. In a later paper the homogeneity range was defined more precisely
~2 tt~ tt~ I I
v

III 0

© (.q
~7 I I
c~

©
6
III

.r"0
;>

t~

.r.

0
e'~
,,.., C,
'=~ A A
.=Z

©
Z

e~

t'~
0

<

455
456 . H. ST~d3LEIN

I00~

9O

tool % BaO

Fig. 13. Quasi-ternary system FezO3-BaO.Fe203 (BF)-MeI10-Fe203 (S). Symbols explained in table 2.

TABLE 2
Compounds of the quasi-ternary system BaO-Fe203-MeO, Me--Divalent cation,
e.g., Fe z+, Zn 2+.

Stoichiometric composition
Compound (mol %)

Symbol Formula BaO MeO Fe203

S 2(MeO.Fe203) - 50 50
BF BaO-Fe203 50 - 50
T (hypothetical) BaO.2Fe203 33.3 - 66.7

M BaO.6FezO3 14.3 - 85.7


M6S 2(3BaO-MeO. 19F~O3) 13.04 4.35 82.61
M4S 2(2BaO.MeO. 13Fe203) 12.50 6.25 81.25
X (M2S) 2(BaO.MeO.7Fe203) 11.1 11.1 77.8
W (MS) BaO.2MeO.8Fe~O3 9.1 18.2 72.7

U (M2Y) 2(2BaO.MeO.9Fe~O3) 16.7 8.3 75.0


Z (MY) 3BaO.2MeO. 12Fe203 17.6 11.8 70.6
Y 2(BaO.MeO.3Fe203) 20.0 20.0 60.0
HARD FERRITESAND PLASTOFERRITES 457

(Lucchini et al. 1980a) and established that the primary magnetic properties were
not changed substantially by the Ca substitution (Asti et al. 1980). According to
Kojima et al. (1980) (CaO)l_x(BaO)x'n(Fe203) has a solubility range of x =
1.0 to 0.6 and n = 5.0 to 5.6. Isotropic magnets with Br = 222mT, (BH)ma~=
10.2 kJ/m 3 (1.28MGOe), BHc = 121 kA/m (1.52 kOe) and 1He = 168 kA/m
(2.12 kOe) were prepared. The subject is also dealt with in section 1.3.7.
The BaO-Fe203-SiO2 system is of particular interest owing to the usual
addition of SiO2 in the commercial manufacture of permanent magnets. Haberey
(1978) and Haberey et al. (1980a) furnished a tentative diagram for air atmosphere.
At 1250°C, up to 0.55% by weight of SiO2 dissolves in BaFe120~9. Any surplus
forms a second glassy phase which is rich in SiO2, has a melting point of about
1050°C and promotes sintering while impeding grain growth, cf. section 2.1.6.
St/iblein (1978) too, found a glassy phase of very similar composition.

1.3.3. SrO-Fe203 system


This system was examined by Batti (1962a) in I bar 02 and by Goto et al. (1971) in
air. Their findings are shown in figs. 14 and 15, respectively. Both diagrams agree
very well.
The homogeneity range is very narrow and in the eutectic range somewhat
enlarged, at most towards the side rich in SrO (Routil et al. 1974). Towards higher
temperatures incongruent melting occurs at 1448°C (1 bar 02) and 1390°C (air),
with the W-phase SrFe18027 (=SrO.2FeO.8Fe203) being formed. Haberey et al.
(1976) likewise observed the formation of the W-phase in annealing in air above
1300°C, while in vacuum annealing above ll00°C Fe304 and 87F5 (or $4F3)formed
with the release of 02. In contrast to the behaviour of BaFe12019 (figs. 7, 8, 10), in
8rFe12019 the corresponding X-phase was only found as an intermediate product
(Goto 1972). Towards lower temperatures SrFe12019 is stable according to
experience hitherto gained.

L'+SF., Liquid
1600 ~~ 1600°t100
oC
lSO0o ~ ~ > , 1520°'10° f
1
1448°?_10°
1400

I
J L÷M M
÷

Fe203
1210o+_10°
1200
S F+M
, , s; "55 :L 75
i

0 20 60 80 I00
Fe203
rnol ~ =
Fig. 14. Phase diagramof SrO-Fe203 (Batti 1962a). Atmosphere: 02.
458 H. STJ~d3LEIN

1600 i

oC

l~O0

SrFe(}-x'L---1225+-~ • L÷M

1200 55.L
M
6"0203
I000

8O0
i 'l

0 20 ZO 60 80 DO
SrO SrFe03x 5 5 M Fe25
- rnol %
Fig. 15, Phase diagram of SrO-Fe203 (Goto et al. 1971). Atmosphere: air.

Towards the Fe203-richer side the two-phase region (SrFe12019+ o~-Fe203)


follows analogous to the BaO-Fe203 system.
On the SrO-richer side the phases S7F5 and $3F2 are given in figs. 14 and 15 as
neighbouring phases, both of them being very close to the composition $4F3
mentioned by Kanamaru et al. (1972). The eutectic temperatures of 1210°C (1 bar
02) or 1195°C (air) as well as the eutectic contents of 53.8 or 55 mole % Fe203 are
close to one another. An SF phase analogous to BF does not Seem to exist (Routil
et al. 1974, Haberey et al. 1976, Vogel et al. 1979a).

1.3.4. SrO-Fe2OB-based systems


Very little has become known on investigations into the SrO-Fe203-MeO system.
As mentioned above, the occurrence of the X- or W-phase was observed
especially for Me = Fe 2+. In the S r O - F e 2 0 3 - Z n O system Slokar et al. (1978b)
found the corresponding X- and W-phases at ll00°C in air.
In the Sr-Fe oxide compounds containing proportions of F e 4+ a r e found more
frequently than in the B a - F e oxides (Brisi et al. 1969, Goto 1972). In fig. 16 the
compounds known from literature are compiled (Haberey et al. 1976). A review
was also given by Sch6ps (1979).
Investigations into the SrO-FeaO3-A1203 system at 1 bar 0 2 and 1200°C were
carried out by Batti et al. (1967); a complete solid solution between SrFe12019 and
SrAl12019 was found.
The SrO-FeRO3-CaO system was investigated by Lucchini et al. (1976). It was
found that at 1100°C in air there is a solid solution with M-structure between
HARD FERRITES AND PLASTOFERRITES 459

SrO

sue°g x
Sr. Fe 0 - 20
Sr Fe 0
SrFe03-~"/~'-"""~- / t S.Ze. o~
, '°2 h
J~ (SF)

Z/\/\ /\ /\

2(FeO2 ) 80 60 gO 20 Fe203
mol % 2(FeO2)
Fig. 16. Compositionsof the Sr-Fem-FeTM oxides mentioned in literature (Haberey et al. 1976).

SrFe120 m and a hypothetical CaFesO13 up to a molar ratio of about 2: 1. On


re-examination the result was not confirmed completely (Lucchini et al. 1980b).
The primary magnetic properties were not changed substantially by the Ca
substitution (Asti et al. 1980). According to Kojima et al. (1980)
(CaO)l-~(SrO)x.n(Fe203) has a solubility range of x = 1.0 to 0.6 and n =
5.0 to 5.6. Isotropic magnets with Br = 220 mT, (BH)max = 10.7 kJ/m 3
(1.34 MGOe), ~Hc = 153 kA/m (1.92 kOe) and jHc = 218 kA/m (2.74 kOe) were
prepared. The subject is also dealt with in section 1.3.7.
Investigations into the SrO-FeaO3-SiO2 system were carried out by Kools
(1978a), Kools et al. (1980), Haberey (1978) and Haberey et al. (1980a). At 1250°C
they found the maximum solubility of SiO2 in SrFe12Om to be 0.6 and 0.4% by
weight. Any additional SiO2 leads to the occurrence of phases which at usual
sintering temperatures are liquid and, similarly to barium hexaferrite (section 1.3.2),
promote sintering and impede grain growth, cf. section 2.1.6. The occurring phases
were also reviewed by Broese van G r o e n o u et al. (1979b).

1.3.5. P b O - F e 2 0 3 system
The phase diagrams in air given by Berger et al. (1957) and by Mountvala et al.
(1962) are shown in figs. 17 and 18, respectively.
Concerning the homogeneity range of the hexaferrite phase, there is only
moderate agreement. While Berger et al. (1957) give no appreciable homogeneity
range, Mountvala et al. (1962) have found such a range from PbO-5Fe203 to
460 H. ST,g~BLEIN

...:-"..J
1400 .,'
/J
/I #
/I
oc /
,' [ t

// / L"r?;÷%%
iI
Liquid , ,~._

!
1200 r / ,,
,PJ t !
/ ," !
!
Gelantineous /
•/
~ I
/ /L÷MI
/ I /
, ~%03
/ ' --- ~ pl
I I A~ "-- I
/
/7 f Viscous or /

\ \\ / Granular ,,,
/ solid
\ \ , / %F+W
\ \ / ~ i ¢../
x " 1/ "P F+PF
- \\ \ /LL%~I..... z: z/"
PbO÷P2F '°2 F÷ Fe203

I P~°:%5 ,
6000 20 ~o 60 8ol lw
P~o 5~ ~% M ~-~o~
mol % =
Fig. 17. Phase diagram of PbO-Fe203 (Berger et al. 1957). Atmosphere: air.

1~00
°C ! L+Fe203
/.. - _, 1315°
/ 1
1200 // L +.NI'" ]1
/I

looo - i o~,

"\\
/L÷%~- %~-.e% I"C%
.M'"
I
,oo~I PbO+L ,,~
\ I[ o I
, 750°
II1 " 7 6 0 ° I

i ..... ,,j..;~_o__j] -;~÷~%% ' "%"~~%%~oo


;'o--:,% ' ,o ,' ,'
Pbo %F P~ M ~o£~
tool % =
Fig. 18. Phase diagram of PbO-Fe203 (Mountvala et aI. 1962). A t m o s p h e r e : air.
HARD FERRITES AND PLASTOFERRITES 461

PbO.6Fe203. The results obtained by Cocco (1955) who found a solid solution
between the boundaries PbO-2.5Fe203 and PbO-5Fe203 show an even greater
variance. It should be taken into account, however, that according to Adelsk61d
(1938) only the composition PbO.6Fe203 explains the measured radiographic data
and density values, not, however, the composition PbO@Fe203, for instance. The
existence of the PbO-rich side of the PbO.6Fe203 phase therefore cannot be
regarded as being proven beyond doubt, especially since equilibrium adjustment
proceeds very slowly and the structures of PbO.6Fe203 and of the neighbouring
phase richer in PbO are very similar and so it is difficult to distinguish between
them radiographically.
The thermal stability of lead hexaferrite is rather small in comparison with that
of the Ba or Sr compound, namely 1250°C (Berger et al. 1957) and 1315°C
(Mountvala et al. 1962); one of the decomposition products is o~-Fe203. Nothing is
known about the presence of FeO-bearing compounds such as X, W, Y etc. in the
PbO-Fe203 system. The differences mentioned are probably attributable to the
high vapour pressure of PbO (Berger et al. 1957, Bowman et al. 1969). Towards
lower temperatures the phase diagrams shown give rise to confusion as they seem
to indicate the decomposition of the compound PbO-6Fe203 below 820° and
760°C, respectively. However, even after prolonged anneals of up to 1000 h in the
range from 650 to 850°C this could not be determined (Berger et al. 1957). The
diagrams should therefore be interpreted to the effect that no formation of
PbO.6Fe203 was observed from the starting materials at these low temperatures.
All workers unanimously mention o~-Fe203 as neighbouring phase for the side
poorer in PbO, but PzF and PF2 for the side richer in PbO, each forming
low-melting peritectics with PbO-6Fe203 at about 900 and 950°C, respectively.
This is of importance for the industrial manufacture of the Ba or Sr hard ferrites
when small additions of PbO are added to the raw mix as a flux.

1.3.6. BaO-SrO-PbO-Fe203-mixed systems


In view of the identical structure and the only slight difference between the lattice
constants of the compounds MO.6Fe203 (M = Ba, St, Pb) (maximum deviation
Ac/c = 6.5%~ after Adelsk61d (1938); see also table 31) it is obvious that complete
miscibility exists in the entire region of stoichiometric composition, cf. Goto
(1972). Special conditions may, however, occur on the side poorer in Fe203 when
the composition is not stoichiometric because the structure and molar ratio
FezO3/MO of the neighbouring phases differ depending on the type of oxide MO.
Batti (1962b) examined the compounds of the BaO-SrO-Fe203 system produced
at 1100°C and found that depending on the BaO/SrO ratio the phases BF (Ba can
in part be substituted by Sr), BSF2 (a small portion of SrO can be substituted by
additional BaO) and $7F5 o c c u r . Later on isothermal sections up to 1235°C were
investigated by Batti et al. (1976).
Batti et al. (1968) synthesized specimens of the BaO-SrO-FezO3-AI203 system at
1400°C in i bar O2 and found that iron can be largely substituted by aluminium in the
entire Bal-xSrxFe12019 region. Equal solubility of CaO was found in the entire range
of BaxSrl-xFe12019 (Sloccari et al. 1977b).
462 H. STJ~BLEIN

1.3. 7. CaO-Fe203-based system


In spite of the close chemical affinity of calcium, strontium and barium no
hexaferrite phase exists in the system CaO-Fe203 (Adelsk61d 1938), probably
because of the smaller ionic radius of Ca. However, a magnetic Ca hexaferrite
phase can be stabilized by the presence of at least 2 mol % La203 (Ichinose et al.
1963, Lotgering et al. 1980). From this material isotropic and anisotropic grades
exhibiting useful magnetic properties can be prepared (Yamamoto et al. 1978a,
1979a) even if La203 containing some Nd203 is used (Yamamoto et al. 1980). The
magnetic properties can be improved by substitution of Ca by Ba (Yamamoto et
al. 1979b) or by Sr (Yamamoto et al. 1978b).

2. Manufacturing technologies of hard ferrites

2.1. Usual technology

The principle underlying the usual manufacturing process is shown diagram-


matically in fig. 19.
The raw materials used are generally the barium and strontium carbonates as
well as natural and synthetic iron oxide o~-Fe203 (rarely magnetite Fe304). In
addition to these main constituents so-called additives such as SiO2, A1203 etc. are
used individually or combined in amounts of about 0.5 to 2.5% by weight. They
serve to control the reaction kinetics, shrinkage and grain growth (see section
2.1.6) but sometimes they also affect the primary magnetic properties of the
hexaferrite phase (saturation polarization, crystal anisotropy energy).
The raw materials are intimately mixed and, if required, granulated or briquet-
ted, and annealed at temperatures of between about 1000 and 1300°C in air
('reaction sintering', 'calcination'). Hexaferrite is thus produced more or less
completely as a reaction product. The reacted mass is crushed and ground to a
powder of sufficient fineness. There are several possibilities for further treatment
depending on which magnet grades are to be manufactured.
(a) Being anisotropic, the highest grades are obtained by wet compression
moulding in a magnetic field. For this purpose the aqueous suspension, whose
ferrite particles are, in the ideal case, single crystals and consist of a magnetic
domain, is poured into the mould cavity. A magnetic field is applied to align the
ferrite particles, thus producing a 'preferred direction' in the suspension. Com-
pression takes place in this state, removing most of the water.
(b) Not quite such high anisotropic grades are obtained by dry compression
moulding of ferrite powder in a magnetic field because with this process the
particles cannot be as easily aligned. The ferrite powder is obtained by removing
the water and drying the ground suspension. The resultant caking of the ferrite
particles impairs the directional effect of the magnetic field, which is the reason
why the dried mass has to be loosened.
(c) For the lowest (i.e. isotropic) grade the powder does of course not have to
be alignable in compression moulding. The dried powder is therefore turned into
an easier-to-process granulate which is compacted in dry condition.
HARD FERRITES AND PLASTOFERRITES 463

Section

J
(4 milling)
dry [ wet 2,1.2

Granulation or

of ready-to- _ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
press hard
powderferrite I Reaction sintering I I 2.1.3.
fromraw ............. 4 ....... ~ ............
materials I ,,r ,orushio ',
I Wetm""ng

O-con'ent
I' I
__/ Ioeogglomerotionl Gronu,ot,on
with binder
/
PressureFiltrati°n]
..........
I magnetic
Drypressingin I ~ Orypressing
in magnetic field ~ ................
field 4 2.1.5.
I~ onisotropic I~ anisotropic I I ~ isotropic |
I magnets l mag,',ets I I magnets I
. . . . . . . - -~ ~-q2 - ~ - - -..7--- ~ --~J- - . . . . . . . . . . .

Production
of shaped I Final sintering I 2.1.6.
from ...........
powderParts ......... [ Gri!ding - i )-21~7.--

Assemblage,
Magnetization ~ 2.1,8.

Fig. 19. Usual production technology of bulk isotropic and anisotropic hard ferrites.

Compression moulding provides a porous compact with a relative density of


about 60% of the radiographic density and only little strength. Indirect shaping by
machining the compact is therefore only possible to a very limited extent. In
subsequent finish-sintering the relative density increases to about 90 to 98%. The
attendant shrinkage (contraction) of the linear dimensions occurs parallel to the
preferred direction 1.5 to 2 times as great as in the direction perpendicular to it.
The compact is then much stronger but also brittle and can only be machined by
grinding, cutting, etc. If necessary, the bulk magnet is completed to form a magnet
system and magnetized. Under certain circumstances the sequence of these two
latter operations can be reversed.
This outline shows that the method usually employed in manufacturing hard
ferrites consists of steps well known in powder metallurgy. Compaction under the
464 H. STJ~d3LEIN

action of a magnetic field is the only technological suitable and useful variant for
magnetic powder which up to recently has frequently been improved process-wise.
The manufacturing operation must be seen as the sum of a number of
interdependent separate steps. Any change at one point affects the subsequent
steps. The following provides details of the individual steps in manufacture and
their interdependence.

2.I.1. Raw materials; main components and additives


According to the molar formula MO.6Fe203 the main constituents are the oxides
of iron, barium and strontium. It was found at an early stage that extreme purity
is not required in production and is even undesirable if optimum magnetic values
are to be achieved. This is one of the reasons why iron oxide, for instance, made
in different ways and from different sources can, in principle, be used. However,
undefined variations in the starting materials must be avoided owing to the
interdependence of the individual steps (fig. 19). In production it is therefore of
major importance that the physical and chemical parameters of the starting
materials remain constant. Other important factors are a minimum price, storage
capability and ease of handling.
For these reasons synthetic and natural iron oxides of the a-Fe203 type
(hematite) have mainly proved successful for production in long runs. Some
pertinent data are compiled in table 3. With these data it must be borne in mind
that purity, grain size, size distribution and shape, apparent powder and tap
density of the iron oxide etc. are very much dependent on the manufacturing
conditions (Gallagher et al. 1973, Stephens 1959, Balek 1970, Gadalla et al. 1973)
and are controlled as much as possible by the manufacturers so that reactivity,
apparent density and the content of impurities, for instance, are matched to
production requirements (Erzberger 1975).
Natural iron oxides have long been used in the manufacture of hard ferrites.
Depending on source, the ores contain varying amounts of impurities ("gangue"),
especially SiO2, which must be reduced to admissible values of less than 1% by
weight. Further attendant oxides may be A1203, TiO2 etc. The particles of these
iron oxides are generally angular with smooth cleavage surfaces, cf. fig. 20, and
only become rounder after a long period of milling, cf. fig. 21. In this way the
reactivity increases as a result of the, at first relatively large, oxide particles being
reduced in size.
Some of the iron oxides used are obtained by spray roasting HC1 pickling
solutions from steel plants according to the formula (Eisenhuth 1968):

4FeCI2 + 4H20 + 02 ~ 2Fe203 + 8HC1.

If the reaction is not fully completed some tenths percent by weight of chlorine
usually remain in the iron oxide which can affect storage and processing owing to
corrosion and its impact on the environment and must therefore be allowed for.
The chlorine content can generally be reduced to below 0.1% by washing with
water. The other impurities obviously depend on the type of steel pickled. Typical
HARD FERRITES AND PLASTOFERRITES 465

,.,d

:Z
'K
e. (2

O
6

e'~ H"~ I
v..a
© oq.
%

q3
oq. oo

cq..
c5

t-z. cq..
06 oo
,7 ¢¢)

e..,

.!

I
tt3
"7

e-,

e-,
c'q oq.
I
cA
G~
o0
i© G~

O
o9
,o O ""
e~ '0<
©

L)

"~.=:
Z r,,t3
466 H. ST~/d3LEIN

Fig. 20. Scanning electron micrograph of natural c~-iron oxide (hematite) as supplied.

amounts are some tenths percent by weight of manganese oxide, for example,
which i s of no significance in the manufacture of hard ferrites. A r o u n d 500°C
roasted oxides from Ruthner process occur as 20-400 ~ m thick hollow spheres of
relatively low apparent powder density, cf. fig. 22, which can be increased
considerably by mechanical treatment so that the oxide is easier to process. The
particles are not compact and very fine (about 0.1 i~m) and spherical, cf. fig. 23.
Owing to their reactivity and low price these oxides have frequently been used for
some years in manufacturing both hard and soft magnetic ferrites (Ruthner et al.
1970, Hiraga 1970, Ito et al. 1974, Ruthner 1980).
In contrast, HC1 regeneration by the Lurgi process taking place at 850°C in a
fluidized bed yields relatively coarse particles with a diameter range of some tenth
of a m m up to some mm. The apparent density of ca. 3 g/cm 3 is relatively high.
These oxides can also be used for hard ferrite production if milled sufficiently to
increase reactivity for the hexaferrite formation process. An advantage c o m p a r e d
with Ruthner oxide is their low C1 content in the order of a few hundredths %.
Pickling with sulphuric acid has lost ground over pickling with hydrochloric acid
since the recycling of these solutions created considerable pollution problems. The
iron oxides obtained in this way therefore need not be discussed at length.
HARD FERRITES AND PLASTOFERRITES 467

Fig. 21. Hematite of fig. 20 after 16 h of wet milling.

It is not u n c o m m o n to use mixtures of different iron oxides in production owing


to their different physical and chemical properties.
Two other processes mentioned in literature for manufacturing Fe203 are worth
noting, namely from pyrite according to the formula (Otsuka et al. 1973):

2FeS2 + 1102 ~ Fe203 + 4SO2,

and the oxidation of carbonyl iron ( O k a m u r a et al. 1952, 1955). Fagherazzi (1976)
provides a review of different processes and also reports on iron oxide produced
in the beneficiation of ilmenite (FeTiO3). Van den Broek (1974, 1977) reports on
experience gained in manufacture using various iron oxides.
Carbonates generally serve as a source of B a O and SrO because they are
chemically stable, inexpensive to store and only have to separate CO2 during
decomposition (Jfiger 1976, 1978, Ullmann 1974). Commercial materials may
further contain impurities such as SiO2, C a O and other compounds. Moreover,
barium carbonate generally contains some strontium carbonate and vice versa. In
principle, other salts, e.g., the alkaline earth nitrates or chlorides, could be used
but the separated gases are corrosive and present a health hazard. Experiments
468 H. ST~BLEIN

Fig. 22. Scanning electron micrograph showing hollow spherical agglomeration of ferric oxide particles
made by spray roasting (courtesy M.J. Ruthner of Ruthner AG, A-1121 Wien).

by Granovskii et al. (1970) with barium acetate, formate, nitrate, hydroxide and
peroxide showed no advantages c o m p a r e d with barium carbonate as raw material.
In manufacturing hexaferrites an excess of BaO, SrO or P b O above the
stoichiometric molar ratio of 1:6 is normally used, e.g., about 10tool% with
barium hexaferrites. This corresponds to a mixture of about 97% by weight of
BaO.6Fe203 + 3% by weight of BaO-Fe203. Such a composition is considerably
easier to sinter than t h e stoichiometric one. However, it must be seen in
connection with the above-mentioned additives which in both phases can b e c o m e
enriched to varying degrees or even form new phases, thus altering the above
ratio, cf. section 2.1.6. Therefore the weighed amount and the additives must be
matched to optimize the manufacturing process.

2.1.2. Mixing; granulation


The steps described below serve to prepare the raw materials so that the reaction
to form hexaferrite can take place in subsequent annealing treatment. As this is a
solid-state reaction there have to be sufficient short diffusion paths for the
reactants. T o achieve this the raw materials must be fine enough (the finer they
have to be, the smaller the proportion of the substance in the total mass is), well
HARD FERR1TES AND PLASTOFERRITES 469

Fig. 23. Scanning electron micrograph of spray roasted oxide particles (courtesy M.J. Ruthner of
Ruthner AG, A-1121 Wien).

mixed, compacted and in contact with one another. If the raw materials are not
fine enough, they can be ground and mixed in the same step. The good mix thus
achieved must be maintained until the reaction occurs, i.e., on the way from the
mixer to the furnace there must be no segregation resulting in local changes in
concentration when handling, charging and shaking the mixture. The equipment
and processes used for this purpose vary and therefore only general aspects are
discussed in the following.
Mixing ('homogenizing') can take place either with a wet or dry process (Ries
1969a). In wet mixing (and grinding) generally using an aqueous suspension
(slurry), vibration, drum or agitator mills (attritors) are used. This mixing method
is extremely effective but requires energy for dewatering and drying (Sch6ps et al.
1976). For this purpose, the suspension can be dewatered either mechanically, e.g.,
in a filter press, and then dried or fed directly to an atomizer.
T h e r e are two possibilities for dry mixing: (1) grinding and mixing, in drum or
ball mills, (2) intensive mixing in an edge-runner mill or in a high-intensity
counterflow mixer with swirler. The first method is used when the raw materials
are not fine enough for the subsequent reaction. Using the second method (fig. 24)
depends on whether the material is to be fed direct to the reaction furnace or as a
granulate. In an edge-runner mill the grinding wheels rotate in a pan in such a
way that there is relative motion between tread and pan bottom. The material is
subjected to friction, comminution, mixing and kneading and is also compacted
470 H. ST~?~BLEIN

and agglomerated to a certain extent. The intensive mixer can only operate with
dry material or convert the powder into a granulate through the addition of
moisture or a binding agent; see further below.
In a continuous or tunnel kiln the material can be m o v e d through the reaction
chamber in boxes or on trays, in bulk or in tablet form. Through the rotary kiln,
however, the material must m o v e without any boxes or trays, in which case, owing
to the fineness of the particles, no uniform and loss-free passage is guaranteed.
The mixture is generally turned into a granulate of pellets several m m in diameter
(Ries 1970, 1971b, 1975a, b). One way of doing this is to use granulating pans having
an inclined, rotating drum. The granulate is formed by spraying liquid, e.g. about
10%, onto the dry mixture so that a liquid film forms on the particles. The particles
granulate when the material on top tumbles down like an avalanche, see fig. 25.

Fig. 24. Diagram of a feeding, mixing and granulating plant for ferrite mixtures with raw material bins
(1), feeding and weighing station (2), edge-runner mill (3), vane feeder (4), belt weigher (5), granulator
(6) and belt conveyor (7) (Ries 1969b, 1971a).
HARD FERRITES AND PLASTOFERRITES 471

Usually an organic binding agent is added to the water to make the granulate less
sensitive to abrasion and disintegration. This must, of course, be expellable during
annealing with a minimum of residue. Certain cellulose products, alcohols, waxes
and alginates are proven binding agents. As a result of granulation the relative
density of the mixture increases appreciably, e.g., to twice the apparent powder
density, as the following example shows. This is a mixture of 47.5% by weight of
natural iron oxide (ce-FezO3), 31.7% spray-roasted iron oxide from HC1 recycling,
17.8% BaCO3 and 3.0% PbO:

Apparent powder density of mixture 1.33 g/cm 3 ~ 27% rel. density,


Tap density of mixture 1.78 g/cm 3 ~ 36% rel. density.
Density of a green pellet, 2.64 g/cm 3 ~ 53% rel. density,
Density of a pellet sintered at 1180°C 3.51 g/cm 3 -~ 66% rel. density.

The high compaction of the particles in the pellets promotes diffusion and
reaction during annealing. A particularly smooth throughput is achieved with a
granulate having a narrow particle-size range, this being produced by screening
and separating the pellets which are too large or too small. Mixing and granulat-
ing can also be carried out in a single unit which first operates as a dry mixer and
then, after the addition of a liquid, as a granulator. The processes and equipment

Fig. 25. Discharge area of a pelletizing pan with granulate (courtesy of Maschinenfabrik Gustav
Eirich, D-6969 Hardheim i. Odenw.).
472 H. STJd3LEIN

for mixing and granulating, with particular reference to the ferrite industry,
having been described in detail by Ries (1959, 1963, 1966, 1969b, 1971a, 1973).
Schinkmann (1960) has also referred to the importance of thorough blending
and shown what damage occurs as a result of local irregularities in the reaction:
warpings, distortions, cracks, pores, formation of vitreous reaction mass and
occurrence of fine, white, acicular precipitates. Discontinuous crystal growth is
also promoted (Arendt 1973a), cf. section 2.1.6.

2.1.3. Reaction sintering ; intermediate products


This production step serves to convert the raw materials into the hexaferrite phase
in the form of sufficiently large crystals. Using suitable manufacturing processes
even a densely sintered compact can be produced which merely needs to be
ground and magnetized before use. This special method for producing magnets is
described in detail in section 2.2.1. Normally, however, the reaction product
should not be of such density so that less grinding work is required in subsequent
crushing and milling. On the other hand, it is generally desired that the reaction
takes plac e to as large a degree as possible. The reaction has been investigated by
numerous workers. One characteristic of all hexaferrites is that these compounds
are not obtained from the raw materials in one single step but that intermediate
products ('precursor phases') with a more simple structure are formed first which
subsequently react to form the hexaferrite phase. The type of intermediate
product mainly depends on whether barium, strontium or lead hexaferrite is
being formed. This is a consequence of the different phase diagrams (cf. section
1.3).
The total reaction C to form the barium hexaferrite phase takes place according
to Suchet (1956) in the two unit steps A and B:

A: BaCO3 + Fe203--~ BaO-Fe203 + C02


B: BaO.Fe203 + 5Fe203~ BaO-6Fe203

C: BaCO3 + 6Fe203~ BaO.6Fe203 + CO2.

The intermediate product which occurs is the monoferrite BaO'Fe203. A number


of later investigations has directly confirmed this reaction sequence (Winkler
1965, Beretka 1968, Beretka et al. 1968, St~iblein et al. 1969, Bye et al. 1971,
Wullkopf 1972, 1973, 1974, Haberey et al. 1973b, K6nig 1974, Gadalla et al. 1975,
Efremov et al. 1977), but others only indirectly because although the authors
recorded comparable results they gave a different interpretation to them at the
time (Sadler et al. 1964, Sadler 1965). Some authors do not exclude the possibility
that intermediate products even richer in BaO than monoferrite may have formed
in an early stage of the reaction without this having been proved beyond doubt. If
BaCO3/Fe203 diffusion couples are used, layers of intermediate products richer in
BaO are in fact found in addition to mono- and hexaferrite (Wilson et al. 1972).
However, it is generally found that the BaCO3 decrease equals the BaO.Fe203
increase, as unit step A requires, and so, at the most, small amounts of such
HARD FERRITES AND PLASTOFERRITES 473

phases richer in B a O could be present. On the other hand, Gadalla et al. (1975)
report that, in addition to the monoferrite, 2BaO.3Fe203 occurs as an inter-
mediate product up to a maximum of 1200°C. In this connection it should be
pointed out that the monoferrite was the first intermediate product to be f o u n d
both in the synthesis of other hexagonal ferrites of the types Y, Z and U and also
in the synthesis of 2BaO-Fe203 (Winkler 1965), i.e., in feeds corresponding to very
different BaO/Fe203 molar ratios. The fact that the 1:1 molar compound
BaO-Fe203 was always the intermediate phase to be found first in appreciable
amounts, both in feeds with an excess of B a O and Fe203, points to the ease with
which this compound can be formed. There are, however, no studies on the
cause(s) of this. It is noteworthy, however, that no intermediate monoferrite was
found when a co-precipitated raw mixture was reacted to produce hexaferrite
(Roos 1979, 1980), see section 2.2.2.
Information varies on the temperatures at which reactions A and B occur,
depending on the prevailing conditions and state of the raw materials. Important
factors are the provenance, manufacture, particle-shape, -size and -range, powder
surface area, impurities, apparent density of the mixture, molar ratio Fe203/BaO
etc. However, there is obviously no clear connection between the above factors
and the reaction kinetics and in particular the reaction temperature. Furthermore,
both reactions are affected in varying ways by these factors (Wullkopf 1974).
Naturally the well-known laws applying to solid-state reactions apply here, too,
especially as, apart from the reaction time the reaction temperature also plays a
certain role.
The above-mentioned studies on reactions A and B provide the following
information on temperatures:
The first, though very small, amounts of CO2 are released from about 400°C
(start of reaction A). Decomposition of barium carbonate thus begins at a
temperature around 350°C lower than when pure barium carbonate is heated
(Bulzan et al. 1976). Radiographically detectable quantities of monoferrite are
generally found only at about 600 to 650°C upwards, with very fine raw material
particles at 100 to 200°C lower, and by 950°C all the barium carbonate has been
converted and reaction A finished. The appearance of hexaferrite at 700 to 800°C
marks the beginning of reaction B which ends with the disappearance of the iron
oxide. Figures vary considerably on this point, ranging from 900 to 1200°C. This is
plausible as the complete reaction is especially dependent on intimate mixing of
the reactants and is affected particularly by the upper end of the FezO3 particle-
size range (Bye et al. 1972). A comparison of the figures shows that the
temperature ranges of both reactions overlap to a certain extent.
The mechanism of reaction step B was studied in detail using diffusion couples
of BaO'Fe203 and Fe203 pellets (Stfiblein et al. 1972, 1973a, 1973b). It was shown
that the hexaferrite crystals in the reaction layers grow in a preferred direction.
Layers with different orientation can be clearly distinguished, see fig. 26. A 0.1 to
0.2 mm thick reaction layer of hexaferrite crystals, whose basal plane is preferably
parallel to the contact surface, i.e., which have a {0 0. l} fibre texture (stage I)
forms on the contact surface of both reactants. The {0 0. l} texture is all the more
474 H. S T ~ B L E I N

Z,-
~c~
3-.9
o

t.
2
/
j.
c I--3,
o~

0.2 0.~ mm
Reaction
interface

BaO. Fe203 q ~. Fe203


15
arb.
units

I0
f • 3 (00.8)

H.O]
/
,

7,
C

\
o0'2 0.2 0.,~ tom
Reac ion
interface
Distance

Fig. 26. X-ray intensity ratios and intensity values across the thickness of the hexaferrite layer in the
couple Fe2Os-BaO.Fe203.
HARD FERRITES AND PLASTOFERRITES 475

p r o n o u n c e d the smoother the contact surface is. The texture forms regardless of
whether the Fe203 contact surface has a random polycrystalline structure or
consists of a single Fe203 crystal cut parallel or perpendicular to the basal plane.
A topotactic mechanism can therefore be excluded as the cause of the {0 0. l}
texture.
This {0 0. l} orientation is retained as the thickness of the layer increases. With
the increase in layer thickness the orientation becomes weaker on the side facing
the BaO.Fe203 whereas the {0 0. l} texture facing the Fe;O3 side fades away. At
the same time a {h k. 0} texture forms where the basal planes of the hexaferrite
crystals lie perpendicular to the contact surface (stage II).
The explanation for this may be found in the anisotropic growth rate of the
hexaferrite crystals which is relatively low along the c-axis but relatively high
perpendicular to it. It is assumed that in stage I a discrete random nucleation of
the hexaferrite first occurs on the Fe203 contact surface and that the incubation
period and growth rate are not dependent on orientation. After a certain growth
the state shown diagrammatically in fig. 27 is reached where only the two possible
extreme cases of orientation are illustrated. After the reaction of the uppermost
iron oxide layer, i.e., at the end of stage I, the {0 0. l} crystallites constitute most
of the surface area. In stage II the reaction can then only proceed perpendicular
to the phase boundary. In this case the {h k. 0} orientated crystals have the best
opportunity of growing. Therefore in both stages the direction of maximum
growth of the hexaferrite crystals is parallel to the direction in which the reaction
rate is highest.
This explanation contradicts a paper by Takada et al. (1970b) in which a
topotactic mechanism is presumed. These authors found that hexaferrite platelets
had grown with their basal planes parallel to small platelet-shaped Fe:O3 crystals.
This result corresponds exactly to the above-mentioned stage I of the diffusion
experiments and can therefore be explained by the anisotropic growth rate of the
hexaferrite. The latter explanation is also given by Kohatsu et al. (1968) from
analogous diffusion experiments with CaO-6A1203, a material crystallographically
similar to hexaferrite, where a topotactic mechanism could definitely be excluded.

SaO'Fe203

Nuclei Reaction interface

aO6Fe25
(differently oriented)

Fig. 27. First stage reaction model for the formation of hexaferrite from monoferrite and hematite.
476 H. STPd3LEIN

A still atmosphere impairs ferrite formation after reaction C because the CO2
obtaining first impedes further decomposition (Mondin 1969, Bye et al. 1971).
Therefore, it must always be ensured that there is an adequate supply of oxygen
during the reaction. According to Heimke (1966) a certain amount of H20 has a
positive effect on the reaction.
Beretka (1968) found no change in the reaction sequence as in formulas A and
B by adding 0.5% NaF, but a 150 to 200°C reduction in the formation tem-
peratures. Bye et al. (1971) achieved similar results by adding 0.5% LiF which
accelerates carbonate decomposition, the reaction as in formula B and the grain
growth and causes the formation of a liquid phase. However, it was later
established that the lithium ferrites LiFeO2 and LiFesO8 were the first products to
form in the reaction (Wilson et al. 1972).
According to Haberey et al. (1973b) monoferrite formation takes place endo-
thermally and hexaferrite formation without any heat change. Bye et al. (1971)
found an activation energy of 209 kJ/mol (50 kcal/mol) for the step determining
the rate of carbonate decomposition. Sadler (1965) gives a similar value of
190.5kJ/mol (45.5kcal/mol) for reactions at temperatures below 735°C and a
value of 306.5 kJ/mol (73.2 kcal/mol) for temperatures above 735°C. It was found
that the reaction can be satisfactorily expressed by the formula derived by Jander
(1927): [ 1 - ( 1 - p)i/312 = kt, expressing the relation between reacted portion p of
ball-shaped particles, reaction rate constant k and reaction period t. According to
Kojima et al. (1969) as well, Jander's formula is the one best suited to describe the
reaction rate. However, between 850 and 900°C activation energies of 201 to
904 kJ/mol (48 to 216 kcal/mol) were found, depending on the type and treatment
of the iron oxide used in the reaction. A very low value of 59-+42k J/tool
(14 _+10 kcal/mol) for the subsequent reaction stage was given by Cho et al. (1975a).
Literature data on the reaction mechanism and kinetics were compiled by Schrps
(1979).
The formation of the strontium hexaferrite phase was investigated by Beretka et
al. (1971) and by Haberey et al. (1976). Beretka et al. (1971) describe the two unit
steps D and E by the following formulas:

D: S r C O 3 + Fe203--~ (SrO'Fe203 + 2SrO.Fe203) + CO2,

E: (...) + 5Fe203-~ SrO.6Fe203,

where, however, only SrO-Fe203 is claimed to appear after reaction in a vacuum.


Apart from the fact that owing to the different molar ratios on both sides the
equations can, at best, only describe the reaction qualitatively, there is consider-
able doubt about the existence of strontium monoferrite (see section 1.3). On the
other hand, Haberey et al. (1976) and Vogel et al. (1979b) found the unit steps F and
G:

F: SrCO3 + ~Fe203 + (0.5 - x)½02-~ SrFeO3-x + CO2,


G: SrFeO3_x+ 5.5Fe203-~ SrO-6Fe203 + (0.5 - x).102,
HARD FERRITESAND PLASTOFERRITES 477

where the intermediate product strontium perovskite SrFeO3_x only occurs at the
reaction temperature whereas 7SrO-5Fe203 (4SrO.3Fe203) was found after
quenching. It may be that the 02 supply recorded in reaction F does not come
from the atmosphere but from carbonate decomposition, i.e. that instead of the
recorded CO2 a corresponding mixture of CO2 and CO is given off (Haberey et al.
1977a, cf. also Wullkopf 1978). The results of Beretka et al. (1971) can, at least
partly, be interpreted by the explanation of Haberey et al. (1976).
As far as the reaction temperatures are concerned, the picture is as follows: The
initial traces of CO2 are found from 300°C upwards, once again appreciably lower
than with the decomposition of pure carbonate (from about 650°C upwards).
Considerable amounts of the intermediate product occur at 600 to 660°C and
SrCO3 can be detected radiographically only below 800°C (end of reactions D and
F respectively). The hexaferrite phase occurs from 800°C upwards and iron oxide
up to about 11500C (start and end of reactions E and G). Owing to the small
number of test results the temperature values should not be regarded as i00%
accurate meaning that they correspond roughly to the reaction temperatures of
barium hexaferrite. According to Haberey et al. (1976) reaction F is strongly
endothermal, reaction G weakly endothermal. This corresponds qualitatively to
the heat changes dffring the formation of barium compounds.
Another reaction sequence than D-E and F-G was reported for co-precipitated
ferric hydroxide and strontium laurate, see section 2.2.2 (Qian et al. 1981).
The formation of lead hexaferrite has also been investigated by several authors
(Berger et al. 1957, Mountvala et al. 1962, Bowman et al. 1969, Mahdy et al. 1976b)
for which the reaction steps H, I and K are given:

H: 2PbO + Fe203--~ 2PbO.Fe203,


I: 2PbO.Fe203 + 3Fe203 ~ 2[PbO-2Fe203],
K: 2[PbO.2Fe203] + 8Fe203~ 2[PbO-6Fe203].

However, both intermediate products are not always found during the reaction,
for kinetic reasons according to Bowman et al. (1969). Some partial reactions can
obviously proceed very slowly, see section 1.3.5. The following temperatures must
not therefore be considered as homogeneity ranges for the state of equilibrium,
but must be seen in the dynamic sense, i.e., longer periods displace the tem-
perature ranges towards lower values and they depend on the reactivity of the raw
materials: PbO is present up to a maximum temperature of 750°C, 2PbO.Fe203
was found between 670 and 850°C and PbO-2Fe203 between 600 and 825°C;
PbO.6Fe203 can occur from 750°C upwards and Fe203 up to a maximum of
1000°C. The relatively high vapour pressure of PbO results in the renewed
occurrence (precipitation) of Fe203 from 1150°C upwards (Bowman et al. 1969).
According to Berger et al. (1957) appreciable losses in weight can occur from
950°C upwards owing to the evaporation of PbO.
Reaction sintering in industrial plants mainly takes place nowadays in internally-
fired rotary kilns with ceramic lining where temperatures of about 1200-1350°C
478 H. ST.~3LEIN

necessary for the barium and strontium hexaferrite reaction can be attained (Petzi
1974b). Cartoceti et al. (1976) have investigated the heat balance of such a kiln
charged with a wet mix and found that less than 10% of the total combustion heat
is utilized for the hexaferrite reaction. The balance is somewhat more favourable
when a dry mixture is used. It must be fed as a granulate to permit uniform
throughput (cf. section 2.1.2) and first passes through the reaction zone, undergo-
ing no abrasion if possible, and then through the cooling pipe below the kiln shell,
cf. fig. 28 (Petzi 1971). Owing to direct firing b y g a s or oil burners the oxygen
partial pressure varies in the kiln chamber but is always lower than in the
atmosphere and especially low in the burner zone, a s is shown in fig. 29 (Petzi
1971). Therefore direct contact between flame and material must be avoided for a
perfect reaction. Temperature and annealing time depend on the reactivity and
particle size of the raw materials used and also on the desired technological
properties of the powder and on the magnetic properties of the magnet grade to
be manufactured. Temperatures of 1000 to 1100°C are, under certain circum-
stances, sufficient for isotropic hard ferrites as unreacted constituents can form
hexaferrite during the final sintering (cf. section 2.1.6) and a small crystallite size
in hexaferrite is admissible and generally desirable and necessary. Anisotropic
hard ferrites require higher reaction temperatures of the mix because the mini-
mum size of all hexaferrite crystallites must be around 1 Ixm. Fine grinding down

Fig. 28. Rotary kiln for calcining hexaferrite obtained from raw materials showing below the cooling
pipe for processed material (courtesy of Fa. Riedhammer, D-8500 Nfirnberg).
HARD FERRITES AND PLASTOFERRITES 479

db temperature-O2-curve
21% 02
sintering_ 1400° c ..- ,...~"'"'"'~' .......................... "-,..-.. 20
ternflerqture1200 18 furnace atmosp.here
"""" T "''... -15
1000 furnace ",,... -14
800 \ ' cooh'ng tube ,/r'~'~ /
12
10
600 ~ M , ' / , O , j . J ~ ' " 86
coo
2oo ~. / ..,..,.,,i- . . . . . . . -2

furnace length 0 1 2 3 ,~ 5 6rn

__Pyr~2_~_-~ charge__
temperature-
me(~surmg_ vibrator
burner
(gas-or oilheated)

Fig. 29. Temperature and atmosphere along the length of the rotary kiln of fig. 28. The atmosphere
was measured in the lower half near the pellets (M1)and in the upper half (M2),where the 02 content
is very low within the burner zone (Petzi 1971).
to a size below 1 }xm (cf. section 2.1.4) produces particles which consist of one
single crystallite and which can therefore be aligned in a magnetic field (cf. section
2.1.5). Accordirlg to Van den Broek (1974, 1977) a reaction temperature matched
precisely to the raw materials used is therefore of major importance because
deviations from it cannot be entirely corrected in subsequent manufacturing. T o
produce plastoferrites (cf. section 4.1) the same considerations as with compact
ferrites apply regarding the particle size to be produced, i.e., it depends on
whether alignable powder particles are needed or not. Moreover, for magnetic
and practical reasons, the powder should only consist of the hexaferrite phase.
Of course, other kiln types can be used for reaction sintering, such as bogie,
end-discharge pusher or batch furnaces as long as the temperature, reaction time
and atmosphere requirements are met. In this case the powder is used, for
example, in bulk or briquet form. Owing to the poor thermal conductivity of
powders the innermost parts must be given sufficient time to react and recrystal-
lize.
Fagherazzi et al. (1972) have described a pot-grate kiln which permits exact
temperature control with little crystal growth and caking. Iron hydroxide with
acicular particles is said to be especially good as raw material. The reacted
pellets are friable giving platelet-shaped single-domain crystals with coercivities
HHc of up to 340 kA/m (4.26 kOe) with BaFe12019 and 455 kA/m (5.7 kOe) with
SrFe12019, without tempering being necessary (cf. section 2.1.4).
A recent further development of the rotary kiln type shown in fig. 28 is called
the passage pendulum kiln, see fig. 30. The reaction tube no longer rotates
continuously in the same sense, but oscillates around an equilibrium position. As
a consequence there are constructional advantages with the energy supply,
enabling a more efficient and compact setup.
480 H. ST,3d3LEIN

Fig. 30. Passage pendulum kiln for calcining of hexaferrites from raw materials at max. temperatures
of 1350 to 1400°C, with adjustable temperature curve and controllable energy supply. Length 6 to 10 m
(courtesy of Fa. Riedhammer, D-8500 Nfrnberg).

2.1.4. Preparation of mouldable powder


The reaction-sintered mass is hard and often coarse-sized and therefore has to be
turned into a mouldable and sinterable powder with crystallite and particle size
fulfilling certain magneto-physical requirements. These requirements stem firstly
from the fact that crystallite size is linked with coercivity. The critical size for
single-domain behaviour is around 1 p~m. There is always grain growth and
recrystallization during sintering and so before sintering the crystallite size has to
be considerably smaller than 1 fxm, e.g., with most crystallites in the size range 0.1
to 0.5 txm. The coercivity of the magnetic material can therefore be influenced by
the intensity and duration of milling. Secondly, the requirements depend on
whether isotropic or anisotropic magnets are to be manufactured. For manufac-
turing isotropic magnets the powder particles can be polycrystalline, while for
manufacturing anisotropic magnets preferably all of them have to be monocrys-
talline so that they can be aligned in the magnetic field (see section 2.1.5). As
torque in a magnetic field is proportional to volume, for good alignment the
particles should not be unnecessarily fine; this also facilitates the escape of air or
milling fluid during pressing and reduces shrinkage during sintering.
The reaction-sintered lumps are crushed and ground to produce the powder.
Using jaw or roll crushers a granulate in particle sizes of one or several
HARD FERRITES AND PLASTOFERRITES 481

millimetres is produced which in the dry state can be reduced in ball or vibration
mills, for instance, to particle sizes of, at the most, 100 or several 100 ixm. In
certain circumstances this can suffice for the manufacture of isotropic magnets
from polycrystalline particles. Particle sizes under 1 Ixm are obtained by batchwise
wet milling, e.g., in roller or vibration mills, or continuously in attritors, for
instance. Figures 31 to 33 show some common types. During milling additions can
be introduced if the feed is to be corrected. S o m e economic aspects of various
types of mill were examined by Maurer et al. (1966) and technical aspects
described by John (1973).
The advantage of the attritor and the vibration mill over roller mills is seen in
the intensive grinding action which gives relatively short milling times (Heimke
1962, Richter 1968, Stanley et al. 1974) and is said to result in relatively little
abrasion (Maurer et al. 1966). As grinding media steel balls of uniform size are
normally used as long as differential wear during operation does not cause certain
size variations. According to Kal3ner (1970) the use of grinding balls with different
diameters gives no advantages. H e also found that the addition of interfacially
active agents to the milling fluid showed no effects on the result, but failed to

Fig. 31. Ball mill (Fa. Dorst Keramikmaschinenbau, D-8113 Kochel am See) having a capacity of 5.3 m 3
(courtesy of Fried. K r u p p G m b H , Krupp W l D I A , D-4300 Essen).
482 H. STABLEIN

Fig. 32. Rotary vibrating mill with a total capacity of 12l (courtesy of Fa. Siebtechnik GmbH, D-4330
Mfilheim (Ruhr) 1).

mention the type of agents tested. The experiments carried out by Tul'chinskii et
al. (1971) are, however, claimed to show that the milling time can be reduced to a
third without affecting the ultimate magnetic values if ammonium carbonate is
added to the grinding water.
After fine milling the water content has to be adjusted or the water removed
depending on subsequent treatment which is in turn dependent on the magnetic
quality to be obtained, cf. section 2.1, fig. 19. For wet compaction water contents
of 20 to 50% by weight are required, which can be obtained by decanting, partial
evaporation or adding water. These water contents roughly correspond to 43 to
15% by volume of ferrite. The high ferrite contents cause the suspension's
consistency to be paste-like, the low ones cause it to be fluid. By way of
comparison, a packing density of about 15% by volume of ferrite is obtained
when a milled suspension settles or the dry powder is poured. For optimal
alignment in a magnetic field the packing density has to be sufficiently low.
If the powder is to be subsequently treated in the dry state the water is either
removed by means of spray driers, whereby the particles can form fine hollow
spheres, e.g., a few 0.1 mm in diameter of correspondingly low apparent powder
density (approx. 1.0-1.2g/cm3), or by filter pressing, evaporation, etc. In the
drying process rather constant particle aggregates with poor magnetic field
alignability are formed which can be broken up in suitable mills (pinned disc mill,
jet mill, etc.) to obtain a powder with good alignability (Kools 1978b) which,
HARD FERRITES AND PLASTOFERRITES 483

Fig. 33. Attritor mill for hard ferrites (courtesy of Fa. Netzsch GmbH, D-8672 Selb).

however, due to its fineness does not flow well and is m o r e difficult to handle
during pressing. If, on the other hand, isotropic hard ferrite is to be manufactured,
the particles are granulated either after drying with the addition of binders and
means to facilitate compaction (cf. section 2.1.2), e.g., by a vibrating screen that
sizes the material at the same time, or by an atomizer with additions of
dispersants, binders and lubricants. Dispersants lower the viscosity and the
required minimum content of water already in amounts of a few hundredths to a
few tenths of a percent. G u m arabic, a m m o n i u m citrate and polyethylenimine was
successfully tried by Vogel (1979) in soft ferrite suspensions and found to be
compatible with polyvinyl alcohol. Binders must give a certain strength to the
484 H. STJd~LEIN

granules to enable good flowability. Granulate with aggregate sizes of smaller


than about 1 mm is easy to handle. The effect of polyvinyl alcohol (0.75 to 3%)
and of polyamine sulfone (1.5%) as binders in soft ferrite powders were described
by Harvey et al. (1980). The drying kinetics of single droplets of ferrite suspen-
sions were investigated by Malakhovskij (1980).
The phenomena deriving from the interaction of milling material, grinding
media and grinding fluid are examined more closely in the following. Reducing
the particle size is accompanied by two undesired effects: the grinding media
undergo abrasion and the material and water react, with earth alkaline hydroxide
being formed and part of the ferrite destroyed.
Figure 34 shows different particle size ranges for powder milled with a feed size
of <1 mm in a ball mill (Bungardt et al. 1968). Logarithmically the particle sizes
largely reflect a normal distribution. Richter (1968) obtained the same results,
establishing a closer reflection of the Gaussian curve than of the Rosin-Rammler
distribution (1933). The deviations from the normal distribution in the fine and
coarse range were greater after milling in the ball mill than after milling in the
attritor. This is held as attributable to the different grinding mechanisms of the
mills: abrasion grinding preferable for the attritor, impact grinding and chipping
grinding preferable for the ball mill. Initially grinding progresses quickly but then
the curves show only slight progress as a point is reachedwhere as many particles
agglomerate as are ground. During ball milling the particle size variations

a:l

"4

"6

)(i 2 2 ~ 6 a101 2 z 6 8100 z ~ pm 810 ~


Particle d i a m e t e r
Fig. 34. Particle size distributions of BaO.5.6Fe20~ after ball milling at different milling times
(Bungardt et al. 1968).
HARD FERRITES AND PLASTOFERRITES 485

decrease somewhat. The authors found, however, that the variations are not
related to the milling conditions and in particular that it is the same after ball
milling as after attritor milling and is therefore related to the material. Neverthe-
less, the proportion of coarse or fine particles can be influenced by the milling
conditions (mill type, grinding media geometry, quantity of grinding balls, etc.).
The particle size distributions shown in fig. 34 were determined using an
electron microscope. Special preparation is required because they are single-
domain particles which cannot be demagnetized by magnetic fields and which
therefore always attract each other magnetically and conglomerate. Horn et al.
(1968) drew the particles magnetically onto a slightly sticky carrier, while Machin-
tosh et al. (1976) electrostatically charged the powder at a temperature over Curie
point and drew it onto a carrier. A g o o d dispersion was attained in both
experiments. Heidel et al. (1977), however, only found a partially satisfactory
dispersion of the powder in the magnetic field. In contrast, determining the size of
the particles via air separation or changing the resistance of an electrolyte
(Coulter counter) gives values which are too large by 1 to 2 powers of ten because
an agglomerate particle size is measured (Bungardt et al. 1968).
The problems encountered in determining the particle size of hexaferrite
powders and the resultant difficulties in describing them have been referred to in
particular by Bungardt et al. (1968) and Kools (1975). Hexaferrite particles
preferably have the shape of platelets because they split relatively easily parallel
to the basal phase of the hexagonal crystal lattice. They mostly lie with this plane
on the carrier (St/iblein 1957) and so the electron microscope captures the largest
surface of the particle. Although further splitting of the particles in the basal
plane enlarges the surface area it does not seem to reduce the particle size. This is
probably one of the reasons why the specific surface area grows noticeably with
the duration of milling, but as seen under the microscope the particle size hardly
decreases (Maurer et al. 1966, Richter 1968). A further reason for these deviations
is that Ba-hexaferrite particles may be surrounded by a barium carbonate/barium
hydroxide film, the formation of which will be dealt with later. Inner surfaces of
these porous layers can give the false impression that the specific powder surface
areas values are excessively high. Kools (1975) compared the particle sizes
calculated from measurements of the surface area or of the permeability to gas
(Fisher sub-sieve-sizer) with those determined with scanning microscopy. Parti-
cularly puzzling was the proportion of very fine particles, probably attributable
mainly to abrasion and foreign phases of the commercial hexaferrite magnets.
Indeed, the particle sizes determined by the various methods were more uniform
when the finest particles had been removed by acid treatment. Nevertheless, the
values obtained by the different methods only agreed qualitatively. The method of
determining the permeability to gas proved itself relatively insensitive to the
content of finest particles.
With a magnetic powder such as hexaferrite powder magnetic quantities can
also be used to examine the effect of milling, especially the coercivity of the
magnetic polarization 1He and the specific saturation polarization Js/p (=47r~r in
the CGS system). Figure 35 shows the she curve according to Heimke's experi-
486 H. ST.~?,LEIN

300
kA
m b"
250,~

200

150 b

J"c
,°° L , t . , ....____ a

50

i )

0 200 ~00 600rain


Millingtime
Fig. 35. Intrinsic coercivity]He vs. milling time of Ba-hexaferrite powder (Heimke 1962). Curve (a):
calcined 4h 1370°C+milled; curve (a'): as curve (a)+ 0.5 h 1000°C;curve (b): calcined 0.5 h 1160°C+
milled; curve (b'): as curve (b) + 0.5 h 1000°C.

ments (1962, 1963). As the milling time lengthens jHc drops, if a low calcined, i.e.,
fine crystalline, material is used. When a high calcined, i.e. coarse-crystalline,
material is milled, however, jHc rises at first. Under further milling jH~ then
drops, as numerous experiments have shown (Tenzer 1963, Bungardt et al. 1968,
Richter 1968, Haneda et al. 1974a, Mackintosh et al. 1976). In most experiments
the maximum jH~ value was between 103 and 143 kA/m [1.3-1.8 kOe], and in one
case (Mackintosh et al. 1976) it was 190 kA/m [2.4 kOe] and is thus evidently
dependent on the particular experimental conditions.
This behaviour is at first surprising. It is expected that as the material is ground
finer the number of multi-domain particles will decrease and that of single-domain
particles increase and as a result coercivity will rise. This is evidently the case for
the initially coarse material. The fall in jH~ must stem from another mechanism.
Heimke (1962, 1963) and most of the authors mentioned attribute it to lattice
defects which permit a relatively easy formation of remagnetization nuclei and
therefore of domain boundaries.
By heat treatment of the ground powder, e.g., ~h of annealing at 1000°C,
coercivity can be noticeably increased, both for the initially coarse and the fine
powder, cf. fig. 35. According to Heimke (1962, 1963) this is due to curing of the
lattice defects. The coercivity is then mainly determined by the crystallite size and
in addition by the material as was found by Richter (1968), which for barium or
strontium hexaferrite powder obtained maximum jH~ values of 240 and 290 kA/m
(3.0 and 3.6 kOe) respectively. Roughly ideal single-domain behaviour of ]/arc~-
600 kA/m (7.5 kOe) can be obtained after very long milling. Tanasoiu et al.
(1976b) obtained a strontium hexaferrite powder with fl-/c = 520 kA/m (6.5 kOe)
HARD FERRITES AND PLASTOFERRITES 487

after 1700 hours of milling and annealing at 1000°C. Similar values are obtained
when the crystallites are produced without plastic deformation, cf. sections 1.2
and 2.2.2. Annealing is of particular importance in the manufacture of powders
for plastoferrites, cf. section 4.1, but after Fahlenbrach (1965), it is unnecessary or
even harmful in the manufacture of compact hard ferrites. The influence of
annealing on the magnetic characteristics of barium hexaferrite powder was also
investigated by Sch6ps et al. (1977). Tenzer (1963, 1965) attributes the drop in
coercivity during milling to the formation of superparamagnetic crystallites and
the jHc rise during annealing to their being sintered and recrystallized.
If superparamagnetic crystallites were the cause of the low coercivity, etching
the powder in acid should produce a noticeable jHc rise owing to preferential
decomposition of the very fine particles. In fact, however, Richter (1968) found
that this only gave an increase of 5 to 15%, i.e., only a fraction of the rises of 50 to
300% attained through annealing for 0.5h at 900°C. Furthermore, X-ray
measurements have shown that the line broadening in milled powders is largely
caused by lattice distortions and to a lesser extent by ultrafine crystallites, and so
lattice defects are regarded as the main cause of the 1He drop. The same
conclusion is reached by Haneda et al. (1974a) on the basis of M6ssbauer
experiments and the angular dependence of the coercivity, by Hoselitz et al.
(1970) on the basis of remanent torque curves, Bottoni et al. (1972) on the basis of
investigations into rotational hysteresis, and by Ratnam et al. (1970) on the basis
of electron microscope studies and the angular dependence of coercivity. Accord-
ing to the latter stacking faults and deformation twins are the major lattice
defects.
Esper et al. (1978) have pointed out the significance of the annealing atmos-
phere in regenerating the coercivity. On tempering the milled powder in an N2
atmosphere the jHc rise is distinctly lower than annealing in air or does not occur
at all. In partial contrast to Richter's (1968) experiments, however, it was found
that jHc rises roughly in proportion to the percent by weight removed by etching and
that at around 70% by weight powder loss a value approximately the same as after
optimal annealing in air is attained. The authors interpret their results in a model in
which the powder particles exhibit two types of lattice defect. Outside an undisturbed
core covering ~ of the particle radius there are "mechanical" dislocations while the
outermost layer covering 10% of the radius is additionally disturbed by oxygen
vacancies.
The second magnetic quantity, specific saturation polarization Js/p, decreases as
milling time increases (Heimke 1964, Bottoni et al. 1972, Haneda et al. 1974,
Mackintosh et al. 1976, Giarda 1976), in extreme cases to almost ~ of the initial
value (Tanasoiu et al. 1976b). After annealing very different behaviour was
established: a further, though only slight drop, a partial or almost complete
restoration of the initial value. There are evidently a number of influences at play
here. Firstly, there is the reaction of the hexaferrite with the water; H a n e d a et al,
(1974a) were in fact able to prove that o~-Fe203 is released after sufficiently long
milling. If there is a surplus of alkaline earth present and this reacts with the iron
oxide released or with particles abraded from the steel balls there has to be a
488 H. STJid3LEIN

corresponding recovery back towards the initial value. The rise in the Fe 2+ content
with the time of milling is, however, not connected with the fall in saturation
polarization, but is due to abrasion of the grinding media (Heimke 1964). A drop
in saturation polarization was, however, also found after strong, dry deformation
of high-coercivity Ba-hexaferrite crystals (Ratnam et al. 1970) and during milling
of Sr-hexaferrite powder in alcohol (Tanasoiu et al. 1976a and b). In these cases,
too, the drop was completely or largely reversed by annealing. It is to be
concluded from this that the superexchange interaction between the five magnetic
sublattices of the hexaferrite is changed during deformation. Iwanow et al. (1966)
and Richter et al. (1968a) gave an alternative explanation to the effect that a new
soft-magnetic phase forms having a Curie point of about 150°C and disappearing
during annealing.
Apart from coercivity and saturation polarization, which can only be measured
in the dried powder, the hysteresis loop of the suspension can be used for directly
assessing the state of the particles (Kools 1975). From this loop the alignability of
the particles in the suspension, which is important to know for the manufacture of
oriented magnets, can be assessed from the ratio between remanence and
saturation polarization. Within several hours of milling the ratio at first rises
rapidly, this being due to the reduction of the crystal aggregates to single-crystal
particles; during continued milling it reaches a maximum value and then starts to
fall slowly.
As already mentioned, during milling in water a reaction takes place between
the barium hexaferrite and the milling fluid. Barium hydroxide is formed and as a
result the pH value of the water may rise to 13 or 14. It was formerly supposed
that this might be due to unreacted remnants from reaction sintering. This is
contradicted by results obtained by Stfiblein et al. (1969) to the effect that BaO
losses also occur after repeated milling and sintering and with sub- and supra-
stoichiometric compositions (Fe203/BaO<6 or >6). It was also found that
BaO.Fe203 is particularly easily decomposed, see also Belyanina et al. (1977).
After multiple milling and sintering a composition almost the same as
BaO.6Fe203 was obtained. Analog milling tests in ethyl alcohol or acetone
produced for both substances only a slight increase in the molar ratio, which can
be explained by the material being milled picking up abraded particles. The
relatively low resistance of the monoferrite to water can be see when boiling it in
water, where BaO losses of over 50% were determined. In contrast, Ba-hexafer-
rite showed itself resistant to boiling water.
The reaction between barium hexaferrite and water was also found by Bungardt
et al. (1968), and after Tanasoiu et al. (1976a) there is an analogous phenomenon
in the case of strontium hexaferrite. The hydroxide is the cause of a number of
undesired effects, some of which have already been mentioned; barium car-
bonate/barium hydroxide films form on the powder particles during drying, giving
the false impression that the specific surface area is too high and that adjacent
particles are agglomerating which causes poor magnetic field alignability and
hinders sintering (cf. section 2.1.6). If a dry powder of good alignability is required
disaggregation has to be carried out.
HARD FERRITES AND PLASTOFERRITES 489

2.1.5. Die pressing; orienting field


The treated powder is turned into a compact in two stages: (1) By applying
pressure at room temperature (pressing) and (2) by annealing at temperatures
above 1000°C at normal pressure (sintering; cf. section 2.1.6). Their relative shares
in the total compaction process can be altered by varying the pressure, thereby
producing different microstructures with the same final density. In principle, both
steps can be carried out at the same time by what is known as hot compaction, cf.
section 2.2.5, but technical difficulties prevent wide-spread application. In this
section we will therefore only consider those phenomena which take place when
hard ferrite powder is compacted at room temperature.
For pressing, powder (or powder suspension) is subjected to pressure in a
largely enclosed space. Normally a die is used which has a continuous opening
('die cavity') of (almost) constant cross-sectional area which is closed by movable
upper and lower punches, cf. fig. 36. This process is called die pressing because, in
addition to compaction, shaping to dimensions similar to those of the part to be
produced also takes place. For very long production runs the walls of the die
cavity are made of hard metal because of the abrasive action of the ferrite
particles. Another possibility is isostatic compaction where pressure is exerted
from all sides with the powder enclosed in a deformable container. Although a
higher degree of compaction can be attained, which can have a positive effect on
the strength of the compacts, the homogeneity of their density, their sintering
behaviour and magnetic characteristics after sintering (Kools 1973, Stfiblein 1973),
this process is too expensive for normal operating practice. Shaping is only
possible to a limited extent, e.g., for cylinders, or in post-compacting parts which
have already been die-pressed.
, The degree of compaction attainable depends on the pressure, the state of the
powder and on friction. Figure 37 shows the density of (almost) isotropic compacts
of barium hexaferrite after die pressing (<5 kbar) and after quasi-isostatic post-
compaction between 5 and 26 kbar (Stfiblein 1973). A relative density of about
85% was achieved with a maximum pressure of 26 kbar. This density could not be
increased by stress-relieving at 400-600°C, by renewed application of pressure or
by using pressing aids. The same density was also achieved when pressing
anisotropic compacts. The compaction behaviour can be expressed, for example,

a) b)
Upperpunch

(or

erpunch /
Fig. 36. Die pressing with relative movement of both punches (a), or of only the lower punch (b) with
respect to the die during densification.
490 H. ST~d3LEIN

by the formula

pp = 2.88 x pO.129, (1)

with pp = green density in g/cm 3 and p = pressure in kbar; another formula is


given in fig. 37. As the compact volume is V - 1/pp, eq. (1) can be re-written as

V x p0.129 = c o n s t , (2)

which has a certain similarity to the equation for an ideal gas with the exponent of
p, however, indicating that compaction is much more difficult c o m p a r e d with gas.
It should also be mentioned that SmCo5 powder was compacted to 82% at
p = 20 kbar (Buschow et al. 1968) and hexaferrite raw mix to 80% relative density
at p = 26 kbar (Stfiblein 1975), i.e., to similarly high values.

~.5
9/cm~

4O
J
3.5
./
~ experimental
t~
o I/op =0.35-0.091gp
• _,op =2.88.p °.7"~
30

2.50 ' , ,
I0 20 kbar 30
Pressing p r e s s u r e p
Fig. 37. Green density pp of isotropic Ba-hexaferrite specimens vs. isostatic pressing pressure p.

Figure 38 shows the magnetic characteristics of the compacts mentioned in fig.


37 as a function of density (Stfiblein 1973). The remanence Br rises linearly and
coercivity ]/-arc falls linearly with the density. The rise in r e m a n e n c e is understand-
able as the specimen m o m e n t is roughly proportional to the density. The
coercivity decrease at first appears to be due to plastic deformation of the
crystallites. However, the jHc difference remains qualitatively the same even after
annealing at 1000°C so that density and interaction effects cannot be ruled out.
Bungardt et al. (1968) and Maurer et al. (1972) studied the effect of grain
fineness on density by crushing powder for different periods (up to 240 h) and in
HARD FERRITES AND PLASTOFERRITES 491

(BH)m

o.3- ,i f

O.f
0 O-- 0
l 0 I t i
20 4.0 9 /c rn s 5.0
Green density.gp
Fig. 38. Remanence Br, (BH)max,normal and intrinsic coercivities 8He and jHc, and fullness factor
pp
vs. green density of specimens from fig. 37.

different mills and then compacting samples at constant pressure. Depending on


the type of mill there was a more or less rapid decrease in apparent density from
about 66 to 47% as fineness increased. Possible causes for this fall with increasing
fineness are poor removal of air during pressing, a rather unfavourable particle-
size range and the drop in density owing to the Ba-hydroxide/Ba-carbonate layers
on the powder particles (cf. section 2.1.4). With the same pressure in the range
p = 1 to 4 kbar isostatic compaction produces 5-8% higher density values than die
pressing (Drofenik et al. 1970).
In commercial manufacturing considerable importance must be attached to
uniform density throughout the compact in addition to the overall density
achieved by compaction. However, die pressing produces non-uniform densities
owing to friction between the powder and the die wall (Gasiorek 1980). As this is
a surface effect, compacts with a high height-to-diameter ratio are particularly
affected. Parts with a height-to-diameter ratio of considerably more than 2 are not
normally manufactured. Sintering results in a largely uniform density, producing
non-uniform shrinkage and thus, under certain circumstances, inadmissible
deformation of the sintered parts. The uniformity can be influenced by controlled
relative movement of both punches and the die to one another and by additions
which facilitate pressing (cf. section 2.1.2). With two-sided pressing (fig. 36(a))
both punches execute a movement relative to the die during compaction, the
lower punch generally being fixed. With the same relative movement of both
punches the zone of lowest density ('neutral zone') is roughly halfway up the
green; sintering then produces a cylinder slightly arched inward but deformation
492 H. ST~d3LEIN

as a whole is the least pronounced. With unequal relative movement of the


punches the neutral zone can be shifted to any other point. Conversely, one-sided
pressing (fig. 36(b)) results in relatively large deformation with the neutral zone
located on the surface of the stationary punch.
For pressing isotropic grades the granulate (cf. section 2.1.4) is fed from powder
reservoirs into the die cavity whose volume determines the amount to be pressed.
As the apparent density of the granulate may be up to about 100% higher than
that of powder, a lower filling height in the die cavity and smaller punch
movements during compacting are sufficient for a given mass of the part to be
produced. However, the largest granulate particle must be appreciably smaller
than the minimum wall thickness of the part to be pressed. With this volumetric
feeding it is obviously important that the apparent density of the granulate
remains constant in order to ensure as few size and weight variations of the parts
as possible. With fairly small presses rates of 1 piece per second can be achieved,
with rotary-table machines up to 3 pieces per second. The pressures applied range
from 0.5 to 2 kbar, thus producing green densities of about 50 to 60%. The low
plasticity of the particles and wall friction cause rejects due to spalling and "press
cones" at higher pressures (Schiller 1968), which cannot be avoided even when
lubricants are used.
It should be pointed out that in die pressing as a result of the one-directional
movement of the punches a slight orientation of the particles ('compaction
anisotropy') occurs when they are single-crystalline and platelet shaped. The plane
of the platelets is aligned preferably perpendicular to the direction of pressing.
Thus the c-axis, being perpendicular to that plane, is preferably parallel to the
pressing direction. The resulting anisotropy increases to a certain extent with
increasing green density. However, no anisotropy is obtained with purely isostatic
compaction even if the particles are single-crystalline and platelet shaped (Gordon
1956, Pawlek et al. 1957b, Stfiblein 1957, Drofenik et al. 1970).
The possibility of aligning the crystallites in a magnetic field is of major
importance for improving the quality of standard commercial hexaferrites. The
basis for this is the relatively high crystal anisotropy energy Ek = K sin20 with
K = 3 x 105 J/m 3 (3 × 106 erg/cm 3) (section 1.2). The diagram in fig. 39 shows that
polarization Js is deflected through angle 0 from its normal position parallel to the
c-axis of the hexagonal lattice when an external magnetic field H is applied at an

c-AxLs

rticle

Fig. 39. Orientational relationship of a fixed platelet-shaped hexaferrite particle showing hexagonal
c-axis, the directions of magnetic field H (angle o~) and spontaneous polarization J~ (angle 0) (c-axis,
H, and Js are parallel to the sheet).
HARD FERRITES AND PLASTOFERRITES 493

angle o~ to the c-axis; c-axis, J~ and H are on the same plane. The torque per unit
volume which ties Js to the c-axis is

Mc = K sin 20, (3)

i.e., it disappears when 0 -- 0 ° and 90 ° and therefore has a m a x i m u m value K when


0 = 45 °.
T h e torque per unit volume which turns Js in the direction of H is

MI-I = H J s sin(a - 0), (4)


and is therefore at a m a x i m u m when Js is perpendicular to H and disappears when
both quantities are parallel. For the special case where a = 90 ° it is easy to
quantify to what extent Js is tied to the c-axis. From the equilibriu m condition
Mc = M n it follows that sin 0 = H J d 2 K = H / H A where HA = anisotropy field
strength (cf. section 1.2) giving, for example, angles 0 = 18.5, 39 and 72 ° in fields
of H = 400, 800 and 1200 k A / m (5, 10 and 15 kOe) respectively and finally 0 = 90 °
for H = HA = 1260 k A / m (15.8 kOe). When a < 9 0 °, with equal field strength
values H, angles 0 are smaller than those mentioned and Js and H do not b e c o m e
parallel until H ~ ~.
If the crystallites were freely rotatable any small field H would be sufficient to
m a k e all c-axes completely parallel in field direction. In fact this free movability is
not present because mechanical and magnetic forces act even with monocrystal-
line single-domain particles of bulk material. According to Kools (1974, 1975) the
monocrystalline particles in a suspension are present spontaneously in the form of
curled, random-orientated chains representing magnetically largely self-contained
and relatively stable entities. These break up somewhat abruptly in fields between
16 and 40 k A / m (0.2 to 0.5 kOe) and then align themselves in the field direction.
According to W i p p e r m a n n (1968) the aligning of crystallites of dry loose bulk
material starts from H = 30 k A / m (0.38 kOe) which even in this state suggests a
similar chain formation of the crystallites. In compacting a further factor is the
frictional forces due to contact between the particles. T h e possibilities of aligning
polycrystalline particles with a r a n d o m arrangement of the c-axes are, of course,
even worse.
In manufacturing oriented magnet grades d.c. fields of about 400 to 800 k A / m
(5-10 kOe) are used, as experience has shown that good crystal orientation can be
achieved in this way and fields of higher strength are considerably m o r e expen-
sive. Fields of higher strength would in fact be desirable to further improve
alignment because if the crystallites are already aligned to a certain extent and
angles a (fig. 39) are therefore correspondingly small (at least under 45 °) a
m a x i m u m torque Mc is achieved when J~ is parallel to H, i.e., only when H ~ ~.
According to W i p p e r m a n n (1968) pulse fields can also be used for aligning. H e
found that m o s t of the aligning was completed when H = 100-150 k A / m (1.26-
1.9 kOe) which was the case after 0.2 to 0.3 ms with the form of pulse used.
To manufacture anisotropic grades either dry powder or a suspension', in most
494 H. STJ~BLEIN

cases aqueous, is used; we therefore distinguish between a dry and a wet


compaction process. In both cases monocrystalline or quasi-monocrystalline par-
ticles are required. With dry pressing filling the die cavity is not without problems
as the fine powder does not flow very well and volumetric filling can produce size
and weight variations in the compacts. The powder can be successfully 'sucked in
magnetically (fig. 40(a); Richter et al. 1968b). In this method the powder reservoir
(5) is pushed over the die (1) and the die cavity closed by the lower punch (4). An
inhomogeneous field is induced by coil (2) which, together with the effect of
gravity and suction caused by the upward movement of the die, ensures that the
cavity is filled, fig. 40(b). The powder fills the cavity fairly uniformly although
under the action of the field it tends towards a radiated filamentary consistency.
After the powder reservoir (fig. 40(c)) is removed, the upper punch (3) lowered,
the orienting field (fig. 40(d)) switched on, and after compacting (fig. 40(e)) and
demagnetization by a counter field, the compact is ejected and the die 'drawn off'
(fig. 40(f)). This sequence of punch and die movements is therefore called the
'draw-off method' (Fischer 1962) and is also applied in wet pressing. In this case
the suspension is usually injected into the closed die cavity through passages in
the die or in a punch and aligned in a magnetic field, the material compacted,
water expelled, the compact demagnetized and the die drawn off. During the
compacting of oriented specimens slight disorientation occurs because the crystal-
lites prevent one another from adopting the optimum magnetic position when
being pressed against one another (Stuijts 1956, St/iblein 1973, Kools 1978b).
The suspension can be pressed into the die cavity at a pressure of several bar.
Recent machines operate at pressures of up to 300 bar (Fischer 1978). As the
filling period is short and a certain amount of compaction and dewatering takes
place in the filling phase shorter strokes and thus shorter cycles are possible. A
particular advantage for segment magnets is the more radial field formation in the
die cavity at the start of the punch movement. Figure 41 shows a hydraulic press
with a maximum tonnage rating of 1 MN and a high-pressure filling device
operating at a maximum pressure of 300bar. In order to further reduce the
pressing time per piece, dies with several cavities can be used both with the wet
and dry methods. With wet pressing the chamber filling method has proved
successful. Here, the suspension is first poured into a central chamber from where
all the die cavities are filled at the same time. This procedure facilitates uniform
filling of all cavities and thus makes for~ uniform density of the compacts.
Obviously such tools are expensive and only worth while for large production
runs.
Odor et al. (1977) suggested a process for manufacturing single segment
magnets with magnetically different zones. A pasty ferrite/water mixture, exhibi-
ting high coercivity after sintering, is fed into the die cavity either at one or both
segment ends while for the other zones a mixture which, once sintered, exhibits
high remanence is introduced into the other parts of the die cavity. A compact is
thus produced which is particularly well suited to withstand the different magnetic
stresses set up in motors.
Strijbos (1974) investigated the dewatering stage in detail. With an anisotropic
HARD FERRITES AND PLASTOFERRITES 495

"• ~ O
8~

09

0J
E ~ .-=
- i e-, ~ A .k

t.f3

-4

©,-
---- z- . . . .

©
496 H. STJ~d3LEIN

Fig. 41. Hydraulic press, tonnage rating i MN, with a high pressure slurry pump for filling pressures of
up to 300 bar for pressing oriented hexaferrites by the wet method (courtesy of Fa. Dorst, D-8113
Kochel; see also Fischer 1978).
HARD FERRITES AND PLASTOFERRITES 497

filter cake the permeability for water is anisotropic and 2 to 2.5 times greater in
the direction of the magnetic field than perpendicular to it. This is a consequence
of the chain-like structure of the suspension which is retained at least during
initial dewatering.
To remove the water during pressing passages (holes) are incorporated in one
or both punch faces which are covered with filters. 10 to 13% by weight of water
remains in the compact which corresponds to the free space between the particles.
This water must be carefully removed (dried off) as otherwise the compacts may
crack. The compaction pressure of around 0.5 kbar usually is smaller than in the dry
process.
Both methods have their advantages and drawbacks. Wet pressing offers the
advantage of better magnetic values due to improved crystal alignment, ease of
preparation of a suspension rather than of a powder, lower compaction pressure
and more uniform density of fairly large pieces. The advantages of dry pressing
are simpler tooling, the possibility of making smaller parts, shorter cycle and
finally dry compacts. Mechanical presses are generally used for dry pressing,
hydraulic ones for the wet method.
The spatial arrangement of the field for alignment must be adapted to the
desired crystal orientation in the compact. Homogeneous magnets with equal pole
strength values (fig. 42(a)) can only be obtained if a homogeneous field is induced
for alignment. If the die cavity is in the diverging part of the field then the
magnets become inhomogeneous (fig. 42(b)) which is reflected by unequal pole
strength values (Steingroever 1966). With segment magnets radial (fig. 42(c)) or
diametral preferred directions (fig. 42(d)) can be produced by an appropriate field.
Principally, field and direction of pressing can either be parallel or perpendicular
to one another. Under certain circumstances, the extremely inhomogeneous fields
at the punch edges can impede uniform powder distribution. If necessary, these
fields can be eliminated by a non-magnetic covering.

2.1.6. Final sintering; shrinkage; grain growth


Final sintering, also referred to as sintering, completes the conversion, begun by
the pressing operation, of the loose powder into a compact shape and produces a

..... y
I.t'I I#II#W

Fig. 42. Anisotropichexaferritecompactshavinghomogeneous(a), inhomogeneous(b), radial (c) and


diametral (d) preferred directionsof magnetization.
498 H. STABLEIN

magnetically favourable microstructure with crystal sizes in the order of mag-


nitude of 1 p,m.
During sintering the relative density increases from about 55-65% to over 90%,
and sometimes even to about 98%. Since the mass of the compact remains
constant, disregarding side effects, such as evaporation of moisture and lubricants,
release of CO2 from residual carbonates, oxidation of abraded metallic particles
from the grinding media, etc., the volume therefore decreases to approximately
two thirds of the initial value. The reduction in linear dimensions due to the
extensive elimination of pores, i.e., shrinkage, amounts to about 16% with the
isotropic piece and about 22 or 13% with the anisotropic piece in a direction
parallel or perpendicular to the preferred direction, respectively.
The crystal recovery, recrystallization and crystal growth (continuous and
discontinuous (abnormal) crystal growth) taking place during sintering are desir-
able in one respect and undesirable in another. While at a relatively low
temperature coercivity jHc increases with crystal recovery, as already mentioned
in section 2.1.4, it falls again as a result of crystal growth when the temperature is
set to the value needed for shrinkage.
The driving force behind all this is the decrease in surface and grain boundary
energy and the potential energy elastically stored in the lattice. At a sufficiently
high temperature it causes the atoms to diffuse through lattice defects, e.g., vacant
sites or grain boundaries, which of course play an important role as regards the
extent and rate of sintering. As practice has shown, the presence of further solid
or, at elevated temperatures, liquid phases is of particular significance.
Samples of high-coercivity barium or strontium hexaferrite crystals, i.e., fine
crystals with few defects, crystallized out of salt melts, for instance, do not
undergo shrinkage even at temperatures of 1300°C, well above the temperatures
normally used (Arendt 1973a). Only after the addition of 2 to 4% by weight of
sintering aids, some of a vitreous nature with the composition Z1-Z2-A1203
(Z1 = PbO, BaO or SrO; Z2 = SiO2 and/or B203) were good sintering properties
obtained. The author assumes that these materials form a liquid phase during
sintering.
The influence of the molar ratio n = Fe203/MO ( M = Ba, Sr, Pb) has been
investigated repeatedly. Dealing with barium hexaferrite, Stuijts (1956) found that
with stoichiometric composition shrinkage is relatively small and grain growth
proceeds normally, whereas even a slight excess of BaO (n < 5.9) permits dense
sintering owing to the formation of BaO-Fe203. In that case, if the temperature is
sufficiently high, abnormal grain growth can occur. Similar results were recorded
by Lacour et al. (1973a, b, 1975a, b) and St~iblein (1978) for barium hexaferrite,
Krijtenburg (1974) for strontium hexaferrite and Pawlek et al. (1957a) and Mahdy et
al. (1976a) for lead hexaferrite. According to Bungardt et al. (1968) the activation
energies for sinter compaction in the range from 1120 to 1280°C amount to 770 +_ 12
or 5 5 6 + 12 kJ/mol (184--3 or 133-+ 3 kcal/mol) for samples of BaO.5.9Fe203 or
BaO.5.3Fe203 milled for a relatively short time. From this and the time dependence
of compaction it is concluded that volume diffusion is the main transport mechanism.
With increasing fineness of the powder, values between 835 and 1256 kJ/mol (200 and
300 kcal/mol) were found which is due to a retardation of the diffusion by BaCO3
HARD FERRITES AND PLASTOFERRITES 499

layers deposited on the particle surfaces, cf. section 2.1.4. Cho et al. (1975b) give
544 + 84 kJ/mol (130 + 20 kcal/mol) as activation energy for the grain growth of
barium hexaferrite sinter specimens, irrespective of whether the material is a pure or
SiO2-bearing ferrite.
Deviation from the stoichiometric composition and the presence of other
materials (additives) must not be viewed separately. From all experience gained so
far as it can be said that these conditions lead, in principle, to a reduction in the
saturation polarization of the magnet material as a result of the formation of
non-magnetic phases or a lowering of the saturation polarization of the hexafer-
rite phases. But since the saturation polarization of the hexaferrites is in any case
already low compared with that of other permanent magnet materials, only
relatively small quantities of these additives are normally permissible. Depending
on the controls to be applied to the manufacturing process and on the properties
specified for a certain magnet grade an optimal compromise must be made. The
precise mechanism by which they act is not known in all cases. Some additives can
be substituted in the hexaferrite lattice, such as A1, Cr and Ga in Fe sites (Bertaut
et al. 1959) whereby the saturation polarization in particular as well as the
anisotropy field and consequently the coercivity of the hexaferrite phase are
affected. In this connection the effect of SiO2 additives was investigated in detail
and it was found that they cause the formation of low-melting and, in some
instances, vitreous eutectics with 50 to 60% mol SiO2 (Haberey 1978, Kools 1978a,
Stfiblein 1978, Kools et al. 1980). In addition to Fe203 the eutectics contain a
relatively large proportion of alkaline earth oxides. This overproportional alkaline
earth consumption explains why the feed in industrial production must have a
10% excess of alkaline earth over the stoichiometric value. Because of the low
melting temperatures, sintering of the magnet specimens takes place in the
presence of a liquid phase. Kools (1978a) attributes the grain-growth-inhibiting
effect to the formation of a solid phase which mainly exists on the most strongly
reacting crystal surfaces parallel to the c-axis. Figure 43 shows which value
combinations can be obtained for remanence Br and coercivity jHc with samples
sintered to different densities if they contain no SiO2 (curve A) or 0.5% SiO2
(curve B) (Krijtenburg 1965, Stuijts 1968). Table 4 shows some of these additives.
Additives of SiO2 and A1203 are therefore quite common in production. The
effects of other materials such as B203 or sulphates are disputed because the
specific conditions of each case probably determine whether the positive effects
outweigh the negative ones or vice versa (Krijtenburg 1965, 1974). Some additives
are already present as desired impurities in the main constituents, such as SiO2 in
the iron oxide raw material, cf. table 3. Others are introduced in the form
mentioned in table 4 or as a compound or combined additive, such as AI-
(Granovskii et al. 1970), Ca- (Miiller et al. 1959, Pingault 1974, H a m a m u r a 1977)
or Pb-silicate (Ruthner et al. 1970), as borate (Arendt 1973a), boric acid+ SiO2
(Harada 1980) etc. Many other substances have been used as additives which were
ineffective or worsened matters at least if present in proportions exceeding certain
values. The latter include, e.g. TiO2, MgO, NiO, SnO2, V205 (Kojima 1955a, 1958,
Chroust 1972, Gadalla et al. 1976).
The sintering rate and density can also be influenced by the partial pressure of
500 H. ST~d3LEIN

N O

~ O'~
~A

+ + ÷ +

m-

+
+ I

.=
+ + ÷ + +

'7

<

e~
<
HARD FERRITES AND PLASTOFERRITES 501

~,~0"~i.
xx x

mT ~ •

dO0 - "~,-~''~ "--o-"Q~°° °

%
x i ° • o° o

"\~.

\
Io
I
I
t
x •

01~ i i xl
28 O0 200 300 kA/m zOO

Fig. 43. Remanence Br vs. intrinsic coercivity jHc of Sr-hexaferrites sintered to various densities.
Specimens along curve (A) contain no silica, those along curve (B) contain ca. 0.5% SiO2 (Krijtenburg
1965, Stuijts 1968).

oxygen in the sintering atmosphere. Tests with SrO.5.5Fe203 (Sutarno et al. 1971),
SrO-5.9Fe203 (Reed et al. 1975, Klug et al. 1978) and PbO-5Fe203 (Mahdy et al.
1976a) have shown that amenability to sintering increases as the partial pressure
decreases, which is attributable to the growing concentration of vacant sites.
Intensive fine grinding prior to sintering is necessary not only because particle
sizes around 1 ixm are a prerequisite for high coercivity jHc (section 2.1.4) but
also because of the significance of surface energy as the motive force of com-
paction. The sinter density that c a n b e obtained increases at first as the fineness of
the powder increases. According to Maurer et al. (1972) it reaches a maximum
with barium hexaferrite powder with a specific surface area of 5 mZ/g. As the
fineness increases further the density can fall again. This may be due to the
increasing influence of the BaCO3 layers (cf. section 2.1.4) because they inhibit
diffusion and disintegrate giving off CO2. If the latter occurs with closed pores
then the C02 gas pressure can lead to local swelling and the surface can develop
blisters (Bungardt et al. 1968).
A cause for the stop of further densification is the start of discontinuous grain
growth. Whereas with normal grain growth the pores are usually located at the
crystal boundaries, cf. fig. 49, fast growth causes them to be enclosed within the
crystals, cf. fig. 50. The further removal of pores within the crystals to achieve
more densification must take place by means of volume diffusion of vacant sites, a
much slower process than the grain boundary diffusion of the vacant sites.
The following graphs are to show the influence of the sintering temperature on
different physical properties of the sintered compact. Figure 44 shows shrinkage,
502 H. ST~/d3LEIN

©
tO

tO 20 preferred dLrection x~ 1 2" "£

~ 10'
tO
ll/,,"Slproferred {
0

1-- #~ ,50 %

o~
"o

J 9-.~---x~-'J~L' ,,.
t2o ~
tlO4
/O

/
6000
Nlmm 2
~ ~ooo

2000

0
0 500 1000 °C 1500
Sintering tempereture
Fig. 44. Shrinkage, shrinkage ratio, apparent density, porosity and Vickers hardness vs. sintering
temperature of anisotropic Ba-hexaferrite specimens (St/~blein 1968a).

density and hardness of anisotropic Ba-hexaferrite specimens, properties which


begin to change from 800-900°C upwards (Stfiblein 1968a). The shrinkage S Ip
parallel to the preferred direction is, in the case under consideration, about 1.6
times greater than the shrinkage S" perpendicular to it. This ratio is a con-
sequence of the platelet-like growth of the crystallites. As the latter can be
controlled by SiO2 additives (Stuijts 1968), the ratio of both shrinkages is,
perforce, also affected. As the SiO2 content decreases, S"/S" increases sharply. In
contrast to the above-mentioned properties the permanent magnet characteristics
(fig. 45) change from a few 100°C upwards, particularly coercivity jHc, as already
described in section 2.1.4 for the annealing of finely milled powder. The other
characteristics also increase to a certain extent, especially those measured in the
preferred direction. As a function of sintering temperature, all magnetic values
attain, in both directions, peaks whose location and amount depend on the
composition and manufacturing conditions. With increasing milling time, for
example, the temperature of the curve peak falls but the peak itself rises (Sixtus et
al. 1956). The fact that the peaks measured parallel to the preferred direction
H A R D FERRITES AND PLASTOFERRITES 503

measured II preferred direction


/ - ,~

,~00 '201 xx~x ;

mT

300 ,~o [.I ~1 I I

c
13
200 ,6o .' 7--~-~/ i /
E
g~

1oo so . . ~.~___+.____÷ /. , :

- : V 7 " - - -~U
Lo'-- b~
I BHc (BH) I , + +.,.+-%. ~
I . _ ../ . . . . +- -- -1"- . . . . . I "+4"4
0 C
0 500 1000 °C 1500
Sintering temperature
Fig. 45. Remanence Br, (BH)max, normal and intrinsic coercivity eH~ and jHc, resp., vs. sintering
temperature of anisotropic Ba-hexaferrite specimens (Stfiblein 1968a).

occur at different temperatures (cf. also Harada 1970, Gershov 1971) is, in
addition to the composition and preparation, the reason why there are anisotropic
grades with high remanence Br and low coercivity jHc and vice versa. While with
most grades the demagnetization curve exhibits a bulge in the preferred direction
the demagnetization curve perpendicular to the preferred direction is a straight
line. In this case the peaks of Br, (BH)ma× and BHc lie at the same sintering
temperature. Similar conditions apply to isotropic grades although only higher
values of the properties occur. Therefore with isotropic magnets there is practic-
ally only one grade, irrespective of provenance and manufacturer, with narrow
ranges for Br, (BH)ma× and BHc, there being only a certain amount of latitude left
for 1He.
While density and magnetic characteristics normally exhibit a smooth curve in
the temperature range from 1100 to 1250°C, Heimke (1958) observed with barium
hexaferrites an abnormal trend around 1150°C, the c~-fi-transformation tem-
perature of CaSiO3, after an amount of this material corresponding to 1% SiO2
had been added to facilitate sintering.
The graph of the properties of anisotropic specimens above the sintering
temperature shows that the temperature must be kept constant to within 10°C in
order to retain uniform quality of the sintered pieces with a given condition of the
compacts. However, the sintering period is also important. With isothermal
504 H. STJ~d3LEIN

Range: : T : j ,r_, i
Sintering temp. : 1000 1100 1170 1200°C
I I I I
A mount of orientation ~-- 1,0
/ ~ 0,9
I l l meesured porallel to
400 320
I1 " perpendiculor tc
mT kA/m ~ direction . tl(i) "/ / .0.8

I O.7

300 24,0
J (a). "
/> " 0.6

¢11
~ 200 .~ 160
E

iI 80

0
2.6 3.0 4.0
i
~-~..~_

........

A p p a r e n t density
g/cm 3
~--

i--

5.0 5.3
0.2

0.1

Fig. 46. Remanence Br and intrinsic coercivity ~Hc vs. apparent density of Ba-hexaferrite specimens
pressed with (a) and without (i) orienting magnetic field (St~iblein 1968a).

sintering relatively coarse powders shrink according to a power law whereas very
fine powders shrink according to a logarithmic law (Bungardt et al. 1968). The
influence of the pressure decreases with increasing compaction (Sutarno et al.
1970c, St~iblein 1973).
Considering the values as a function of density provides a more quantitative
view of the phenomena during sintering. The remanence values, measured
parallel and perpendicular to the preferred direction, and coercivity jHo,
measured parallel to the preferred direction, are shown in this way in fig. 46. The
values marked (a) relate to the same specimens as in figs. 44 and 45, whereas
those marked (i) refer to specimens which were manufactured from the same
powder as (a) but without the application of a magnetic orienting field. The
degree of alignment is shown as a dotted line. ~: = 1.0 corresponds to the
saturation polarization, numbers ~ < 1.0 to fractions of it. With the isotropic
specimen the degree of alignment is sc = 0.5. There are three density ranges, I, II
and III, which correspond roughly to temperature ranges up to 1000, 1000 to 1200
HARD FERRITES AND PLASTOFERRITES 505

and above 1200°C, respectively, and in each of which characteristic phenomena


occur.
The remanence values only increase in temperature range II almost linearly but
not quite proportional to the density. There are two reasons for this: the shearing
effect caused by porosity (Denes 1962) and the change in the crystal texture
caused by crystal growth (Rathenau et al. 1952). If the shearing effect is con-
sidered mathematically the dot-dash curves obtains in fig. 46. For the (almost)
isotropic specimen this gives degrees of alignment of sc = 0.55 and 0.464 measured
parallel and perpendicular to the direction of pressing respectively, regardless of
specimen density, as is reasonably expected. Conversely, the degree of alignment
of the anisotropic specimen only rises slightly at first in range II and then
considerably in range III (Reed et al. 1973). This sharpening can also be
quantitatively verified radiographicaUy, as fig. 47 shows, and also with the angular
dependence of the remanence (Stiiblein et al. 1966a).

50-
c~

8 40.
"s
30-

20.
J
10-
g.
o I I I
11100 12100 1300 oIc t~O0
Sintering temperature
Fig. 47. Pole densities of the basal plane of barium ferrite parallel to the preferred direction (St~iblein
et al. 1966a).

Figures 48--50 show the structure of the specimens sintered at 1000, 1225 and
1385-1395°C. It is easy to see the transition from the initial, very porous, state (fig.
48) to a rather dense microstructure with at first relatively slight, continuous
crystal growth and intercrystalline pores (fig. 49). This is followed by the ap-
pearance of a discontinuous, pronounced laminar crystal growth with in-
tracrystalline pores (fig. 50).
The phenomena in ranges I to III (fig. 46) can therefore be summarized as
follows:
Range ! is characterized by the curing of lattice defects, which can be best seen
in the substantial rise in coercivity and a smaller rise in remanence. Compaction
of the specimens and a relatively slight grain growth take place in range II.
Coercivity gradually drops and the crystal texture of anisotropic specimens
becomes somewhat sharper. In range III exaggerated crystal growth is the cause
of the sharp drop in coercivity and the noticeable sharpening in the texture of the
506 H. STABLEIN

Fig. 48. Micrographs of Ba-hexaferrite specimens sintered at 1000°C, not etched (Stfiblein 1968a): (a)
anisotropic specimen cut parallel to preferred direction, (b) as (a) but cut perpendicular to preferred
direction, (c) isotropic specimen.

Fig. 49. Micrographs of Ba-hexaferrite specimens sintered at 1225°C, etched with 50% HCI at 80°C
(Sti~blein 1968a): (a) anisotropic specimen cut parallel to preferred direction, (b) as (a) but cut
perpendicular to preferred direction, (c) isotropic specimen.

Fig. 50. Micrographs of Ba-hexaferrite specimens, etched with 50% HC1 at 80°C (Sti~blein 1968a): (a)
anisotropic specimen, sintered at 1395°C, cut parallel to preferred direction, (b) as (a) but cut
perpendicular to preferred direction, (c) isotropic specimen, sintered at 1385°C.
HARD FERRITESAND PLASTOFERRITES 507

anisotropic specimen. That this sharpening is linked with an increase in


remanence in spite of a sharp jHo drop, shows that the nucleation for remag-
netization is made more difficult, i.e., it only begins with a negative field. On the
other hand the local stray fields of the crystallites in the isotropic specimen
promote the formation of remagnetization nuclei in positive fields (Rathenau et
al. 1952) so that remanence becomes less than half of the saturation polarization.
Further tests showed that the same magnetic values can be achieved in range II
by using various temperature/time combinations, provided that sintering was
adjusted to produce equal density values. It is therefore unimportant for the
magnetic values whether a shorter annealing time is used at higher sintering
temperature or vice versa and how high the heating and cooling rates are.
Shrinkage and grain growth cannot therefore be influenced independently by
varying the sintering temperature. This can be most simply interpreted by
assuming a common elementary process for both phenomena (Stfiblein 1968a).
Reference has already been made above to the anisotropic shrinkage of
oriented specimens which is caused by the anisotropic growth rate of the hexafer-
rite crystals and which results in their platelet-like shape, cf. section 2.1.3 and figs.
49 and 50. Other properties, too, are more or less anisotropic, cf. sections 3.3 and
3.4, in particular thermal expansion, which amounts to about 14 or 10 x 10-6/K
parallel or perpendicular to the c-axis. This results in particular difficulties in
manufacturing toroidal magnets with a radial preferred direction (Kools 1973).
Pieces which have been die pressed in the normal way break when heated owing
to inhomogeneity of density and alignment. The inhomogeneities can be reduced
by isostatic compaction and the compact strength increased so that the pieces can
be heated without breaking. The stresses in the toroid which then occur with
anisotropic shrinkage during sintering can balance each other out by creep; thus
producing an unbroken piece at the end of the sintering cycle. During cooling,
however, stresses occur as a result of the different coefficients of expansion. These
stresses can result in radial cracks at the hole (maximum tangential stress) and in
tangential cracks with an average diameter (maximum radial stress). They can be
avoided if the ID/OD ratio of the toroid exceeds 0.80 to 0.85 because then the
maximum stresses during cooling are below the ultimate strength.
In large-scale commercial production sintering is usually carried out in electric-
ally- or gas-heated continuous kilns (Petzi 1974a, 1975, 1980) where the compacts
are stacked on plates and pushed through, see fig. 51. Because air is needed as
sintering atmosphere these kilns are simpler than those for sintering soft magnetic
ferrites which require a protective atmosphere. Barium and strontium hexaferrites
are predominantly sintered at 1200 to 1250°C for several hours and the heating
and cooling steps both require another 5 to 10 hours, with larger pieces even
longer. Recently, however, also faster and more economic sintering techniques
have been considered. Lead hexaferrites can and must be sintered at a tem-
perature several hundred degrees lower even if only to minimize evaporation
losses of PbO. Batch kilns are used for smaller quantities and in the laboratory.
Fairly large batch kilns can also be used if special burners (jet burners) and rapid
air circulation are employed (Remmey 1970, Bohning 1978).
508 H. ST~d3LEIN

Fig. 51. Electrically heated pusher type kiln for sintering of hard ferrites. Ceramic conveyor trays are
automatically returned. Length of furnace 11 m, useful width and height 280 × 100 mm 2 (courtesy of
Fa. Riedhammer, D-8500 N~rnberg).

2.1.7. Machining (grinding etc.)


After sintering the magnet parts have dimensions which correspond only to a
varying degree to the required dimensions. With good manufacturing methods
they lie within a range below the limits listed in table 5. According to the German
magnet standard D I N 17410 (May 1977) these limits are admissible for the supply
of unground magnets unless manufacturer and purchaser agreed otherwise.
Dimensional variations are the result of variations in the raw materials supplied
and individual production steps, inhomogeneities in the material and processing
conditions. This needs to be explained in more detail. With isotropic homo-
geneous shrinkage sintered pieces and compacts are exactly similar in the
mathematical sense and dimensional variations from piece to piece only occur as a
result of different degrees of shrinkage, e.g., owing to different green and/or sinter
densities. The situation is similar with anisotropic, homogeneous shrinkage where
there are merely different degrees of shrinkage for different spatial directions as is
the case with oriented magnets. Additional dimensional variations occur with
inhomogeneous shrinkage. A more or less deformed line stems from a straight
edge, a more or less concave or convex surface from a plane, two parallel edges or
surfaces are inclined towards one another or warped, concentric circles become
HARD FERRITES AND PLASTOFERRITES 509

TABLE 5
Permissible deviations in size of hard ferrite magnets (in mm) to D I N 17 410 (May
1977).

Isotropic hard ferrites Anisotropic hard ferrites


perpendicular in perpendict/lar in
to pressing pressing to pressing pressing
Rated size direction direction direction direction*
from to ___ ± -4-

4 0.25 0.30 0.25 0.30


4 6 0.25 0.30 0.25 0.30
6 8 0.25 0.30 0.25 0.30
8 10 0.30 0.40 0.30 0.40
10 13 0.30 0.40 0.30 0.40
13 16 0.30 0.40 0.35 0.45
16 20 0.30 0.40 0.45 0.55
20 25 0.30 0.40 0.55 0.70
25 30 0.35 0.45 0.70 0.90
30 35 0.40 0.50 0.80 1.00
35 40 0.45 0.55 0.95 1.20
40 45 0.50 0.60 1.10 1.35
45 50 0.60 0.80 1.20 -
50 55 0.70 0,90 1.30 -
55 60 0.75 1.00 1.45 -
60 70 0.90 1.10 1.65 -
70 80 1.10 1.35 1.90 -
80 90 1.25 1.55 2.15 -
90 100 1.40 1.70 2.40 -

* Wet pressed hard ferrites are machined on the pole faces.

non-concentric and deformed etc. The causes for this are inhomogeneities in the
material and varying processing conditions. Further configurational deficiencies
are burrs on the edges, undesired sintered-on loose particles etc.
In all these cases where these geometric defects are intolerable the piece must
be machined. This is especially advisable for the pole faces of the magnet when
otherwise unnecessary air gaps occur in the magnetic path which absorb too much
magnetic energy. Owing to the brittleness of the material only such machining
methods can be considered which produce very fine chips, e.g., grinding, tumbling,
lapping or polishing. Diamonds are predominantly used as grinding medium,
owing to their hardness even compared with ceramics, and sometimes silicon
carbide. Good cooling is always necessary. In recent years highly advanced
processes for grinding segment magnets have been used where, for example, the
magnets, lying behind one another on guideways, are pushed through a gap
formed by 2 grinding wheels. Both inside and outside radii can be ground in one
operation.
A number of fundamental investigations were carried out on the very com-
plicated mechanism of grinding ferrites (Broese van Groenou 1975, Veldkamp et
al. 1976, Broese van Groenou et al. 1977). Zones of tensile and compressive stress
510 H. STJ~3LEIN

with plastic deformation, brittle fracture and crumbling occur. Scratching on the
basal plane results in more crumbling than on the prism face; other factors such as
power requirement, and specific grinding energy are anisotropic, too. One parti-
cular result was that the specific grinding energy does not depend very much on
the circumferential speed of the grinding wheel. During high-speed grinding with
speeds of up to roughly 100 m/s a faster removal rate is possible at constant shear
force, i.e., grinding time saved, or less energy can be applied with the same rate of
removal, thus producing less deformation in the piece.
Reference should be made at this point to the manufacture of magnets by
indirect shaping. Normally the sintered piece already has its final shape and size
("direct" shaping), apart from corrective grinding. It may, however, be advisable
or necessary to produce smaller parts from larger, pressed or sintered stock, i.e., as
semi-finished parts, e.g., because they have better magnetic properties or because
a cylinder with diametral preferred direction is easier to manufacture in this way.
Assuming the pieces are sintered, precision machining methods are used in this
process, too, e.g., parting off by cut-off wheels of pieces less than 1 mm thick. With
pressed pieces actual machining is relatively easy to carry out but owing to their
low strength they require careful handling.

2.1.8. Magnetizing and demagnetizing


In as-manufactured condition a permanent magnet is either totally non-magnetic
or only incompletely magnetized. It is usually magnetized after being incorporated
into the magnet system. The required amount of field strength effective in the
permanent magnet is 2 to 3 times as great as its coercivity jHc (St/iblein 1963,
Dietrich 1969). The direction and spatial directional distribution of the magnetiz-
ing field are brought into line with the desired polarized state of the permanent
magnet. How well this must be done depends, among other things, on the
material. With isotropic magnets the remanence accurately reflects the directional
distribution and any inhomogeneities of the magnetizing field after it has been
switched off. With anisotropic magnets, however, the preferred direction largely
determines the remanent state. Figure 52 shows that remanence is independent of
/3 where deviations/3 of the magnetizing field from the preferred direction are not
too great. This is due to the fact that the polarization direction in every crystallite
in the remanent state runs parallel to the local c-axis. With complete alignment of
all crystallites remanence would be independent of angle/3 (apart from a certain
small range around/3 ~ ~-/2). With actual specimens it depends on the degree of
alignment. With the well-aligned 1350°C specimen in fig. 52 remanence only drops
from/3 ~ 50 to 60 ° upwards, whereas with the less well-aligned 1100°C specimen it
starts falling from/3 ~ 30 to 40 ° upwards.
In many cases a permanent magnet is not used in the fully magnetized state.
One reason for this may be that a certain magnetic flux is required of the magnet.
However, in large production runs geometrical and magnetic variations occur
from piece to piece. These are eliminated by defined, individual weakening
('calibration').
Another reason is the higher stability of the weakened condition (Gould 1962,
HARD FERRITES AND PLASTOFERRITES 511

Preferred axis

Remanence~ kfagnetizing field

~ Remanence r2

Perfect alignment
~" 1.0 _ / _ ~ ~ 5 0 o c
1250°C~\
,

si, feting mp t e\
\\
E
~ O~

IlO0°C
-1250oC
1350°C
Q~ i i i i i

30 ° 60° ~0 o

Deviation from preferred axis

Fig. 52. Relative remanence values rl and r2 of oriented barium hexaferrite, measured parallel and
perpendicular, respectively, to the preferred direction, after magnetizing at /3° from preferred axis
(St~iblein et al. 1966a).

Dietrich 1968). This condition is achieved by partial demagnetization, generally


using an alternating field of sufficient amplitude which diminishes slowly enough,
and less frequently by means of an opposing field, rotating field or by thermal
treatment. Defined weakening is not simple, particularly with anisotropic grades
owing to their rectangular hysteresis loop, if the alternating field is applied
parallel to the preferred direction as the dependence of the field amplitude is
relatively large, cf. fig. 53. This disadvantage can be avoided when the position is
perpendicular to the demagnetizing field and preferred direction. However, higher
field strengths are required for the same degree of weakening. The above-
mentioned methods can, of course, be used for complete demagnetization ('neu-
tralization') too.
The various weakening or demagnetizing methods result in different domain
structures (Tanasoiu 1972) even if the overall macroscopic condition is the same.
This can be seen, for instance, in the shape of the magnetization curve, cf. fig. 54
(St/iblein 1970). Specimens demagnetized thermally or with an a.c. or d.c. field
react completely differently to any applied field. Moreover, with a.c. demagnetiz-
ing field weakening the temperature and field direction with respect to the
preferred direction are important.
Magnetization and demagnetization techniques have been described in
numerous papers and text books (Rademakers et al. 1957, Dambier et al. 1960,
Underhill 1957, Parker et al. 1962, Knight 1962, Schiller et al. 1970).
~00
rn7

300

¢ 200 ~ 2

I00

00 200 400 600 kA/rn


No --
Fig. 53. Remanence Br of an anisotropic hexaferrite magnet with intrinsic coercivity 1He = 143 k A / m
after full magnetization and subsequent demagnetization in a gradual diminishing a.c. field of initial
amplitude H0. Curve (1): field parallel to preferred axis, curve (2): field perpendicular to preferred axis.

400-
mf

300-

L200-
.7

100. s

O kA/
Fig. 54. Magnetization curves and parts of the outer hysteresis loop (first quadrant) of an anisotropic
barium hexaferrite specimen having Br = 364 mT (3.64 kG) and jHc = 240 k A / m (3.02 kOe), measured
parallel to preferred axis, after different methods and conditions of demagnetization (St/iblein 1970).
Curve (1): thermal; curve (2): a.c. at -196°C, ± preferred axis; curve (3): a.c. at -196°C, [[ preferred
axis; curve (4): a.c. at 20°C,.A- preferred axis; curve (5): a.c. at 20°C, I] preferred axis; curve (6): a.c. at
300°C, ± preferred axis; curve (7): a.c. at 300°C, II preferred axis; curve (8): d.c., starting from outer
loop, third quadrant.

512
HARD FERRITES AND PLASTOFERRITES 513

2.2. Special technologies

In terms of the magnetic values attainable, the manufacturing process described in


section 2.1 has proved to be highly economic but it is not the only one that can be
employed. In the following further possibilities for the manufacture of hard
ferrites are described. In table 6 they are compared diagrammatically with the
TABLE 6
Technologies available for manufacturing magnets.

D~
hnique

0
"i"~
Manufacturing
to
step

Mechanical m i x i n g
P r e c i p i t a t i o n of the raw materials

Reaction in furnace
Reaction in fluidized bed
Reaction in salt bath

Crushing, grinding

Die p r e s s i n g at room temperature


S h a p i n g by rblling, extruding

i
Hot p r e s s i n g

Sintering

Further p r o c e s s i n g
514 H. ST]kBLEIN

conventional process. Some of these processes require fewer main operations, e.g.,
the sing!e-sintering technique where reaction and final sintering take place simul-
taneously, or hot pressing of the powder where shaping and sintering are carried
out at the same time. In other processes one operation is replaced by another.
This includes the precipitation techniques where mechanical mixing is at least
partly replaced by a precipitation reaction, the fluidized bed processes where the
reacting mass is suspended in the form of fine particles in a stream of gas, or
plastic working (rolling, extruding) which, compared with the pressing of a powder
or a suspension of powder particles, offers different possibilities. In the salt bath
processes both mixing and reaction takes place together in a molten bath.
Naturally, the processes can be combined, for instance, co-precipitation with hot
pressing or single sintering with rolling. A review on non-conventional powder
preparation techniques of ceramic powders was given by Johnson (1981). While
offering some advantages, these special technologies generally entail major draw-
backs. With further advances in technology, it may be that one or the other
process will be employed to a greater extent than at present.

2.2.1. Single sintering technique


The characteristic feature of this technique is that only one single sintering
operation takes place, cf. table 6. This means that during sintering the raw
materials must react to form hexaferrite (cf. section 2.1.3) and, at the same time, a
dense, true-to-shape body must be produced (cf. section 2.1.6). Good mixing and
high reactivity of the raw materials are, of course, particularly important in this
process. This is facilitated or ensured by intensive mixing and milling of the mix
and by the addition of substances (e.g. SiO2, PbO, B203) which promote sintering.
Shrinkage on sintering is about 1.5 times that in double sintering as the raw mix
which is shaped generally has a density on pressing of between 2.0 and 2.6 g/era~
at the usual pressures of around 0.5 kbar compared with a typical density of
pressed ferrite powder of 2.7-3.3 g/cm3. This has to be taken into account in the
sizing of the press-working tools. The sintering temperature is roughly equal to
that applied in double sintering. Milling time and sintering temperature have to be
adapted to one another, as is shown in fig. 55 (Stfiblein 1971), to ensure that a dense
and, at the same time, high-coercivity body is produced. Single-sintered parts show a
greater tendency to inhomogeneous shrinkage (distortion) than doubled-sintered
parts. The extent, however, largely depends on the raw material used and the
processing conditions.
Isotropic or at the most weakly anisotopic magnets are normally obtained by
means of single sintering. More detailed investigations have, however, shown that
appreciably anisotropic parts can also be manufactured owing to the fact that the
platelet-shaped hexaferrite crystals grow on plane Fe203 surfaces (section 2.1.3). It
was found that sintering always produces an isotropic magnet if the raw mix
contains iron oxides of isometric shape (spherical, cubical) regardless of other
conditions under which manufacture takes place. Anisotropic magnets, however,
are obtained when mixes are reacted where the iron oxide particles are
anisometric (acicular, platelet-shaped) and if these were oriented during shaping
H A R D FERRITES AND PLASTOFERRITES 515

oc
2.2 0.8 0.2 0 0 0 0.8
1300

7.~ 7.5 Z6 Z9 3.6 0.8 1.2


1250
f ..... . .

E 7.3 Z5 Z6 ', 8.3 8.5 "8.'~ 6.0


" 1200 • ~\ • • jJ •
\ /
L.. 89
6.5 6.9 72 7"6. 01 ./?6 6o
1150

2/ 2.8 ~.5 56 6.6 go 4.8


1100
i t i i i i I

1 2 4 8 16 32 h 6~
Milling time
Fig. 55. (BH)m.x value in kJ/m 3 vs. milling time (vibration mill, wet) and sintering temperature.
Material: mixture of synthetic a-Fe2Oa (Bayer 1660) and BaCO3.

and retained their configuration up to the hexaferrite reaction. Acicular iron


oxides occur, for instance, when moisture is carefully withdrawn from the acicular
a-FeOOH (goethite). However, the particle shape of the natural iron oxides
(hematite) is as a rule isometric. Nevertheless, there are noteworthy exceptions
probably connected with the formation of the iron ore deposits where the
particles show a propensity for cleavage along the basal plane.
Processing is basically the same as with the isotropic magnets. Owing to the
shape of the particles only a few special requirements have to be met. Milling is
not only important for increasing the activity during sintering but also because it
permits the shape of the particles to be influenced. Depending on the oxide used,
particulate aggregates, for instance, not capable of being aligned can be separated
into single alignable particles or an existing anisometry of the particles can be
destroyed so that these lose their alignability. In shaping the anisometric particles
have to be aligned. In dry or wet pressing this occurs when the punch moves into
the die; this is also the case with the platelet-shaped hexaferrite particles (press
working anisotropy, cf. section 2.1.5). The particles can also be aligned by rolling
or extruding a plasticized raw mix. Suitable plasticizing agents are soft waxes, for
instance, with fusion points of around 45°C, which are easy to shape at room
temperature and can be easily evaporated or drained after shaping. Figure 56
shows the influence of the sintering temperature on the demagnetization curve of
a specimen which originally consisted of a mixture of SrCO3 and acicular a-Fe203.
Within a narrow temperature range (a few multiples of 10 degrees) the transition
takes place from the high coercivity to the low coercivity condition,_ the duration
of annealing also making itself felt. This transition is due to abnormal grain
growth. In the case described the growing crystallites were fairly well aligned so
that growth at the expense of the small, poorly aligned crystallites led to an
improvement in texture and thus to an increase in remanence. The remanence
516 H. STJifl3LEIN

400
rnT
300

200

tO0 ~
1 6 l~J _._
b

-400 kA/rn -300


o.~
Field strength H
- tO0 ~-

- 200

-300

- 400

-500
Fig. 56. Demagnetization curves of differently sintered specimens. Starting material: mixture of
synthetic, needle-like c~-FeaO3 (Bayer) and SrCO3, wet pressed.

~00
mT
(BH)max=15.9 k J j 300

200

100
b
Field strength H / ~,I ,~," "~-lb
-30'9 kA/m ~00 /.J~'-~'O0 /
k
too

200

300

-400
Fig. 57. Demagnetization curves of differently prepared specimens after sintering at 1230°C. Starting
material: mixture of natural hematite with pronounced cleavability along basal plane and BaCO3,
Curves (la) and (lb): anisotropic magnet due to wet milling and wet pressing of the mixture, measured
parallel and perpendicular to pressing direction. Curve (2): isotropic magnet due to dry milling and dry
pressing.
HARD FERRITES AND PLASTOFERRITES 517

values attainable correspond to those of double-sintered, anisotropic magnets. As,


however, this high anisotropy occurs owing to crystal growth and thus at the
expense of coercivity, the magnetic values of single-sintered anisotropic specimens
are, as a rule, not as good as those of the double-sintered specimens. It should be
pointed out that with poorly oriented specimens abnormal crystal growth does not
lead to an improvement in remanence. This behaviour is analogous to that of
double-sintered magnets (section 2.1.6).
Figure 57 shows the demagnetization curves of a mix subjected to a different
treatment. Natural iron oxides and barium carbonate were used. After 16 hours of
wet milling in a vibration mill, wet die pressing and sintering at 1230°C an
anisotropic magnet was obtained with a (BH)max value of 15.9 kJ/cm 3 (2.0 MGOe)
in the preferred direction. After 22 hours of dry milling, dry die pressing and
sintering at 1250°C, however, an isotropic body with a (BH)max value of 7.2 kJ/m 3
(0.9 MGOe) was obtained. Even better values can be achieved with synthetic iron
oxides or hydroxides (Takada et al. 1970b, Esper et al. 1974).
Brief mention may also be made of another manufacturing method which is
similar to both single and double sintering. However, in this process anisometric
iron oxide particles are also used and the mix die pressed. After reaction, pressing
is repeated in the same die without prior milling, which is analogous to cali-
bration, and sintering repeated. This process enables oriented magnets to be
produced without using any magnetic field (DE-OS 2 110 489).

2.2.2. Precipitation techniques


With these techniques at least one of the metal ions derives from a feed substance
obtained by precipitation or hydrolysis from an aqueous solution. The term
co-precipitation is used when all cations are derived from a common solution. The
aim of these techniques is to obtain as intimate mixing of the feed materials as
possible from the very start. This facilitates the formation of the compound in the
subsequent reaction step owing to the short diffusion paths. The reaction can take
place at relatively low temperatures and under optimum conditions a powder can
be obtained containing single domain particles of correspondingly high coercivity.
The next logical step would be to obtain the hexaferrite rather than the mix of
feed substances from the aqueous phase. This appears to be possible but its
practicability cannot yet be assessed for hexaferrites.
The advantages offered by these techniques are the following: they can be used
even with contaminated raw materials (self-cleaning effect), there is no carry-over
of impurities by abrasion in milling and a narrow particle-size range is obtained.
The large water requirements must be regarded as a disadvantage for commercial
applications.
In actual operating practice, an aqueous solution of one or several salts is used.
As metal salts, especially chlorides (not with lead) nitrates, oxalates, and acetates
can be considered. The hydroxides or carbonates of these metals, which are
practically insoluble in water (exception: Ba(OH)2), can be recovered from the
solution. This takes place either by precipitation by means of an alkaline solution
(e.g. NaOH, NH4OH) and/or water-soluble carbonate (e.g. Na2CO3, ammonia
518 H. ST~kBLEIN

carbonate) or passing NH3 and C O 2 gas through the solution or by means of the
thermal hydrolysis of the salts. Another technique is called freeze drying, because
a thin stream of aqueous solution is squirted into a cold, immiscible liquid forming
freezed granules there. The granules are separated from the liquid and the ice is
sublimated, while the raw mixture is retained (Schnettler et al. 1968). The particle
size of the precipitated product depends not only on the type of product but
particularly on the conditions under which precipitation takes place and can range
from about 0.01 to 10 txm. The temperature necessary in the subsequent hexafer-
rite reaction depends on the particle size. Hexaferrite single-domain particles of
very high coercivity can only be obtained from very fine feed substances which
react at temperatures of 800 to 900°C.
Table 7 shows some of the results published in literature. The precipitation
techniques can be subdivided into chemical co-precipitation, chemical partial
precipitation, thermal hydrolysis, electrolytic co-precipitation and hydrothermal
treatment. Most of the tests were carried out using chemical precipitation.
In the tests carried out by Sutarno et al. (1967) the metal salts separated in very
different particle sizes (table 7). While the iron hydroxide was amorphous, the
carbonates of barium, strontium and lead were present in particle sizes of around
5 ~xm. As a result, segregation occurred with larger mixes and the formation of
hexaferrites took place in the same temperature range as with mechanically mixed
feeds. Obviously it is of no advantage if only one raw material is highly reactive.
All raw materials must exhibit a sufficient degree of reactivity and steps must be
taken to ensure that they are intimately mixed.
The tests carried out by Haneda et al. (1973b, 1974b) led to amorphous pre~
cipitation products and the main quantity reacted to form hexaferrite after
annealing at 800 to 850°C. After two hours of annealing at 925°C, a particle size of
0.15 ixm (BET-method), a specific saturation polarization of Js/p = 85.16 mT cm3/g
(o-s = 67.8emu) and a c0ercivity jHc of 480 kA/m (6 kOe) were obtained with
non-oriented specimens. The hysteresis loops measured and calculated according
to Stoner-Wohlfarth (1948) compared favourably from which the conclusion can
be drawn that there was coherent reversal of magnetization in these obviously
almost ideal crystals. Further processing by pressing and sintering or by hot
pressing (cf. section 2.2:5) produced partly oriented, high-coercivity specimens.
Similar results were attained by Gordes (1973) and Roos et al. (1977) who, in
addition, succeeded in increasing the coercivity of the particles to 1He = 510 kA/m
(6.4 kOe) and the specific saturation polarization to JJp--88.8mTcm3/g (O's=
70.7 emu) by strong etching in hydrochloric acid with an attendant loss of material
of about 70%. The hexaferrite formation of the co-precipitated mixture took
place between 700 and 800°C possibly without BaFe204 as an intermediate phase
and yielded a small grain size distribution (Roos 1979, 1980), see also section
2.1.3. Mee et al. (1963) used a chemical precipitation technique (not described in
detail) for making platelet-shaped Ba- and Sr-hexaferrite particles with diameters
from 80 to 150nm and coercivities of 425 and 456kA/m (5.35 and 5.75 kOe)
respectively. Goldman et al. (1977) investigated the influence of different
manufacturing parameters in chemical co-precipitation for magnetically soft fer-
rites. Qian et al. (1981) prepared Sr-hexaferrite by annealing a co-precipitated
HARD FERRITES AND PLASTOFERRITES 519

mixture of ferric hydroxide and strontium laurate for 70 days at 550°C. The
mixture was prepared from iron nitrate, strontium nitrate, lauric acid and am-
monium hydroxide. Intermediates are y-Fe203 and solid solutions of SrO in
y-Fe203. No magnetic or grain size data were reported.
While the above-mentioned co-precipitation method mainly served to produce
magnetically ideal hexaferrite particles, tests for chemical partial precipitation
were conducted with a view to establishing possible advantages in the industrial
manufacture of hard ferrites. The iron oxide is not precipitated but exists from the
very start as a solid phase. The alkaline earth component is either introduced in
the form of a solution, e.g., St(NO3)2 (Sutarno et al. 1969) or Ba(OH)2 (Erickson
1962) or initially as a solid phase SrSO4 which is re-precipitated into SrCO3 by ion
exchange (Cochardt 1966). In the latter process, the fact that SrCO3 is even more
insoluble in water than SrSO4 is utilized so that the ion exchange takes place in
the presence of Na2CO3 or (NH3)2CO3. The process has long been known for
making SrCO3 from SrSO4 (Gallo 1936, Ullmanns Encyclopfidie 1965, Sutarno et
al. 1970b) but was used by Cochardt (1969) for making strontium hexaferrite. In
this H - C process (H for hematite, C for celestite = natural strontium sulphate) the
raw materials iron oxide and celestite are ground as an aqueous suspension
together with Na2CO3 in a mill, with size reduction, blending and chemical
reaction taking place simultaneously (Steinort 1974). An analogous approach
cannot be used for barium hexaferrite because BaSO4 is more insoluble in water
than BaCO3.
By thermal hydrolysis of a Ba- and Fe-acetate solution in contact with paraffin
oil at a temperature of 300°C Metzer et al. (1974, 1975) produced amorphous
precipitates with a specific surface area of more than 100 m2/g and particle sizes
from 10 to 20 nm from which hexaferrite powder with a specific surface area of
5.7 m2/g, a particle size of 0.2 Ixm and a coercivity jHo of 420 kA/m (5.3 kOe)
could be obtained by annealing.
Beer et al. (1958) described a continuous electrolytic co-precipitation process
for making powders for isotropic barium hexaferrites. This process uses metallic
cathodes and an alkaline electrolyte; no detailed information, however, is given.
By means of freeze drying Miller (1970) produced barium hexaferrites starting
with a solution of iron oxalate and barium acetate. Other substance combinations
turned out to be more disadvantageous. Specimens with different Fe203/BaO
ratios having different additives of SiO2, A1203, PbO, Bi203, CaO or TiO2 were
dry or hot pressed without application of a magnetic field.
While all the processes described above furnish a raw mix which needs to be
heat treated for the formation of hexaferrites, the hydrothermal process (Takada
et al. 1970a, Van der Giessen 1970) furnishes barium hexaferrite particles directly
as precipitate. No magnetic data, however, were given. With the DS process the
techniques of co-precipitation and spray calcining are carried out simultaneously, see
section 2.2.4.

2.2..3. Melting techniques


As explained in section 1.3, it is not possible to directly separate Ba(Sr)Fe~:O19
crystals from a melt of alkaline earth and iron oxide with a cation ratio,
520 H. STABLEIN

(#)
L

"d

o .= II

a3
~=~

t'-
0

~o~
0 0

n
=

i
to~ ~ h

#,=

=
0 0

!D.

,.-1
H A R D FERRITES A N D PLASTOFERRITES 521

tt3

0
>

•. ~ ~.s
•~.~ ~o ..~ ~o~
so ~ o~ - ~., ..~ '7:J

>

0
~.~ 0
i ~o
0 .~

0
.'=

,~ t'¢3
0
<

c-
O

p~ o.. =
o 0
[]
522 H. ST~A3LEIN

Ba(Sr):Fe = 1:12, at least not in air. This was confirmed in tests carried out by
Bergmann (1958) who attempted to make BaFe12019 and found a lack of oxygen
in the reaction product. With BaFea2019 and, presumably, also SrFelzO19, higher
oxygen pressures are needed to avoid the formation of primary phases containing
Fe 2+. From melts richer in B a O or SrO, however, hexaferrite can be separated as
a primary phase (Kooy 1958, Goto et al. 1971). This process supplies monocrystals
in sizes of up to several millimeters. Whether high-coercivity crystals in sizes from
0.1 to 1 ~ m can also be made is not known.
Non-crystalline solid specimens of the systems BaO-Fe203 and PbO-Fe203 can
be made in a larger range of compositions by splat-cooling the melt (Kantor et al.
1973). On heating a eutectic specimen (40mol % BaO, 60tool % Fe203)
BaO.Fe203 crystallizes at 610°C and at 770°C BaO-6Fe203 also (Monteil et al.
1977). No detailed information was given on their magnetic properties, however.
In an analogous way amorphous specimens of the system SrO-Fe203 can be
prepared, too, in which SrO.6Fe203 crystals are formed at temperatures of at least
720°C (Monteil et al. 1978).
Relatively large crystals can be crystallized out of molten fluxes, e.g., from melts
containing NaFeO2 or PbO, Bi203, B203 or alkali halide/earth alkali halide
(composition, for instance, as with Arendt 1973b). These processes are used to
make crystals for scientific purposes, e.g., for studying domain configurations and
wall movements. In the past few years a number of melting processes has
become known in which technically interesting aspects play a role. They are
compiled in table 8.
According to Routil et al. (1969, 1971, 1974), Ba- and Sr-hexaferrites can be
made directly using their sulphates, i.e., from the most important minerals of
barium and strontium without it being first necessary to convert them into
carbonates. In addition to the raw materials sulphate and iron oxide only NaCO3
or other suitable additives are required. Stoichiometric hexaferrite crystals crys-
tallize from a large range of compositions; two possibilities are shown in table 8.
Typical reaction conditions are 1 hour at 1200°C in air. The formation of
strontium hexaferrite was investigated in detail and NazFe204 and 7SrO.5Fe203
were found as intermediate products. Depending on the composition of the raw
mix and temperature, the process can be operated as a straight-forward solid state
reaction or as a reaction in which a molten component is present. The iron oxide
used need not be particularly fine and an excess of A1203 and SiO2 impurities is
converted into water-soluble Na-compounds (self-cleaning). Phase separation of
the reaction product is carried out by leaching and magnetic separation. Depend-
ing on reaction conditions, the hexaferrite particles can be in millimeter-sizes.
Magnetic properties are not mentioned.
The process of Wickham (1970) is based on the normal raw material mix of iron
oxide and Ba- or Sr-carbonate which can react in a melt composed of 80 mol %
Na2SO4 and 20 mol % K2S04 at 940°C. In this composition, the sulphate mixture
has the lowest melting point of 845°C. Fine crystalline reaction products
BaFe12019 or SrFe~20~9 can therefore presumably be made although no in-
formation is furnished on this.
H A R D FERRITES A N D PLASTOFERRITES 523

S"

,-£

.~ ....,
< r.~

Eo
©
© ~ , ,.-!
'a oq ~
I °'1 l--- ( ~

0 ,.., ,-~
~'a ,%- ~!D ¢)
0
5 -~ '11-S ~C

o~
~3
o
z

2
o o
d
r.~

O~ + ~
0
=~+-~
~2 ?
© ©

2 ~ ""4' ,m

'~o ~+~.~ c~

+~-
ou~=

..0 o

0
o
0 © ¢)

¢)
E
,,~ <, O.
r~9 r.,,,9 5
524 H. STPd3LEIN

The same raw materials iron oxide and carbonates were used by Arendt (1973b),
but preferably a salt melt consisting o f 50 mol % NaCI and 50 mol % KC1. The
best magnetic properties were obtained under the reaction conditions ½h at 1000 to
1050°C in air, e.g., BaFe12019 powder with up to jHc = 340 kA/m (4.3 kOe) and
theoretical saturation polarization. The quality of the iron oxides used is of special
importance when high-coercivity hexaferrite crystals are to be produced because
the Fe 2+ content must be kept sufficiently low. In this case, an aqueous FeC13
solution is used, precipitated with a solution of 50 mol % NaOH and 50 mo! %
KOH, Ba- or Sr-carbonate is admixed followed by spray-drying which produces
an intimate mixture of the reaction couples and of the melting agent. After
further processing as described, hexaferrite powder was obtained with theoretical
saturation polarization and coercivity up to 430 kA/m (5.4 kOe) for BaFea2019 or
up to 475 kA/m (5.97 kOe) for SrFe12019.
Rather than using a salt bath, Shirk et al. (1970) used a glass melt in making
BaFe12019 crystals. They quenched a melt consisting of 26.5moi% BzO3,
40.5 mol % BaO and 33 mol % Fe203 between brass rollers as 100 ~xm thick strip
so rapidly that the glassy state was retained from which the BaFe12019 crystals
precipitated in the form of hexaferrite by heat treatment- maximum 45 % by
weight- which corresponded to the total iron oxide charged. The particle size can
be varied within wide limits by heat treatment and superparamagnetic, single, or
multi-domain behaviour generated. After annealing at 660 o~--690°C, a coercivity
jHc = 206 or 230 kA/m (2.6 or 2.9 kOe), being constant between 77 and 300 K, was
attained; after annealing at 820°C, however, a maximum coercivity of 425 kA/m
(5.35 kOe) was attained at 300 K which again, however, showed the usual tem-
perature dependence (see section 3.2).
From a similarly composed melt (21.24mo1% B203, 47.23mo1% BaO,
31.53 mol % Fe203) Laville et al. (1980) produced ca. 50 Ixm thick, amorphous
ribbons by quenching between steel rollers. After annealing at 707°C the ribbons
showed permanent magnetic characteristics a s B r = 45 mT (450 G), ~Hc = 31 kA/m
(390 Oe), and jHc = 320 kA/m (4 kOe).
A hexaferrite phase also crystallizes by annealing at least at 667°C amorphous
specimens of 30 mol % NazO, 10 mol % BaO, and 60 mol % Fe203 (= eutectic
composition of the quasibinary_ system Na20-BaFe12019) exhibiting jHc = 320 to
366 kA/m (4.0 to 4.6 kOe) (Laville et al. 1978).
Other papers deal with the amorphous state only of specimens of the system
B203-BaO(PbO)-FezO3-GeO2 (Ardelean et al. 1977, 1980, Burzo et al. 1980,
Moon et al. 1975, Syono et al. 1979, Hayashi et al. 1980). Similar results were
reported for Ba-Sr-hexaferrite crystals (Oda et al. 1982).
0.3 ixm crystals of Sr-hexaferrite were obtained by annealing a SrO-BzO3-SiOz-
Fe203 glass at 800°C for 5 h (Kanamaru et al. 1981).
To sum up it can be stated that some of the melting techniques exhibit a
number of attractive aspects because the mixing and the reaction of the raw
materials is greatly facilitated by a liquid phase present in a fairly large quantity,
cf. table 6 in section 2.2. Technically, the possibilities are probably not yet fully
exploited. A disadvantage is that by-products occur per force and that water
consumption is relatively high.
HARD FERRITES AND PLASTOFERRITES 525

2.2.4. Fluid bed and spray techniques


Using the conventional solid raw materials, the aim of these techniques is to
control the hexaferrite reaction from the start so that only single crystals of single
domain size are produced. This reduces milling requirements and the resultant
risk of contamination by abraded particles and tempering to increase coercivity
(cf. section 2.1.4) may perhaps also be dispensed with because the powder already
has a high coercivity. Two processes are compiled in table 9.

TABLE 9
Fluid-bed and spray techniques.

Manufacturing Properties of the


method Procedure reaction products Reference

Fluid bed 40 to 500 txm (micro-) Slight milling required. Sirone et al.
granulate from the 0.5 ~m size, platelet- (1972),
raw materials reacts shaped Sr-hexaferrite Fagherazzi et al.
in an air stream, particles with max. (1974),
e.g. 2 to 3 h at jHc = 445 kA/m (5.6 kOe) Giarda et al.
980 to 1050°C and Js/p = 63 mTcm3/g (1977)

Spray Milled aqueous Easy to separate Ruthner (1977)


calcining suspension of the 40-200 txm agglomer-
raw material, e.g. ates from 0.5-1 Ixm
Fe203 and Ba- sized, platelet-shaped
carbonate, sprayed crystallites with an
into hot fumace, apparent powder
Reaction conditions density of
e.g. 5 s at 1050°C 1.0 to 1.6 g/cm3.
No magnetic data.

The fluid bed process uses fine granulate which is made to react in a hot air
stream. Granulation is necessary to avoid segregation of the raw material owing to
the different behaviour of the particles if suspended individually in the air stream. In
spray calcining, however, an aqueous suspension is sprayed into the hot reaction
chamber giving a fine subdivision through the formation of droplets. Both
processes can be operated continuously (Sironi et al. 1972, Ruthner 1977, 1979).
According to Fagherazzi et al. (1974) and Giarda et al. (1977), particularly high
coercivity strontium hexaferrite powders with jHc up to 445 kA/m (5.6 kOe) are
obtained in the fluid bed process if a mix of SrCO3 and acicular a - F e O O H is
used. The latter can be precipitated from a FeSO4.7H20 solution by an alkali. By
a proper choice of the raw materials and of the reaction temperature powders are
prepared which can be processed either to compact ferrites or to plastoferrites
(Giarda et al. 1978).
Spray calcining is possible, too, by using a proper solution, see also section 2.2.2.
By feeding and spraying a solution of I mole Sr(NO3)2 and 12 mole FeC12 into a 900°C
reactor a hard ferrite powder with jHc -- 400 to 430 kA/m (ca. 5.1 to 5.4 kOe) was
obtained (DS process; Dornier 1979).
526 H. ST,XA3LEIN

2.2.5. Hot pressing and hot deformation techniques


In hot pressing (also known as pressure sintering) a porous mass is compacted
under the simultaneous action of pressure (>> 1 bar) and temperature (~ 800°C). In
this way the microstructure can be influenced as desired to a much greater extent
than is possible in a two-stage process of pressing at room temperature (section
2.1.5) followed by sintering (section 2.1.6). Owing to the effect of the pressure, the
temperature can be 100 or even several 100°C lower than in normal sintering so
that crystal growth is reduced. In many cases the process produces particles of
very low porosity and/or particularly small crystallite size.
Such a structure is also desired in hard ferrites in order to obtain at the same
time higher remanence and high coercivity (Stuijts 1970, 1973, Jonker et al. 1971).
Compaction is usually effected in a die with punches, similar to die pressing at
room temperature (section 2.1.5). The hot isostatic method could, in principle,
also be used but nothing is as yet known about its use in the manufacture of
hexaferrites.
Contrary to hot pressing, hot working ideally uses a dense body, the shape of
which is changed in the plastic stage. Press forging (upsetting) is such a possibility
in which the width is enlarged at the expense of the height of the part. An inverse
change of shape is obtained in extruding which produces a thinner and longer
part. While such forming is impossible with hexaferrites at room temperature, it
can be carried out to a limited extent at elevated temperature.
Table 10 and 11 give data from literature on hot pressing and hot forming. It
should be taken into account that in some cases both operations took place more
or less simultaneously. As the outlines show, very differently prepared starting
specimens were used, namely from raw mixes, from reacted hexaferrite powder or
from pre-oriented and pre-sintered specimens. The composition of the specimens
also varies greatly and has not been defined in all cases. The following processing
parameters are mentioned: pressing 5 seconds to 1 hour; temperature preferably
1000 to 1200°C; pressure 35 bar to 12 kbar; tool material A1203, graphite, ZrO2,
SiC, Fe.
It can be seen that the magnetic values attained vary within wide limits. This is
attributable, on the one hand, to the different conditions of the specimens used
and, on the other, to the different conditions in hot forming.
For a given specimen density, the type and extent of crystal orientation is, of
course, of decisive importance for the remanence. With the specimens previously
subjected to orientation in a magnetic field a higher remanence is generally
obtained than with non-pre-oriented specimens. Flow of the material and crystal
growth during compaction play a considerable role in the formation of the
texture. According to Von Basel (1981) oriented grain growth is the main
reason for the increase of the degree of texture with increasing pressing time.
Even die pressing of a raw mix gives anisotropic sintered bodies with a c-axes
fibre texture. Such a fibre texture seems to be even more pronounced in press
forging because the specimen is then not only shortened axially, but also widened
radially. The opposite occurs in extrusion where the specimen is elongated axially
and made smaller radially. The c-axes therefore orient themselves preferably
perpendicularly to the direction of extrusion and a c-axes ring fibre texture is
HARD FERRITES AND PLASTOFERRITES 527

formed. With such a texture the remanence perpendicular to the direction of


extrusion can at best only reach the 2/7r part (approximately 64%) of the
saturation polarization, see fig. 70(b).
The highest jHc values after hot forming are attained if high-coercivity ferrite
powder is used which was obtained by co-precipitation (section 2.2.2). Haneda et
al. (1974b) produced in this way dense specimens with iHc = 400 kA/m (5 kOe)
which corresponds to about 83% of the initial value of jHc = 480 kA/m (6 kOe) of
the powder used.
The possible reaction between the hexaferrite and the die wall is a difficulty
which these hot techniques entail. The presence of magnetite and other not
identified phases impairs the saturation polarization and the permanent magnet
characteristics and causes, particularly towards the abscissa, a buckled, e.g.
concave, course of the demagnetization curve in the second quadrant so that the
(BH)max value and the coercivities fl-/~ and jHc are adversely affected to a fairly
great extent. In such cases, these characteristics can be substantially improved by
an oxidizing annealing treatment in air, e.g. for two hours at 950°C (St~iblein
1974).

2.2.6. Rolling and extrusion techniques


Ferrite powder is usually shaped by pressing using a die and punch as described in
section 2.1.5. This technique enables practically all shapes of magnets required in
various applications to be manufactured. The typical feature of this technique is
that well defined components can thus be made. In contrast, rolling (calendering)
or extrusion produces semi-finished products of any length or width which then
has to be cut or stemmed to the desired dimensions in a further operation. While
these techniques are frequently used for plastic-bonded hard ferrites (section 4.1)
they are of a more limited importance for the compact ferrites.
Ferrite powder alone is not amenable to rolling or extruding. A binding agent
has therefore to be added to make the powder cohesive, to impart movability to
the particles and plasticity to the mix and which can be removed on sintering
without leaving any or few residuals. For this purpose, the same substances can be
used as in granulation (cf. section 2.1.2), but in a much higher concentration, e.g.
50% by volume. Some possible combinations of binding and plasticising agents
were mentioned by Schat et al. ~(1970). Strijbos (1973) investigated the mechanism
of removing carbonaceous residuals of burnt binding agents. Table 12 gives some
data from literature. A particularly important feature of these techniques is that
they enable anisometrically-shaped powder particles (acicular or platelet shape) to
be oriented by the shearing forces set up between the rolls or in the die and in this
way to make the product anisotropic. Using this method, Carlow et al. (1968)
oriented Si3N4 whiskers by extrusion with the whisker axes being arranged parallel
to the direction of extrusion. With hexaferrites the platelet shape of the particles
is utilized (cf. section 2.1.5). The hexaferrite powder used must consist of alignable
crystallites. In the "Ferriroll" process (BH)m,x values of up to 28kJ/m 3
(3.5MGOe) are said to have been attained by rolling or calendering, with
0.25 mm thick foils having been stacked until the desired thickness of the part was
reached (anonymous, 1967). For this purpose, lead hexaferrite powder with
528 H. STJid3LEIN

P
,...i

-4
o~ ~
£
0 o

~z
p
~E

~E
e~

"0

& ..~ ~..~ "~ ~ •~ -'~ ~ .=


0 0

"~~
~ ~ ~ o

o
©

~
o ~~§o o

e-

~a
HARD FERRITES AND PLASTOFERRITES 529

.-Z

D~

cr~ +
~'" ~z
t---

,.j
r, .i ~
¢'4

1 1

0 r~

~ ~ ~
-~ ~ , ~ . ~
~ . ~
ii
i.!
o

='~~
530 H. ST)i~BLEIN

t~
8 ~

C)

~~
~z ~z C)

e..

~'~ ~'~
-..~
&
;:h
•- . ~ .o ~ ~
-: ~=~

0 ,.~ ,- ,.9o
r~ ~ c ~ '~

~.~.~,~
o "~

0 r~
HARD FERRITES AND PLASTOFERRITES 531

,.A
t~
,-.A

e.,
O

tt3

e-,
e-, 0 i

O ?
,' O ,"'

"L" ,9 e.,

'~ O

~5

,A
"~ ©

~ O

e-, ;~
,"

09 e'~ t.N
e'.,
r/)

~::~ .,-.,
©

O
532 H. ST~d3LEIN

t'--

O ,...i

,;4

..=
o m ,m
~oq.

t"q o ~ O

..=

e-

+
o ..=

,= ~ '~ ~ ~ o
# i e~

"6 o
•,. ~ ~ ~ E
g
~ o'z=- 7 O " O e~

25

oe,,~

e-.

-.=
~,
:.m
H A R D FERRITES A N D PLASTOFERRITES 533

oo

O0

0 ~
~0

~ ~.~
~0~

"~ 0 o~

°~
534 H. ST~d3LEIN

Z
{-q

'6

z.. ©

~= ~ ~
O

¢'5

o 0

O c-
'6
o

~oo ~
-o
~NI

o ,-'~
.~ 0 © "D
o

~a
0
0. <

0
2
o
.o

B ~
HARD FERRITES AND PLASTOFERRITES 535

particles of a pronounced platelet shape was obviously used. The planes of the
platelets become preferably aligned parallel to the rolling plane as a result of the
shearing forces. By a similar procedure radially oriented cylindrical magnets were
prepared from Sr-ferrite and used in,stepper motors (Torii et al. 1979, 1980, Torii
1981, Saito et al. 1981). Much less anisotropy is achieved when barium ferrite powder
is used. The manufacture of strips, tubes and segment magnets with a uniform,
exactly radial, preferred direction was described by Schiller (1968) and Richter et al.
(1968b). In segment magnets (BH)max = 10.5 kJ/m 3 (1.3 M G O e ) was reached in radial
direction (Schiller et al. 1970). Not only hexaferrite powder but also a raw mix can be
used. If anisometric iron oxide particles are employed, then anisotropic magnets are
obtained under certain circumstances after single sintering for the reasons described
in section 2.2.1. In this way, St~iblein (1974) attained (BH)max values of up to 16 kJ/m 3
(2 M G O e ) in Ba- and Sr-hexaferrites.

2.2. 7. Preparation of thin layers


Layers with thicknesses of about 0.1 to 100 Ixm are of no importance at present in
permanent magnet engineering. Investigations so far have been carried out with a
view to their use as masking material in the manufacture of integrated semicon-
ductor circuits (Taylor et al. 1972) or as material for microwave and millimeter
wave applications (Glass et al. 1977, 1978), for bubble stores, and for studying the
reaction mechanism of the formation of Ba-hexaferrite (St~iblein et al. 1972,
1973a, 1973b, cf. section 2.1.3). Some data are compiled in table 13. Permanent
magnetic characteristics are not available.
In this connection, it may be pointed out that hexaferrite materials were also
studied as substrata for thin epitactical layers (Stearns et al. 1975, Glass et al.
1977, 1978). For this purpose the system (St, Ba) (Fe, A1, Ga)12019 was thoroughly
measured by Haberey et al. (1977b). It was found that Sr(Fe, A1, Ga)12019 always
crystallizes as a single phase with a magnetoplumbite lattice and furnishes wide
latitude for independently adjusting the lattice constants and the magnetic satura-
tion polarization. When barium and aluminium are present together, a miscibility
gap occurs because instead of BaO-6A1203 the compounds BaO.4.6A1203 and
BaO.6.6A1203 are present with a different structure. Later on Haberey et al.
(1980b) prepared transparent magnetic SrFesALO19 foils, 3 to 10 p~m thick, on
non-magnetic, transparent SrGaa2019 single crystals.

3. Technical properties of hard ferrites

3.1. Magnetic characteristics at room temperature; standardization

The behaviour of the flux density B and the polarization J under the action of
their own demagnetizing field - H (and possibly of an additional external oppos-
ing field) is a significant factor in the actual use of permanent magnets. It is
usually assumed that these 3 quantities run parallel or antiparallel to one another
although this is not always the case in real magnet systems. The relevant magnetic
states of the permanent magnet are then shown in the 2nd (and 4th) quadrant of
536 H. STJ~d3LEIN

the hysteresis loop, the "demagnetization curve", in the two possible represen-
tations B(H) and J(H).
The maximum demagnetization curve physically possible with the charac-
teristics Br, (BH)m~, ~-/~, jH~, /Xrec and (BpH)m~* can be described by the
saturation polarization J~ and the crystal anisotropy constant /£1 as well as the
associated anisotropy field HA under the following conditions, cf. section 1.2:
(1) The specimen consists of a pure hexaferrite phase, i.e., there are neither
pores nor foreign phases.
(2) There is no interaction between the crystallites and polarization is reversed
by coherent spin rotation (Stoner-Wohlfarth model, Stoner et al. 1948).
(3) The c-axes of all the crystallites are completely oriented in the anisotropic
specimen and distributed at random in the isotropic specimen.
Tables 14 to 16 contain references for the three above-mentioned quantities J~,
/£1 and HA. SrM? presents the most favourable combination. The optimum
limiting values listed in table 17 are based on the values given by Jahn (1968). The
associated demagnetization curves are marked in figs. 58 and 59 with "S-W". For
BaM and PbM specimens the optimum limiting values have to be modified in line
with the respective values for J~ and/£1.
The demagnetization curves of commercial magnet grades are also plotted in
figs. 58 and 59. The relevant characteristics are listed in table 18. The demag-
netization curves reach the limiting curves satisfactorily to a greater or lesser
extent for the following reasons only:
(1) The actual saturation polarization is not only correspondingly smaller owing
to the presence of pores and non-magnetic phase constituents (cf. section 2.1.6).
T A B L E 14
Saturation polarization Js of Ba-, Sr- and Pb-hexaferrites, m e a s u r e d with single crystals (S) and
commercial specimens (C).

Specimen Js in mT* at
used 0 (K) r o o m temperature Reference

~Casimir et al. (1959)


Ba-M 660 475 LStuijts et al. (1954, 1955)
B a - M (S) - 480 Smit et al. (1955)
B a - M (S) 716 -+ 50 460 _+ 12 Jahn (1968), see also Jahn et al. (1969)
B a - M (C) - 427 Hempel et al. (1965)
B a - M (C) - 422-448 St~iblein et al. (1966a)
B a - M (C) - 433 Voigt et al. (1%9)
S r - M (S) 704-+ 38 472-+ 9 Jahn (1968), see also Jahn et al. (1969)
P b - M (S) - 462-466 Voigt (1969)
P b - M (S) 618 438---4 P a u t h e n e t et al. (1959)
N6el et al. (1%0)

* I m T & 10 G

* (BpH)max, the m a x i m u m " d y n a m i c " energy product, is defined as the m a x i m u m value of the product
of Bp and H. Here, - H represents the abscissa of a point on the demagnetization curve and Bp the
flux density, obtained with the disappearance of field strength - H .
t For designation see section 1,3, page 450.
H A R D FERRITES AND PLASTOFERRITES 537

TABLE 15
Constants K1 and K2 of crystal anisotropy of Ba-, Sr- and Pb-hexaferrites. S = single crystal specimen.

KI* in kJ/m 3 at
Specimen
used 0(K) room temperature Kz/K1+ Reference

Went et al. (1952)


BaM - 310 - Stuijts et al. (1955)
BaM - 300 - Gerling et al. (1969)
BaM - 270 <5% Giron et al. (1959)
BaM (S) 440 ± 30 313 _+9 <3% Jahn (1968), see also Jahn et al. (1969)
SrM (S) 465 ---20 346 ± 7 <3% Jahn (1968), see also Jahn et al. (1969)
PbM (S) 282 220 <2% Pauthenet et al. (1959)
N6el et al. (1960)

* i kJ/m 3 ~- 10 4 erg/cm 3.
t Because of K1 >>K2 with hard ferrites it is sufficient in all practical cases to consider only the first term
K1 = K and to neglect all higher ones, as was done in sections 1.2 and 2.1.5.

T h e s a t u r a t i o n p o l a r i z a t i o n o f t h e h e x a f e r r i t e p h a s e , t o o , c a n d e c r e a s e if, f o r
example, aluminium or chromium ions are inserted on lattice sites of the iron. The
s a t u r a t i o n p o l a r i z a t i o n o f c o m m e r c i a l s p e c i m e n s is t h e r e f o r e o n l y a b o u t 9 0 % o f
t h a t o f t h e p u r e h e x a f e r r i t e p h a s e , cf. t a b l e 14.
(2) A s r e g a r d s t h e c o e r c i v i t y j H c t h e r e a r e r e l a t i v e l y g r e a t e r d i f f e r e n c e s t o t h e
above limiting values at least in commercially manufactured specimens, because
t h e m a g n e t i z a t i o n is n o t r e v e r s e d b y c o h e r e n t r o t a t i o n o f t h e s p i n s b u t b y

TABLE 16
Anisotropy field HA of Ba-, Sr- and Pb-hexaferrites at room temperature,
measured with single crystals (S) and commercial specimens (C).

Specimen HA*
used (MA/m) Reference

BaM (S) 1.35 Smit et al. (1955)


BaM (S) 1.29 Silber et al. (1967)
BaM (S) 1.27-1.38 Jahn (1968), see also Jahn et al. (1969)
BaM (S) 1.35 Dixon et al. (1970)
BaM (C) 1.35 Hempel et al. (1965)
BaM (C) 1.22-1.24 Voigt et al. (1969)
BaM 1.43 Mamalui et al. (1975)
SrM 1.54 De Bitetto (1964)
SrM (S) 1.40-1.51 Jahn (1968), see also Jahn et al. (1969)
SrM 1.63 Mamalui et al. (1975)
~Pauthenet et al. (1959)
PbM (S) 1.i0 tN6el et al. (1960)
PbM (S) 1.11-1.15 Voigt (1969)
PbM 1.31 Mamalui et al. (1975)

* 1 MA/m --" 12.57 kOe


538 H. STfid3LEIN

"C

,.-d

~q
• I"'-


©

~ ÷~
,,.-,

d,

t~

©
H A R D FERRITES AND PLASTOFERRITES 539

S!W 2 lailb
J 400
i mT .0

-I
3alla I 4
I o
200
o
/
~ ~"
.~i.
/ 0
-I gO0 -600 -400 -300 k A Im -200 -100 0
Field strength =
Fig. 58. Demagnetization curves J(H) of commercial hard ferrite grades (curves (la) to (4)) compared
with ideal coherent rotation model (Stoner-Wohlfarth curves "S-W" fitted to Js = 472mT and
/£1 = 346 k j/m3):
(la) grade KOEROX 360 (anisotropic);
(lb) grade K O E R O X 380 (anisotropic);
(2) grade K O E R O X 300 K (anisotropic);
(3a) grade KOEROX 300 (anisotropic);
(3b) grade K O E R O X 330 (anisotropic);
(3c) grade KOEROX 330 K (anisotropic);
(4) grade KOEROX 100 (isotropic).
(KOEROX is the registered trademark of Fried. Krupp GmbH, D-4300 Essen, for hard ferrites.)

.f
,oo!
mT ~

.t t" 200 ~-
.0
1 "1 Ib j
.f
f 3¢~2

-,~00 -300 kAlm -200 -100 00


Field strength -- ~-

Fig. 59. Same as fig. 58 but as B(H) graph.

nucleation of domains magnetized in opposite directions and by Bloch wall


displacement• These processes take place in sometimes considerably smaller
opposing fields than the coherent rotation (Haneda et al. 1973a, Roos et al. 1980).
(3) The c-axes of the anisotropic specimens are never completely aligned. This
means a reduction in the above-mentioned limiting values of both Br and jHc.
B r - c o s o~ decreases as the angle a between the c-axis and the magnetic field
increases (fig. 39). The dependence HA(a), however, for coherent spin rotation is
540 H. ST~BLEIN

TABLE 18
Remanence Br, maximum dynamic energy product (BvH)max, maximum static energy
product (BH)max, coercivity eric, intrinsic coercivity jHc and relative recoil permeability/zrec
of commercial hard ferrite grades*. Manufacturers' sales literature often gives the ranges
of magnetic characteristics prevailing in mass production. The values given below were
obtained from favourably sized specimens of the upper quality level. Minimum values of
standardized grades are shown in tables 20 and 22. Nominally isotropic grades can exhibit
slight anisotropy, see text on page 540.

B~ (BpH)max (BH)max BHo JHc ]abrec


Grade (mT) (kJ/m3) (kJ/m3) (kA/m) (kA/m)

Isotropic:
KOEROX** 100 230 35 9.1 151 240 1.18
Anisotropic:
KOEROX 360 400 67 30.2 175 176 1.06
KOEROX 300 375 67 25.3 185 186 1.07
KOEROX 300 K 370 78 25.3 224 225 1.06
KOEROX 330 370 88 25.3 249 250 1.06
KOEROX 330 K 360 92 24.0 263 300 1.06
KOEROX 380 400 102 30.4 260 265 1.05

* 1 mT~ 10 G; I kJ/m3 --"0.1257 MGOe ~ 1/8 MGOe; i kA/m ~ 12.57 Oe ~ 100/80e.


** KOEROX is the registered trademark of Fried. Krupp GmbH, D-4300 Essen, for hard
ferrites.

very strong, especially with small a, e.g., HA(10 °) ~ 0.67HA(0°), see curves " S W "
in figs. 73 and 74.
C o m m e r c i a l "isotropic" specimens often exhibit slight directional orientation
which derives f r o m the anisometric particle shape and the theological p h e n o m e n a
during m a n u f a c t u r e (cf. section 2.1.5). M e a s u r e d in the preferred direction, this
influence can offset the disadvantages m e n t i o n e d u n d e r (1), page 536, cf. curves 4 in
figs. 58 and 59.
T h e r e m a n e n c e can be quantitatively calculated f r o m the orientation dis-
tribution of the crystallites, which is d e t e r m i n e d by radiography, assuming that
after magnetization the spins in each crystallite lie parallel to the nearest c-axis
and there is no interaction b e t w e e n adjacent domains. Slightly t o o high Br values
were f o u n d in such a c o m p a r i s o n (Stfiblein et al. 1966a). T h e difference can be
attributed to anisotropic internal shearing (Stfiblein 1968a). If the r e m a n e n c e is
a s s u m e d as given, then u n d e r the action of an increasing opposing field changes in
magnetic state first occur, as described by the S t o n e r - W o h l f a r t h theory. Devia-
tions are o b s e r v e d with the grades f r o m H ~ - 1 4 0 k A / m (~--1.76 k O e ) shown in
figs. 58 and 59, because i n h o m o g e n e o u s magnetization processes then occur.
These set in so suddenly with the anisotropic grades that the polarity of a b o u t ~ of
the m a g n e t v o l u m e is reversed within a field c h a n g e of s o m e k A / m (several
10 Oe). Quantitative calculations or predictions are no longer possible in this case.
Figures 58 and 59 contain the demagnetization curves of only one isotropic but
several anisotropic grades (No. 4 and l(a) to 3(c)). T h e isotropic grade is easy
HARD FERRITES AND PLASTOFERRITES 541

to manufacture and the course of its demagnetization curve largely corresponds to


what is physically possible, at least between points Br and uric. This is the
important region in actual applications and thus there is no need for further
isotropic grades. The picture is different with the anisotropic grades. In principle,
it is easy to manufacture grades of high remanence and low coercivity and vice
versa (including transitions) either by appropriate selection of the manufacturing
conditions (especially the sintering conditions, cf. section 2.1.6 and fig. 45) or by
the type and amount of additives (cf. section 2.1.1). In order to obtain both
properties with values near their upper limits, all parameters relating to com-
position and manufacturing technology must be matched precisely. For example,
the highest-quality grades corresponding to the curves lb and 3c are manufac-
tured by wet pressing (cf. section 2.1.5) using strontium hexaferrite. Higher
manufacturing costs are, of course, reflected in the prices with the result that the
necessary quality criteria will not be set higher than absolutely necessary in view
of the geometric and magneto-physical circumstances and possibilities. This not
only applies for use at room temperature but for the entire temperature range
under consideration (cf. section 3.2) including all other influences (cf. sections 3.3
to 3.5).
In order to obtain the optimum magnetic values the material must be com-
pletely magnetized. This requires field strengths at least double to treble the
coercivity jHc (Stfiblein 1963). Here, the actual field strength in the material is
meant, i.e., any existing demagnetizing field has to be applied additionally. If
magnetization or remagnetization is incomplete, inner hysteresis loops with
correspondingly lower characteristic values will be passed through, cf. figs. 60(a)
to (d). It is worth noting that the innermost loops have only a narrow lancet-like
shape, even with anisotropic grades. Here, predominantly reversible mag-
netization changes take place. The rectangular shape of the outermost loop is only
reached gradually. Inner hysteresis loops of hard ferrites (and other permanent
magnet materials) were measured by Dietrich (1969).
It may be appropriate at this point to make a few remarks on national and
international material designations and classifications. Every manufacturer has, of
course, his own trademark(s), cf. table 19. The various qualities are designated by
additional numbers or letters. This may be purely continuous numeration without
any reference to magnetic values. In certain countries, however, the number has a
direct relationship to the (BH)m,x value (so far always expressed in the CGS
numerical value for the unit GOe) and a letter can represent the high-remanence
grade with R, B etc. and the high-coercivity grade with K, H etc. For example, a
ferrite grade "100" means a (practically) isotropic material with a (BH)max value
of approximately 100 × 104 GOe (= 1 MGOe ~ 8 kj/m3). Analog designations for
anisotropic materials with correspondingly higher (BH)m~x values are "250" to
"430". Moskowitz (1976) has listed the tradenames of hard ferrites and other
permanent magnet grades.
In recent work on standards for permanent magnet materials a designation
system is contemplated which is understandable internationally, is related to the
most important permanent magnet characteristics and takes into account the ISO
542 H. STABLEIN

,~7 .6 -5 -4 -3 -2 -1 0 t 2 ~ ~, _5 6 koe z
4OO
kG mT
3 300 "",
2~
2 2oo ~,

1 too ~.
I

0 0

.! -100

-2 -200

-3 -300

:~oo-4oo-4oo-3'~o -240 -160 -80 0 80 160 240 320 400 kAIm 5FrO
400

Fig. 60(a)

-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 kOe6
,~00
kG mT
300

2 200 .~

1 too g-

, -100

-2 -200

-3 ~ / ill' -300
_ jj F
~ J
-400
-40o-4o~ 4'20-22o -~so :eb o ob 160 240 320 kAIm 480

Fig. 60(b)

Fig. 60. Outer and inner hystereses loops (hysteresis family) and magnetization curve of various
commercial hard ferrite grades: (a) grade KOEROX 100, (b) grade KOEROX 360, (c) grade KOEROX
330 K, (d) grade KOEROX 380.
H A R D FERRITES AND PLASTOFERRITES 543

-7 -6 -5 -Z, -3 -2 -1 0 1 2 3 4 5 6 kOe 7
4 4OO
kG mT
3 300

lj
2 2oo ~
I ~oo ~.

-I - 100

-2
J -200

-3 -300

-400
-56o -4'8o -,bo- - 33o -24o -/6o -8o 0
i

8O /6o 2'~o 3~ ~bo k;a~.,560


Fig. 60(c)

-7 -6 -5 -,~ -3 -2 -1 0 1 2 3 4 5 6 kOe 7
4 dO0
/#f
kG mT
3 300 "~

2
j'/l tf~-- 200
.o

I I' /~ ......
I I /" 1oo g.
0
/ - J- ' Field strength H
-I 100
L'
-2 200

-300

-400
-560 -480 -400 -320 - 2Z~O -160 -80 0 80 160 240 320 400 kAIm 560

Fig. 60(d)
TABLE 19
Standard designations and trademarks of sintered hard ferrites from various
countries. Some trademarks are of historical interest only.

Country Standard designation Trademarks

Federal Rep. Hartferrit Cekasyd


Germany DF
Dralodur
Ferram
KOEROX
Ox
Oxidur
Oxit
Oxyd
Roboxid
SAF-Ferrit
Unox
German Democ. Rep. Maniperm
France Oxonyte
Spinal
Spinalor
Great Britain Caslox
Feroba
Magnadur
Italy Oximag
Sinterox
Japan MPA FB
MPB Ferrinet
FXD
SSR
YBM
Netherlands Ferroxdure (FXD)
Spain Ferdurite
Ferribarita
CSSR Durox
USSR Bi; BA -
USA Ceramic Arnox
Ceramagnet
Cromag
Duramax
F
Ferbalite
Ferrimag
Genox
Indox
M
Magnite
Mi-T-Mag
MO
Moldite
Westro

Korea KF
(South)

544
HARD FERRITES AND PLASTOFERRITES 545

(International Standardization Organization) recommendation R 1000 on the


introduction of the SI system (Syst6me Internationale d'Unit6s). As an example,
table 20 shows the hard ferrite grades covered by D I N standard 17 410 of May
1977 in the Federal Republic of Germany. These are characterized by a pair of
numbers. The first n u m b e r indicates the minimum (BH)max value in kJ/m 3, the
second n u m b e r the minimum value of the coercivity ~Hc/lO in kA/m. The
selection of these two values allows for the fact that the (BH)ma~ value is the most
important characteristic for static applications where the state of the p e r m a n e n t
magnet is constant in terms of space and time, whereas for dynamic applications
the coercivity jHc represents a measure of and a reference to the stability and the
load capacity of the magnet in alternating counter fields.
Proceeding from this national work on standards the International Elec-
trotechnical Commission (IEC) has recently started work on magnetic materials in
its technical committee 68. Initially, a classification of all magnetic materials was
drawn up (IEC publication 404-1, 1979) containing the typical values or value
ranges of the characteristics of interest. Table 21 contains the values given for
hard ferrites. T h e lower limits of the value ranges largely match the minimum
values of D I N 17 410, table 20. The p e r m a n e n t magnet materials were specified in
another I E C document (68 C O 24, 1980) and the values in table 20 largely
adopted. Only the grade 26/18 was added, cf. table 22, thus showing that
standards of quality are in good agreement worldwide*.
T h e characteristic values for p e r m a n e n t magnets are usually determined by
passing "slowly", quasi-statically through the demagnetization curve or h y s t e r e s i s
loop. However, the polarization J obtained with a constant field strength is
actually dependent on time and one therefore speaks of "thermal after-effect". In
a n u m b e r of p e r m a n e n t magnet materials the polarization decreases according to
the formula AJ = S log t + C where S is proportional to the absolute temperature,
C is a constant and t the measuring period. With anisotropic specimens, J is
highly dependent on the working point and the direction of m e a s u r e m e n t accord-
ing to Dietrich (1970). A J is much smaller above the knee of the demagnetization
curve than below it, i.e. on the steep gradient of the hysteresis loop, and much
smaller a J values are observed in the jHc range if m e a s u r e m e n t s are m a d e
perpendicular to the preferred direction.
A difference must be made between the thermal after-effect and the natural
stability, which means the change in the state of magnetization of a p e r m a n e n t
magnet or p e r m a n e n t magnet system in its own demagnetizing magnetic field
following prior magnetization (and, possibly, stabilization). The results obtained
by various authors are compiled in table 23. As the specimens used were not

* National standards or standard-like agreements on hard ferrites in other countries are:


-Japanese Industrial Standard (JIS) C 2502 (1975) "Materials for Permanent Magnet";
-Magnetic Materials Producers Association (MMPA) Standard No. 0100-75, Evanston/Ill., USA,
"Standard Specification for Permanent Magnet Materials";
-Fachbereichstandard TGL 16541/01 (1976), GDR, "Hartmagnetische Werkstoffe, Oxidische Sin-
terwerkstoffe".
- GOST 24 063-80, USSR, "Magnetically hard ferrites" (contains only Ba-hexaferrites).
/
546 H. STJ~d3LEIN

© 0"5
r'-,.

~3 2
t"Q
t'q

o
Z
2
0

e-,

F_,

E
C

~ooll

"~ ~222
N NNNNNN
HARD FERRITES AND PLASTOFERRITES 547

2
~.oo.
8

~D

~.. o ~

e', ,~
I
¢'.1

~<
,d

~<
t'q
Cq

p.z

Cal~
k;

qZ:
rqeq (D

2
O

_= E
.~

~-=
<
548 H. ST~,BLEIN

©
oo :2
03
t"-

© 2
0'3

r,.)

tt3

C)
t'q

~-.x k

"G

e-,
£
~8

,,.-,
~D
"'G

e..,
HARD FERRITES AND PLASTOFERRITES 549

E -~ .r"

e-,
e-,
+1 +1
w. ,~.
ii
q

w. ~. ce m.
%
O
e~
e~
e-,

×
(..q

,.tZ

tt3

e'~ q3

II

o
,<

~D
eq ~ (-.q
E
©
t"q
©

0
£
cq ~ cq
e-

~h
tt'5 0
¢q ¢q
©

.&.~ o o
O

~D

O
c,.)
Z
550 H. STJ~3LEIN

adequately characterized in all cases, it is not possible to make a quantitative


comparison. Qualitatively, it can, however, be concluded that the natural stability
increases as coercivity increases. Kronenberg et al. (1960) established a logarith-
mic dependence of the flux density on time. According to Dietrich (1970) the
instability disappears if the working point is close to remanence or coercivity 8He
and has its maximum value when the working point is close to the (BH)max point. The
problem of natural stability was dealt with theoretically by N6el (1951).

3.2. Influence of temperature on magnetic properties

The magnetic state and magnetic properties of a magnet below its Curie tem-
perature Tc (~450°C for hard ferrites) are, for various reasons, dependent on its
temperature history and the temperature at which they are measured.
The temperature history includes the sintering treatment where the density and
the grain size of the magnet and thus its remanence and coercivity, inter alia, are
adjusted by control of the sintering temperature and period, cf. section 2.1.6. In
contrast to metallic permanent magnets hexaferrites do not require any further
heat treatment with the result that only the structure produced by sintering
determines the magnetic properties at the service temperature. Irreversible
changes in the structure and thus in the magnetic properties would only occur if
the magnet were again reheated to sintering temperature, i.e., to around 1200°C.
Hexaferrites are thus the permanent magnets with by far the most stable struc-
ture.
The temperature history can also influence the directional distribution of the
domains. After magnetization the polarizations of each crystallite should be
oriented parallel to the direction of the adjacent local c-axis, i.e., 0 ° to a maximum
of 90 ° to the magnetizing field. This occurs when the nucleation field strength is
n o t exceeded anywhere. If, however, it is exceeded, polarization changes irrever-
sibly to the opposite direction and angles exceeding 90 ° are observed. With
commercial anisotropic magnet grades the nucleation field strength is practically
identical to the coercivity jHc since the hysteresis loop is more or less rectangular, cf.
figs. 58 and 60. Irreversible changes in the directional distribution of the spins and
thus irreversible changes in the flux emitted from the permanent magnet therefore
occur when the (local) coercivity is no longer greater than the (local) demagnetization
field. With hexaferrites the coercivity drops as the temperature decreases, in this case
with the danger of irreversible magnetization changes occurring. The term "irrever-
sible" refers only to the instantaneous change which cannot be reversed by
temperature changes but by re-magnetization only.
With the reversible magnetization changes there is a clear relationship between
the magnetic characteristic value o r magnetic state and temperature. For a not too
large temperature range A T the change ABr in the remanence and AjHc in the
coercivity can be regarded as linear and described by a temperature coefficient c~:

1 ABr 1 AjH¢
a(Br) = Br AT ' and a(sHc) = iHc A T '
H A R D FERRITES AND PLASTOFERRITES 551

where the initial values Br and ~Hc are usually related to room temperature. In the
following the temperature dependences of Br and of jHc and then the tem-
perature dependence of the entire demagnetization curve will be examined.
Figure 61 shows the saturation polarizations Js of BaM, SrM and PbM as well as the
remanences Br of anisotropic and isotropic commercial barium ferrites as a function
of temperature. Further similar data are given by Rathenau et al. (1952) (BaM),
Richter (1962) and Schiller (1965) (anisotropic Ba-ferrite) and Fahlenbrach (1963)
(isotropic Ba-ferrite). All curves are straight-lined in a wide temperature range and
with a gradient which corresponds to a temperature coefficient of about a(Js)=
(B 0 = - 0 . 2 % / K near room temperature. Both temperature coefficients are equal
when the polarization of the material is rigid, i.e., no polarization reversals take
place. If and at what counter field H v this behaviour is present can be seen from the
fact that the gradient of the demagnetization curve from the remanence up to this
field Hp lies at or at least near the theoretical value of /~e~ (cf. table 17).
The temperature dependences of B~ and Js are then determined by the same
mechanism.
In order to assess the jH~ temperature dependence it is useful, as in section 3.1,
to first observe the behaviour with coherent magnetization reversal (Stoner-
Wohlfarth theory). The temperature dependence of the coercivity jH~ is then

600
3 2
mT

-<..
.£ 200
2

i
400
i
,K 6o0 ~3
r

- 200 0 200 °C 400


Temperature

Fig. 61. Saturation polarization Js and remanence Br vs. temperature. Curve (1): Js of BaM (Casimir et
al. 1959); (2): Js of SrM (Jahn 1968); (3): Ys of PbM (Pauthenet et al. 1959); (4): Br of anisotropic BaM
specimen (Ireland 1959); (5): Br of isotropic BaM specimen (Hennig 1966); (6): Br of isotropic BaM
specimen (Went et al. 1952).
552 H. STPd3LEIN

given by the temperature dependence of the values H A = 2K/Js for ideally


oriented specimens and HA = 0.48 (2K/J,) for isotropic ones. The values found by
various authors for BaM and SrM are shown in fig. 62 (curve (a), (b) and (c) of A).
Here, the possible influence of a demagnetizating field owing to the platelet-like
shape of the ferrite particles (cf. section 2.1.6) has been disregarded. If this
influence is taken into consideration, however, and if the crystallites are idealized
as extended platelets (fig. 39), then the effective anisotropy field strength is given
by the equation H A = 2K/Js-Js/I, Zo for oriented specimens and HA = 0.48
(2K/Js-Js/I, zo) for isotropic specimens which are shown in fig. 62 under B.
Figures 63 and 64 contain measured temperature dependences of jHc for
anisotropic and isotropic specimens, respectively. The anisotropic specimens fall
considerably short of the theoretical values of fig. 62, the best attaining about
40%. With the isotropic specimens, on the other hand, agreement with the
theoretical values is better. This applies especially to specimens which were
manufactured by co-precipitation (section 2.2.2) and which are listed here for
comparative purposes. The curve measured by Mee et al. (1963) on relatively
loosely pressed powder largely corresponds to the one theoretically predicted for
curve B(c) in fig. 62. Compact specimens, however, only tend to approach the
theoretical values at elevated temperatures and then only to a poor degree. This is

1600 20

kA kOe

1200
[anisotropic I --
2K '
I
"~ ( ~ Js Is, I
~~, 800 ..I " i tO

I
I

.oo j®' '

Ii
ti
9 200 z;O0 600 ~ 800K
0 ~ i ~ t I I I I I

-200 0 200 400 - *C


C
Temperature
Fig. 62. Calculated anisotropy field HA for coherent spin rotation vs. temperature for anisotropic and
isotropic specimens. Curves (A) were calculated without the influence of a demagnetizing field,
whereas curves (B) take this influence into account. (a) SrM (Jahn 1968), (b) BaM (Casimir et al. 1959),
(c) BaM (Rathenau et al. 1952, Rathenau 1953).
HARD FERRITES AND PLASTOFERRITES 553

0 200 400 600K


i i i
50O i i i i
6
k_~A 1
m kOe
~ 400 5

300
7/ 4

o 3
,~ 200
2

i 0
-200 0 200 400 °C
Temperature =
Fig. 63. Measured intrinsic coercivity jHc of anisotropic hexaferrite specimens vs. temperature. Curve
(1): SrM (Jahn 1968); (2): BaM (Sixtus et al. 1956). a Ip and a I refer to the same specimen, measured
parallel and perpendicular, rasp., to the preferred direction; (3): Hennig (1966), (a) SrM, (b) BaM, high
]Hc grade, (c, d) BaM, high Br grade; (4): K O E R O X 330 K (see also fig. 65(b)); (5): K O E R O X 360
(see also fig. 65(d)).

0 200 400 600 K

8
600
kA ",w 7
m
kOe
5O0 " 6

I ZOO f ~
2b
5

4
jHc 300
/ / \"il 3
20O
2
tO0 l

0 0
-200 -I00 0 I00 200 300 ZOO 500oC
Temperature
Fig. 64. Measured intrinsic coercivity ]Hc of isotropic hexaferrite specimens vs. temperature, com-
pared to calculated anisotropy field HA.
Curve (la) HA = 0.48 (2K/J~ - Js/P-0), same as curve (Bc, isotropic) of fig. 62.
Curve (lb) HA = 0.96K/J~, same as curve (Ac, isotropic) of fig. 62.
Curve (2a) 30% by v o l . - specimen SrM (Mee et al. 1963).
Curve (2b) 30% by v o l . - specimen BaM (Mee et al. 1963).
Curves (3) B a M (Sixtus et al. 1956).
Curves (4) B a M (Went et al. 1952).
Curve (5) BaM (Rathenau et al. 1952, Rathenau 1953).
Curve (6) BaM (Fahlenbrach 1963).
Curve (7) BaM (Hennig 1966).
Curve (8) K O E R O X 100, see also fig. 65(a).
554 H. STABLEIN

attributable to the formation and displacement of Bloch walls. The ever increasing
deviation towards lower temperatures is related to the fact that the critical particle
size decreases with the temperature (Rathenau et al. 1952, Rathenau 1953, Sixtus
et al. 1956), i.e., an increasingly larger proportion of the crystallites tends towards
spontaneous Bloch wall formation.
As can be seen from figs. 63 and 64, the curve gradients, e.g. in the range from
0 to 100°C, vary considerably, even with roughly the same jHc values at room
temperature. For the commercial specimens listed there (jHc> 100kA/m),
changes AIHdAT of 0.4 to 1.1 kA/m per K (5 to 1 3 . 8 0 e per K) are found without
these values exhibiting a distinct variation with the coercivity. The temperature
coefficients of these specimens lie in the range of a(jH~) = 0.15 to 0.50%/K where
c~(jH~) tends to increase when jH~ drops, it must be assumed that the temperature
variation of jHc depends on the precise constitution of the structure and thus on
the details of the manufacturing process. However, this does not appear to have
been studied in detail.
The K O E R O X materials shown in fig. 62 exhibit in the range from - 4 5 to
+85°C average AjHdAT values of 0.67kA/m per K (8.3Oe per K) for the
isotropic specimen K O E R O X 100 and 0.79 to 0.94 kA/m per K (9.9 to 1 1 . 7 0 e per
K) for the anisotropic grades, i.e., a change fairly independent of coercivity. The
average temperature coefficients for the same temperature range are a(jHc)=
0.27%/K for the isotropic grade and 0.28 to 0.48%/K for the anisotropic grades
where the approximation a(jH~)-1/jH~ applies. In general, the temperature
coefficient for these materials can be approximated from t~(jHc)-- 86/jHc in % / K
(jH~ in kA/m). Here, the factor 86 is 100 times the average of the above-
mentioned range from 0.79 to 0.94 kA/m per K. (In CGS units: a ( j H c ) = 1.08/jHc
in % K with jHo in kOe and an average value of 1 0 . 8 0 e / K . )
In practice, a permanent magnet is usually neither in the Br nor jHc state, but in
between. In order to evaluate its temperature behaviour the change of the
demagnetization curve with temperature must therefore be studied. Figure 65
contains the demagnetization curves of some commercial hard ferrite grades
- m e a s u r e d at temperatures of -45, +25 and +85°C. Described ideally, the demag-
/ n e t i z a t i o n curves always consist of two almost straight sections, a Br and a jHc
section, which are linked by a transition section ("knee"). The gradients of both
sections hardly depend on temperature. As regards the gradient of the Br sections
/Xrec~ 1 is always found for anisotropic grades in accordance with table 17 and even
for the isotropic grade there is only a change in gradient of about +2% or - 2 % for the
temperatures - 4 5 and +85°C, respectively, compared with the value at room
temperature. Therefore, approximation curves for other temperatures can be simply
plotted from the demagnetization curve for a certain temperature and from the
temperature behaviour of Br and jHc.
The behaviour of the magnetic flux as a function of temperature can be
described very simply in an idealized diagram (cf. Schwabe 1957). Figure 66(a)
contains two idealized demagnetization curves B(H) of an anisotropic hard ferrite
for the temperatures T = room temperature and T ' < T. All domains are uni-
formly oriented on the Br curve sections P4Br and P;B'r, as shown at F1. In the iHc
H A R D FERRITES AND PLASTOFERRITES 555

B"H #)kJ/m3
a 16 8
400
mT
300 t
Qa

2OO rn
i11

MGOe 100
-5kOe 2 ~--~ - ----=-3~ - -I
0
-400kFA
n._
_
.A -300 - ~ -100~
- 100

--200

+~~25oC -45°C
~ 85oC

B'H in kJlm3
40 32 24 16
ZOO

300 !

B.H/in
MGOe ~ - 2 200
5
4 // I00
3 /
-5kOe -4 2 - - 3 " -1
0
- 400k~A
k - 300 / - 200 . -100 /
Field strength I00

_-200

+85°C

Fig. 65. Demagnetization curves B(H) of some commercial hard ferrite grades, measured at -45, +25,
and +85°C. Grade KOEROX, (a) 100, (b) 330 K, (c) 300 K, (d) 360, (e) 380.
556 H. ST.~i~BLEIN

B. H in k Jim 3
C ~00

300
B-H in
MGOe / 200 ~-

3 ~ 7 100
-5kOe -~ 2 -~3 /

,oo _ -
0

tO0
Field strength " I I /
_ -200
-~5°C
+25. C
J +85~C

B.H inkJ/m 3
d 4OO

3O0

B'H in
200

I00
-5kOe
-i ~ -3 ~
/, 0
-Z OOkA -2,
m
-I00
Field strength
-45*6" 200

/ +25°C
////~85"C

Fig. 65 (continued).
H A R D F E R R I T E S A N D PLASTOFERR1TES 557

B. H in kJ/rn 3 500
e ~0 32 2~ ~, mT
400

300
T
b
B.H in 200
MGOe
5~
4 lO0
3~ tu
-5kOe -4 -2 -l
, 0
-400 k_AA - 200 -I00
m - 100
Field strength

-200

,25°C
÷85"C
Fig. 65 (continued).

state the polarity of 50% by volume of the magnet is reversed, which is shown
diagrammatically in F3. A continuous transition, e.g. represented by F2, takes
place on the vertical jHc curve sections P4jgc and P~jH'c. The absence or
occurrence of irreversible losses with a temperature cycle T ~ T ' ~ T obviously
depends on whether the load line corresponding to the actual working point lies
above or below $2.
With a load line lying above $2, e.g. at S1, the changes in B and H of the
magnet are completely reversible. As is shown in sections (b) and (c) of the
diagram, the flux density changes from B I ~ B ~ B 1 and the associated field
strength from H1 ~ H~ ~/-/1.
With a load line lying below $2 but still above $4, e.g. at $3, the reactions during
cooling are still reversible up to the intermediate temperature where the knee PK
of the associated demagnetization curve lies on $3. At the intermediate tem-
perature flux density and field strength in the magnet pass through extreme values
BK and HK respectively. Further cooling results in the state P; with the values B ;
and H ; owing to the irreversible change in polarity of some of the domains. On
reheating to T there is no change in the domain distribution established in this
way. By plotting the point D this is expressed graphically in such a way that the
section ratios are P4D/P4 jHc = P;P;/P~ j H ' c and the transition from D to P~ lies on
$3, where the gradient of DP~ (just like that of P4B,) is equal to # .... On heating,
both B and H drop to B~ and H~ respectively. The total changes for T ~ T ' ~ T
are thus characterized by B 3 ~ B K ~ B ~ B ' ~ (part b) and H 3 ~ H w ~ H ; ~ H ~
(part c).
558 H. STJi,B L E I N

B~

B,
B,
B,

F, B,,
B5

-H

lq-
F~

F,
~.ure

=/./It

Fig. 66. Reversible and irreversible changes of flux density (b) and field strength (c) of an oriented
hard ferrite, shown with idealized demagnetization curves (a) of fig. 65(d) for T = 25°C and T ' =
-45°C. Load line examples $1, $3, $5 cause 3 different kinds of demagnetization behaviour of a fully
magnetized specimen on temperature cycling T ~ T ' ~ T, divided by the limiting cases S: and $4.
Inserts F1, F2, F3 show schematically fully magnetized, partially demagnetized, and completely
demagnetized state, respectively.

At a load line below $4, e.g. at $5, domains with reversed magnetization are
already present in the initial state P5 whose volume fraction is numerically equal
to the section ratio P4Ps/2 P4jHc. By cooling down from T ~ T' this fraction
increases irreversibly to P~P~/2 P~ jH" causing flux density and field strength in the
magnet to drop to B~ and H~ respectively. After reverting to the initial tem-
perature the state P~ is resumed with B~ and H~. Point P~ can be determined via
point E with P4E/P4 jHc = P~P~/P~jHc analogue to point D. The total changes for
T ~ T ' ~ T are thus given by B s ~ B ~ B ~ (part (b)) and H s ~ H ~ H ~ (part (c)).
The position of point P~ can be found in the same way as that of P~ and Pg.
However, the position of points P2, P~, P~ and P~ can also be determined by a
different approach. As the respective initial states P~, P;, P; and P; had the same
field strength and their associated flux densities dropped by almost the same
extent on heating, the states after cooling must also have the same field strength.
HARD FERRITES AND PLASTOFERRITES 559

The position of P2 and thus the desired field Strength are derived directly from the
load line $2 passing P~.
The changes shown in parts (b) and (c) run reversibly proportional to a (Br) on
the sections marked with double arrows whereas the changes on the sections
marked with single arrows are irreversibly proportional to a(iHc).
The ratios shown idealized and schematically in fig. 66 must be modified for the
actual conditions. Firstly, the knee of the demagnetization curve is in practice
curved, cf. fig. 65(a) to (e). As a result the curve at BK and HK (fig. 66(b) and (c))
runs into a flat peak. Secondly, the gradients of the 1He sections are finite and the
associated field strengths do not therefore coincide exactly. Thirdly, the demag-
netization conditions in terms of magnitude and direction are not generally
constant throughout the entire magnet volume with the result that various
shearings are assigned to the individual volume elements. For example, Schwabe
(1957) found that the most unfavourable points of a particular hard ferrite
loudspeaker system are subject to a demagnetization field which is twice as strong
as the mean field strength. Finally, inhomogeneous material properties can be
caused by inhomogeneous manufacturing conditions (e.g. in the degree of orien-
tation and in the grain size distribution), i.e., in this case a family of demag-
netization curves must be expected. If, as a result, inhomogeneous demag-
netization conditions and/or inhomogeneous material properties occur, the tem-
perature behaviour of a permanent magnet system can be qualitatively estimated
satisfactorily only to a greater or lesser extent.
Table 24 and fig. 67 show the influence of the working point, the coercivity jHc
and the maximum cooling temperature on the extent of the irreversible losses
with anisotropic hard ferrites. As expected, the irreversible losses increase as
working point, coercivity and temperature decreases.
In practice, the described temperature behaviour can thus be allowed for by the
choice of material and design.
According to the above, irreversible losses may only occur when cooling takes

T A B L E 24
Irreversible loss of flux density of oriented
hard ferrites with intrinsic coercivities of
jHc = 170 to 1 9 0 k A / m (2.13 to 2.38kOe)
and with different length to diameter ratios
l/d after having been cycled between 20°C
and various lower temperatures (Gershov
1963).

Irreversible loss, %
l/d -20°C -407C -77°C

0.18 19.1 29.0 39.7


0.39 13.3 24.0 35.8
0.73 10.1 18.5 28.8
1.0 0.0 0.0 1.7
1.3 0.0 0.0 0.9
560 H. ST~d3LEIN

1.5 2 25 kOe 3.5

</
1 u
O

%
+20°C
tO

20-

30.
t~
-20°C~ y
oJ
~0-

50-
-77°C L
6C
I00 150 2O0 2so ~A/~ 3oo
D

Intrinsic coercivityjH c at 20°C


Fig. 67. Irreversible loss of flux density after cycling between 20°C and indicated temperature of
oriented hard ferrites having a length to diameter ratio of lid
= 0.18 vs. intrinsic coercivity jHc at 20°C
(Gershov 1963).

place after magnetization but not during heating up to 400°C, the latter being
true, independent of the working point (Tenzer 1957, Assayag 1963, Gershov
1963). Irreversible losses during cooling can be avoided or at least reduced by
selecting a material with a sufficiently high coercivity jHc and/or by employing a
configuration with a sufficiently long magnet so that at the minimum temperature
even those unit volumes with the strongest demagnetization are not influenced
irreversibly. If these measures are not possible or inadequate, the loss expected
later in service can be anticipated. With ferrites it is sufficient to cool the magnet
only once to the lowest expected temperature. In principle, it would also be
possible to anticipate the losses by partial demagnetization in a magnetic field but
no information is available in literature on this point.
The irreversible temperature variation is determined by the material and cannot
be anticipated but merely influenced by the design. As is usual in permanent
magnet technology, the negative temperature variation of the useful flux can be
reduced or made positive, also with hard ferrites, using a shunt placed parallel to
the useful flux and made of a material with a highly temperature-dependent
polarization. The latter is, for example, required with eddy-current systems
(speedometers, brake systems for electricity meters) in order to compensate also
for the temperature dependence of the electrical resistance of the eddy-current
material, e.g., c~(A1)= +0.46%/K and c~(Cu)= +0.43%/K. Such materials ("com-
pensation materials") have been known for some time (St~iblein 1934). Common
compensation materials are, for instance, Fe-30%Ni alloys. Their status and
development potential are described, inter alia, by Schiller (1965) and Fahlenbach
(1972). Owing to the relatively high temperature coefficient a(Br) of the hard
ferrites, however, the cross section of the shunt must be larger than with the
metallic materials with their 10 times smaller a(Br) value. This means more
HARD FERRITESAND PLASTOFERRITES 561

compensation material as well as a lower useful flux level of the compensated


system. In fact, hard ferrite magnets have hitherto scarcely been used for
applications where the temperature coefficient of the useful flux has to disappear
or be positive. In this connection it must be noted that the temperature variation
of the air gap flux of a system is determined not only by the temperature variation
of the permanent magnet but also by that of the iron conductors. Their (smaller)
temperature coefficient plays an important role, especially with flux densities of
above about 1T (Fahlenbach 1963). However, the useful flux is drastically
reduced with this procedure so that it is not suitable for general use.

3.3. Influence of deviation from the preferred axis, of mechanical stress, and of
neutron irradiation on magnetic properties

Influence of deviation from the preferred axis


In anisotropic specimens the magnetic characteristics are, owing to the alignment
of the c-axes, generally more or less strongly dependent on angle a between the
preferred direction of the specimen and the direction of measurement or use. The
degree of this dependence is therefore significant for the application of the
magnets. In addition, the angular dependence permits certain conclusions to be
drawn on the magnetic state and the mechanism of polarization reversal. This
dependence presupposes that the magnetizing field and the direction of
measurement are parallel to each other. On the other hand, fig. 52 shows the case
where the direction of measurement of the magnetic flux was held constantly
parallel or vertical to the preferred direction and merely the direction of the
magnetizing field changed.
In exactly isotropic specimens, of course, the characteristics show no angular
dependence. It should be pointed out, however, that nominally "isotropic"
specimens may be somewhat anisotropic owing to the flow conditions in powder
compaction, cf. section 2.1.5. Depending on the details of the manufacturing
process, the magnetic fluxes of the products at present industrially manufactured
can differ by up to 20% from each other parallel or perpendicular to the direction
of pressing, cf. fig. 46. Differences also obtain in the manufacture of plastic- or
rubber-bonded hard ferrites in extruding, calendering or die-pressing, cf. section
4. The angular dependences discussed in the following also apply, albeit to a
lesser extent, for such "isotropic" specimens.
Figure 68 shows the angular dependence of the remanence Br(a) for some
specimens sintered at various temperatures (St~iblein et al. 1966a). The remanence
of the pressed state, which largely corresponds to that of the 1100°C specimen, is
not shown (Stfiblein 1965). With complete alignment remanence has to run
B r - cos a. Although the alignment is incomplete, Br roughly follows this function
for sufficiently small angles because cos a changes only slightly in this area.
Incomplete alignment is most noticeable in the area of a--~ 90°. As more and
more disoriented crystals disappear as the sintering temperature rises, the cos a
curve is increasingly approached. Joksch (1964) recorded the same angular
dependences for some commercial grades.
562 H. ST.Sd3LEIN

~ q250°C
:1350oC
COS, 0 ~ ,

00o 10° 20 ° 30 ° ,~0° 50 ° 60 ° 70 ° 80 ° 90 °


Angle ct to preferred axis

Fig. 68. Measured relative remanences B~ vs. angle c~ to preferred axis for anisotropic BaM
specimens, sintered at various temperatures (St~iblein et al. 1966a). 1100 ° specimen: Br = 330mT;
jH~ = 246 kA/m. 1250 °` specimen: Br = 395 mT; jH~ = 142 kA/m. 1350 ° specimen: B, = 405 mT; jH~ =
41 kA/m.

If the spatial orientation distribution of the c-axes is known, B,(a) can be


calculated from it, assuming that the polarization of each crystallite lies parallel to
the nearest c-axis after magnetization. The quantitative calculation of the three
specimens in fig. 68 showed good agreement, cf. fig. 69. Slight deviations between
calculation and experiment point to anisotropic internal shearing owing to non-
spherical pores.
The influence of the texture sharpness on the remanence was calculated for
some model arrangements of the c-axes (St~iblein 1966). With a "normal fibre

- ~ I . . i
LO ~ smtenng terr )erature

O~
(J
¢p
• ................. 1 ...............
. . . . . .
a,,gnment . . . . . . . . . . . . . .

¢.
OJ

Q: >1100°C
-1250oc
>1350°6,
0..0; -- a cu a i from (oo.aj

,o 30 ° 60 ° 90 o

Deviation from preferred axis


Fig. 69. Measured relative remanences Br vs. angle o~ to preferred axis for the same specimens as in
fig. 68, compared to values calculated from c-axis densities.
HARD FERRITES AND PLASTOFERRITES 563

texture", as arises in normal powder pressing in the magnetic orienting field


(section 2.1.5) the curves shown in fig. 70(a) are obtained. The frequency dis-
tribution f ( a ) of the c-axes is assumed as f ( a ) ~ cos 2i a. i = 0 means the isotropic
state, rising i the growing degree of orientation. By analogy, Br can be calculated
for the "ring fibre texture" in which the c-axes are preferably parallel to one
plane, but within this plane arbitrarily arranged, cf. fig. 70(b). In this f ( a ) - sin 2i a
was applied. The actual relationships can be described reasonably well by these
model functions, but m o r e precise results are obtained by using a series
development according to Legendre polynomials (St/iblein et al. 1966b).
T h e r e are various methods of radiographically determining the function f ( a )
(St/iblein et al. 1966b, 1971, Stickforth 1975, Willbrand et al. 1975). Frei et al.
(1959) and Shtrikman et al. (1960) described the determination of f ( a ) from Br
measurements. T h e accuracy is unsatisfactory, however, as in well aligned speci-
mens the r e m a n e n c e only reacts weakly to changes in the degree of alignment, cf.
also fig. 70(a). The angular dependence of the ferromagnetic resonance, too, can be
used to determine f ( a ) (Hempel et al. 1964). If the entire or the precise function

E O.5 • •

00o I0 o 20 ° 30 ° 40 ° 50 ° 60 ° 70 ° 80 ° 90 °
A n g l e c~ to p r e f e r r e d axis

~.o
b
o

O5

:_-5 /
o
0o 10o 20 ° 30 ° 40 ° 50 ° 60 ° 70 o 80 ° 90 °
A n g l e c~to p r e f e r r e d a x i s
Fig. 70. Relative r e m a n e n c e s Br vs. angle c~ to preferred axis calculated for two models (Stfiblein
1966): c-axis distributed preferentially along a preferred direction (a), or along a preferred plane (b).
Increasing parameter i means increasing sharpness of orientation, see text.
564 H. ST]~BLEIN

f(a) is not required but a qualitative measure of the orientational order only, the
value of f = ( p - p 0 ) / ( 1 - p 0 ) is used sometimes (Lotgering 1959). Here, p =
I(OOl)/E I(hkl), i.e, p is the ratio of the sum of all (00/) X-ray intensities and the
sum of all (hkl) intensities (including 001) of the specimen considered, and p0 is
the analogous expression for a random (isotropic) specimen, f-values range from 0
(isotropic) to 1 (completely aligned).
The angular dependence of the (BH)~, value of some commercial specimens is
shown in fig. 71. In all specimens it is very well reflected by a cos 2 a curve. This is
to be expected because of (BH)ma~ ~ B 2, as long as the demagnetization curve
B(H) in the second quadrant shows a pronounced knee, /x0BH~ is greater than
B~/2 and the relatively weak #~eja) dependence is neglected.

Ioo
%

t~

so
Q

0
0o 10° 20 ° 30 ° 40 ° 50 ° 60 ° 70 o 80 ° 90 °

A n g l e ct to p r e f e r r e d axis

Fig. 71. Relative (BH)max value of commercial hard ferrite grades vs. angle c~ to preferred axis (Joksch
1964, Stfiblein 1965). Absolute (BH)m,×(0 °) values range from 17.3 to 26.5 kJ/m 3 (2.2 to 3.3 MGOe), cf.
fig. 72.

As fig. 72 shows, the curves of the relative coercivity BHc vary to a large extent
for the individual specimens of fig. 71 in contrast to the curves of the (BH)max
values. With rigid magnetization in opposing fields up to at least BH~, ap-
proximately she ~ Br is expected. The magnet of curve (3) behaves roughly in this
way. If ~ / c < Br/tXo#.... with increasing a firstly jHc(a) and then B r ( a ) determine
the angular dependence of BHc(a).
The curve of the relative intrinsic coercivity ]Hc(a) is shown in fig. 73 for some
commercial hard ferrite specimens. By comparison, fig. 74 shows the depen-
dences for powder specimens manufactured by various methods. In both cases
the angular dependences of two extreme cases are also plotted: coherent mag-
netization reversal after Stoner-Wohlfarth (SW) and nucleation and growth of
reverse domains (Kondorsky 1940). In the latter case the effective coercivity is
given by the component of the field applied in the direction of the easy axis, i.e.
] H c - 1/cos a. Commercial hard ferrites show a curve between both extreme cases
which is far from SW behaviour. The lower jHc is, the greater is the tendency to a
better approximation of the cos -1 a curve. The findings of Ratnam et al. (1972), in
HARD FERRITES AND PLASTOFERRITES 565

5
..~

\
\
0o 10 o 20 ° 30 ° gO o 50 ° 60 ° 70 ° 80 ° 90 °

A n g l e cc to p r e f e r r e d axis
Fig. 72. Relative coercivity sHe of commercial hard ferrite grades vs. angle c~ to preferred axis. Curves
(1) to (3): Joksch (1964); curves (4) and (5): Stfiblein (1965). Same specimens as in fig. 71 having the
following characteristics:

Br (BH)max BSc jgc


Curve (mT) (kJ/m 3) (kA/m) (kA/m)

1 390 26.5 137 139


2 370 24.0 152 154
3 310 17.3
.........................................
204 227
4 380 25.2 153 155
5 345 20.8 191 193

which the coercivity is described very precisely by jHc(a) = 3 COS - 1 0 g k A / m for 100
to 200 fxm B a M crystals, fit into this pattern. O n the o t h e r hand, high-coercivity
p o w d e r s m a n u f a c t u r e d by precipitation processes show a certain t e n d e n c y
towards the S W curve, cf. curves (5) and (6) of fig. 74. All in all, it is clear that the
reversal of magnetization in the hard ferrites (as in all o t h e r p e r m a n e n t m a g n e t
materials) occurs, at least in the main, t h r o u g h n o n - c o h e r e n t reversal processes,
cf. B e c k e r (1967).
T h e angular d e p e n d e n c e of the p e r m a n e n t permeability was m e a s u r e d by
Joksch (1964), /~rec rising as o~ increases and, vertical to the preferred direction,
being 20 to 30% higher than parallel to it. This agrees well with the calculation
f r o m the m o d e l of c o h e r e n t reversal of magnetization, cf. Table 17. This m o d e l
therefore correctly describes the actual b e h a v i o u r as long as no incoherent process
has taken place.

Influence of mechanical stress


Mechanical impacts, stresses and vibrations do not influence the magnetic state of
hard ferrites. A s s a y a g (1963) f o u n d no effect b e y o n d the m e a s u r e m e n t accuracy of
566 H. ST.~J3LEIN

150 I / /4 r

I00 - . - _ x _ _ , : ×

~ ~ 3
\
8 \
~o

50

oc

0 N
0o ~o 20 ° 30 ° ~0 o 50 ° 60 ° 70 ° 80 ° 90 °

A n g l e ct to p r e f e r r e d axis
Fig. 73. Relative intrinsic coercivity sHe of commercial hard ferrite grades vs. angle c~ to preferred
axis. Curve designation as in fig. 72. Shown are also theoretical dependences according to
K o n d o r s k y - ( 1 / c o s a ) or to Stoner-Wohlfarth (SW) model. SW(0 °) = 1470 kA/m for SrM, cf. table 17.

75o%
cos ~ / I

50

0 x
0o I0 o 20 ° 30 ° 40 ° 50 ° 60 ° 70 ° 80 ° 90 °

Angle cL to p r e f e r r e d axis
Fig. 74. Relative intrinsic coercivity ~H~ of non-commercial powder specimens. (1) Pressed BaM
specimen having Br = 199 roT, (BH)max = 6.5 kJ/m 3, nHc = 90 kA/m, sH0 = 96 kA/m; density = 3.0 g/cm 3
(St~iblein 1965). (2) BaM powder, ball-milled, oriented, she(0 °) = 253 kA/m (Ratnam et al. 1972). (3)
Same as (2), but additionally anneaed at 950°C, ~Hc(0°) = 294 kA/m. (4) BaM powder, ball-milled,
oriented, she(0 °) = 120 kA/m (Haneda et al. 1973a). (5) BaM powder, co-precipitated, oriented, annealed
at 9 2 5 ° C , ~Hc(0°) = 490 kA/m (Haneda et al. 1973a). (6) BaM powder, crystallized from glassy borate
phase at 820°C, leached in dilute acetic acid, oriented, jHc(0 °) = 534 kA/m (Ratnam et al. 1970).
SW(0 °) = 1470 kA/m for SrM, cf. table 17.
HARD FERRITES AND PLASTOFERRITES 567

1% on isotropic and anisotropic hard ferrites caused by semi-sinusoidal impact with


acceleration amplitudes of 150 to 500 g lasting 1 ms, uniform acceleration up to 25 g
and vibrations of 10 to 5 000 H z with accelerations up to 20 g. For the m e c h a n i s m of
such effects on magnetostriction and the m a x i m u m elastic change attained, see
Lliboutry (1950).

Influence of neutron irradiation


Irradiation of isotropic and anisotropic Ba-hexaferrite specimens with fast n e u t r o n s
having energies of at least 1 M e V impaired all magnetic characteristics drastically
(Chukalkin et al. 1979). Specimens irradiated by 1.2 x 1024 m 2 s h o w e d only B~ of ca.
38%, (BH)max of ca. 13% and BHc and 1He of ca. 30% of the respective starting values.
Supposedly, irradiation causes Fe 3÷ cation vacancies in 2b lattice sites, which are
most i m p o r t a n t for the spin order. This o r d e r is turned f r o m an originally collinear
structure into a helical and then into a block angled one (Chukalkin et al. 1981).

3.4. Various physical and chemical properties

T h e most i m p o r t a n t magnetic characteristics were discussed in sections 3.1 to 3.3.


In the present section various physical and chemical properties and data are
c o m p i l e d which are of interest for classifying the materials and evaluating their
b e h a v i o u r and which are in part widely dispersed and difficult to find in literature.
T h e physical relationships cannot be dealt with in detail here so that reference
must be m a d e to the relevant literature.

Magnetostriction
Values for a B a M single crystal were given by Kuntsevich et al. (1968), cf. table
25. In particular it can be seen that turning the polarization f r o m the z direction
(= c-axis) into the x direction (= parallel basal plane) p r o d u c e s contraction in the
x direction and dilatation in both the y and z directions. T h e same authors f o u n d
in isotropic B a M specimens saturation magnetostrictions of hll = - (9 -- 0.5) × 1 0 -6
(direction of m e a s u r e m e n t parallel to the field) and h . = + ( 4 . 5 _ _ 0 . 5 ) × 10 -6
(direction of m e a s u r e m e n t perpendicular to the field). In contrast, R a t h e n a u
(1953) gives h ~ 20 × 1 0 - 6 with magnetization in the basal plane. T h e d o m i n a n t
TABLE 25
Saturation magnetostriction of a BaM single crystal magnetized and measured
along various directions (Kuntsevich et al. 1968). x, y, z = rectangular coordinate
system; z-axis parallel to hexagonal c-axis; x-axis parallel to one of the a-axis in the
basal plane.

Direction of Saturation
magnetization f i e l d measurement magnetostriction × 10 6

•~A x x - (15 _+0.5)


hB x y + (16 _+0.5)
hc x z +(11 _+0.5)
hD 45° to X and z 45° to x and z -(13+0.5)
568 H. ST~d3LEIN

contribution to magnetostriction originates from the Fe 3+ ions on 2b lattice sites


(Kuntsevich et al. 1980).
The saturation magnetostriction of the single crystal can be derived from the
values for AA... AD in table 25 for any direction of measurement and polarization,
cf. Mason (1954).

Resistivity
The specific resistivity p (and the specific conductivity o- = 1/p) of hard ferrites are
many orders of magnitude higher (and lower, respectively) than in metallic
materials and show the temperature dependence known from semiconductor
materials, cf. fig. 75, curve (a). Commercial ferrite specimens usually have a d.c.
resistivity of at least 10 6 ~ c m at room temperature (Went et al. 1952, Stuijts et al.
1955). This is, however, subject to strong fluctuations depending on composition
and manufacturing conditions, but generally this does not have an adverse effect
on applications because there is obviously no connection with the permanent

1000 200 2,0


'500 ' 100 0 - 5 0 -I00-125 -150oC
I 1

T mS i0~
(7 ,t,SGHz,~ • GHz tO2

lO~ 1 ~ I0kHz" 106


Hz ,200

x_-OHz ~crn
0 1 2 3 ,~ 5 6 10___..
3 8 9 I0~
I K
-.f ,.
Fig. 75. Specific conductivity o- and specific resistivity p vs, reciprocal absolute temperature T (and vs.
temperature in °C). Curve (a): D.C. values of a commercial, non-oriented BaM specimen of
composition BaFe31~gzO17.38 (Went et al. 1952, Haberey et al. 1968). Curves (b): Effective values of a
non-oriented BaM specimen of composition BaFe 3+ 12.59019.89 for various frequencies (Haberey et al.
1968).
HARD FERRITES AND PLASTOFERRITES 569

magnetic characteristics. Special agreements should only be made with the


manufacturer in those (rare) cases where emphasis is placed on high insulating
properties, for instance. It must also be noted that resistivity also depends on the
frequency, cf. fig. 75, curves (b). In this example the effective resistivity drops at
room temperature by about 3 powers of ten if a change is made from the d.c.
measurement to the 1 GHz a.c. measurement. At lower temperatures the ratio is
even greater, at higher temperatures smaller. According to measurements by
Rupprecht et al. (1959) the resistivity of anisotropic BaM specimens is up to about
one power of ten higher in the preferred direction than perpendicular to it, at
least in the frequency range 105 to 107Hz. Less anisotropy was found by
Dullenkopf (1968) in nominally isotropic specimens where the resistivity was up to
30% higher in the direction of pressing than perpendicular to it. This is possibly a
result of the crystal orientation caused by die pressing, cf. section 2.1.5.
In order to explain the resistivity behaviour let us take a model from a coated
dielectric in which highly conductive hard ferrite crystals are surrounded by grain
boundaries (barrier layers) of poor conductivity (Haberey et al. 1968, Dullenkopf
et al. 1969). Conductivity is attributable to electrons transient between Fe 3+ and
Fe 2+ ions. In the polycrystalline state the resistance to d.c. and low-frequency
current is largely determined by the properties of the barrier layers and may be
increased by post-annealing at 400-700°C in an oxidizing atmosphere or decreased
in a reducing atmosphere. Here, no connection with the Fe 2+ content was
observed (Dullenkopf 1968), but influencing this by substituting Ti, Mn and Co is
claimed to be possible (Dtilken 1971). We have a better understanding of the
resistance mechanism in the super-high frequency range, where the resistance is
inversely proportional to the Fe 2+ content (Dullenkopf et al. 1968) and which, in
extreme cases should equal the d.c. resistance (averaged over all spatial direc-
tions) of a single crystal owing to capacitive short-circuiting of the barrier layers.
The d.c. resistance was measured by Zfiv6ta (1963) for BaM and PbM. In both
cases the resistance parallel to the c-axis (P0 was roughly ten times that per-
pendicular to it (p±) in the range from -150 to 200°C. At room temperature
Pll ~ 700 l~cm and p± --~70 l~cm were found for BaM, Ptl ~ 50 ~ c m and p± ~ 5 ~ c m
for PbM.
Studies on the dielectric behaviour (e.g. Rupprecht et al. 1959, Haberey 1967,
Haberey et al. 1968, Dfilken 1971, Vollmerhaus et al. 1975), the thermo-electric
behaviour (e.g. Zfiv6ta 1963, Bunget et al. 1967, Dullenkopf 1968, Dullenkopf et
al. 1968, 1969) and the magnetic after-effects of hard ferrites (e.g. Haberey 1969,
Dfilken et al. 1969, Dfilken 1971) were carried out, partly in connection with the
electrical conductivity.

Linear thermal expansion


The extent of linear thermal expansion is of great interest in practice, firstly owing
to the relationship with crack formation as a result of thermal shock, and secondly
owing to the necessity to combine various materials in magnet systems exposed to
specific temperature ranges. With hexaferrites the conditions are more difficult
owing to the considerable anisotropy of thermal expansion. One particular
570 H. ST~3LEIN

consequence of this anisotropy is that crack-free toroids with a radial preferred


direction can be manufactured, if at all, then only with thin walls (Kools 1973, cf.
section 2.1.6). When hard ferrite magnets are cemented to metal components it is
recommended to use an elastic binding agent to compensate for the different
amounts of thermal expansion (Hamamura 1973).
The coefficients of thermal expansion are compiled with references in fig. 76
and table 26. a increases continuously from 0 K up to a maximum value at Curie
temperature both parallel and perpendicular to the c-axis, and then drops slightly.
Thermal expansion is appreciably greater along the c-axis than in the basal plane.
This is observed not only with specimens pressed in a magnetic field, cf. curves (1)
and (2) in fig. 76, but also with die-pressed, nominally "isotropic" specimens
which became slightly anisotropic during die pressing, cf. curves (3) and (4) (cf.
also section 2.1.5). Conversely, an anisotropy of only 1% was found in the
specimen relating to curve (5) (Buessem et al. 1957). As the comparison with the
thermal expansion of iron in fig. 76 shows, considerable expansion differences are
liable to occur when the material is cemented to the pole faces of isotropic and
anisotropic magnets.

0 500 K
i ~ i ! i i i I D i

15.'106 /
#

.. ,.;,4

,'3/, 5

•.,4
!'
II ~

0 ' ' , , ~ ,~ , , I
0 500 °C
Temperature
Fig. 76. Linear thermal expansion coefficienta of commercialhexaferrite specimens vs. temperature,
compared to a of iron.
Curve (1) SrM fl preferred axis
Curve (2) SrM ± preferred axis / (Van den Broek et al. 1977/78)
Curve (3) BaM ]1pressing axis ~ /
Curve (4)BaM _Lpressing axisJ (Clark et 1976) al.

Curve (5) BaM, isotropic (Buessem et al. 1957)


Curve (6) Iron.
H A R D FERRITES A N D PLASTOFERRITES 571

.~ ~ - ~ ~ ~ ~ _~ ~ ~ ~ ~
×

© o

o O o
t~
O

~.~
Z
©
e~.s

o o o ~~ ~~~ ~

r,
2s
~" 0 ~ ~ ~.~
09

,--k

8
0
"-d

I I l l l I

°~

e~

= ~0
e~

©
.=
O

*
572 H. STJkBLEIN

Other thermal properties


4 W/(m • K) (Clark et al. 1976) and 5.5 W/(m • K) (Valvo 1978/79) were given for the
thermal conductivity of B a M and 0.84 J/(gK) ( T D K 1978) and 0.714 J/(gK) (Torii
et al. 1979) for the specific heat.

Elastic properties
T h e characteristics of various moduli are compiled in table 27. At least the
modulus of elasticity is anisotropic (Kools 1973, Iwasa et al. 1981). According to
Cavalotti et al. (1979) this modulus depends to a relatively small extent on
composition and manufacturing conditions. The t e m p e r a t u r e dependence of all
three m0duli is very low according to R e d d y et al. (1974); at -193°C the values
were only a few percent lower than at r o o m temperature. Iwasa et al. (1981)
determined the temperature dependence of BaM specimens as - 2 8 . 9 (isotropic),
- 3 7 . 7 (]]c) and - 5 2 . 2 N/mm2K (2c) between r o o m t e m p e r a t u r e and 900°C.
Poisson's n u m b e r of an isotropic B a M specimen was given as 0.28 (Clark et al.
1976) and as 0.24 (Iwasa et al. 1981).

TABLE 27
Elastic moduli in kN/mm2 of commercial BaM specimens at room temperature.

Property Specimen Value Reference

Young's modulus isotropic 198 ]


anisotropic, IIpreferred axis 177 ~ Kools (1973)
anisotropic, ± preferred axis 211 J
isotropic 151.9 Reddy et al. (1974)
isotropic 138 Clark et al. (1976)
isotropic 130-140 Cavallotti et al. (1979)
isotropic 183 }
anisotropic, [[ preferred axis 154 Iwasa et al. (1981)
anisotropic, L preferred axis 317

Rigidity modulus isotropic 63.5 Reddy et al. (1974)

Bulk modulus isotropic 83.4 Reddy et al. (1974)

Mechanical strength
The tensile and flexural strength values compiled in table 28 exhibit very large
fluctuations both within the series of m e a s u r e m e n t s made by one author and
between various authors. Most of these fluctuations are probably attributable to
the ceramic, brittle nature of the materials and the particular difficulties encoun-
tered in measuring. The fluctuations almost completely cover up the effects of
composition, manufacturing conditions and anisotropy. It can be stated with
reasonable certainty that both the tensile and flexural strength values are less than
150 N/ram 2. According to Wills et al. (1976) strength increases with the quality of
the surface finfsh.
H A R D FERRITES AND PLASTOFERRITES 573

TABLE 28
Strength values in N/mm 2 of commercial hard ferrites at room temperature.

Property Specimen Value Reference

Tensile anisotropic BaM; 11preferred axis 36.3--+ 7.8


Kools (1973)
strength anisotropic BaM; ± preferred axis 73.4 ± 10.8 /
isotropic BaM 4.9-7.8
IHamamura (1973)
anisotropic BaM 11.8-18.6 J
isotropic BaM 55 Clark et al. (1976)
isotropic, anisotropic M 19.6-49.0 TDK (1978)

Flexural isotropic BaM 6.8-9.8


Gershov (1963)
strength anisotropic BaM 2.5-8.0 l
anisotropic BaM, SrM; I preferred axis 86.3 _+7.8 Kools (1973)
isotropic BaM 77 ± 33
Hamamura (1973)
anisotropic BaM, SrM 88-147 J
isotropic, anisotropic M 29.4-88.3 TDK (1978)
isotropic BaM 60-130 Cavallotti et al. (1979)

Compressive anisotropic BaM 735 ± 78 Kools (1973)


strength isotropic BaM 440 Clark et al. (1976)
isotropic, anisotropic M >686 TDK (1978)

On the other hand, the compressive strength determined by various authors


exhibits comparatively small variations. As expected, the values are considerably
higher than the tensile and flexural strength values.
For the impact strength (notched-bar impact test) Hamamura (1973) found
values between 2.4 and 9.8 kgm/cm 2 without any distinct effect of anisotropy but
proportional to the density.
The critical stress-intensity factor K~c was determined in connection with studies
on crack formation and the grindability of hexaferrites, cf. section 2.1.7. The
values in table 29 clearly show an influence of anisotropy. Cleavage along the

TABLE 29
Critical stress-intensity factor Krc in MN/m 3/2 of commercial hard ferrites at room temperature.

Specimen Value Reference

Anisotropic SrM, fracture surface ]]preferred axis


2.12"8[) Veldkamp et al. (1976)
Anisotropic SrM, fracture surface ± preferred axis
Anisotropic SrM, fracture surface II preferred axis 2.1-2.5"
1.5-2.1" Veldkamp et al. (1979)
Anisotropic SrM, fracture surface 2 preferred axis
Anisotropic BaM, fracture surface II preferred axis 2.83-+0.101
Anis0tropic BaM, fracture surface ± preferred axis 0.96±0.05~ Iwasa et al. (1981)
Isotropic BaM 1.57 ± 0.04J
Isotropic BaM 1.3-3.2 Cavallotti et al. (1979)

* Maximum values are obtained for molar ratio FeaO3/SrO ~ 5.5.


574 H. S T ~ 3 L E I N

basal plane is easier than perpendicular to it. According to Veldkamp et al. (1976)
K~c depends on the molar ratio Fe2OjSrO = n with anisotropic SrM. It slowly
increases from n = 4.4 to a maximum at n ~ 5.5 and drops even more quickly
when n is higher. The figures in table 29 are averages; the related variations
reached a maximum of about _ 1 g N / m 3/2. A dependence on the molar ratio was
also found with isotropic BaM specimens, but the K,c maximum occurred at
n ~ 4 . 5 or lower (Cavallotti et al. 1979). The same authors also observed an
influence of the pressing process: with dry-pressed specimens K~c=
2.73 _+0.24 MN/m 3/2, with wet-pressed specimens K~ = 2.05 MN/m 3/2. According to
Iwasa et al. (1981) KI~ of BaM specimens decreased between room temperature
and 1000°C linearly and rather slightly and between 1000 and 1200°C drastically.
In the first mentioned region the temperature dependence is 6.23--_ 0.03 (isotro-
pic), 0.42 -- 0.03 ([]c) and (8.38 -+ 0.06) x 10 .4 MN/m 312K (±c), and for the second
region in the order of 10 -a MN/m 3/2 K. No anomaly was found around the Curie
temperature. Fracture surface energies of 6.35 + 0 . 3 2 (isotropic), 2.82-+0.30 (][c)
and 11.92-+ 0.84 J/m 2 ( ± c ) w e r e deduced.

T A B L E 30
Hardness of commercial hard ferrites at room temperature.

Hardness Specimen Value Reference

Vickers Anisotropic SrM, II preferred 8.6 k N / m m ;


axis Veldkamp et al.
Anisotropic SrM, L preferred 5.6 k N / m m 2 (1976)
axis
SrM 6.5 k N / m m 2 Broese van Groenou
et al. (1979a)
Anisotropic B a M ca. 6 k N / m m 2 see fig. 44
SrM, single cryst, no. 33, 10.3 ± 1.0 k N / m m 2
IIc-axis
SrM, single cryst, no. 33, 7.5 ±0.5 k N / m m 2 Jahn (1968)
± c-axis
SrM, single cryst, no. 57, 9.6 ± 1.0 k N / m m 2
II c-axis
SrM, single cryst, no. 57, 8.3 ± 0.5 k N / m m 2
L c-axis
PbM, single crystal 11.0 k N / m m 2 Courtel et al. (1962)

Ritz SrM 24 k N / m m z Broese van Groenou


et al. (1979a)

Rockwell, Isotropic B a M 73-76 }


Scale A Anisotropic B a M 72-80, Gershov (1963)
Rockwell, Isotropic B a M 80-90 Cavallotti et al. (1979)
Scale N
load 147 N

Mohs M 6-7 Schiller et al. (1970)


HARD FERRITES AND PLASTOFERR1TES 575

Hardness
Table 30 shows hardness values determined by various tests. Like most of the
mechanical properties hardness is thus anisotropic. Figure 44 shows the hardness
as a function of the sintering temperature and thus of the density.

Lattice constants, X-ray density


Lattice constants of M-ferrites are compiled in table 31. There is only a slight
difference for the Ba, Sr and Pb compounds so that the volumes of the unit cells
only differ by a maximum of 1%, giving values for the theoretical densities of 5.3,
5.1 and 5.6 to 5.7g/cm 3 respectively. A density of 5.59g/cm 3 obtains for the
mineral magnetoplumbite with the lattice constants given by Berry (1951) and the
(idealized) composition given by Blix (1937) as PbFev.sMn3.sA10.sTi0.sO19.

TABLE 31
Lattice constants of M-ferrites in nm (at room temperature).

Specimen a c c/ a Reference

BaM 0.5876 2.317 3.943 Adelsk61d (1938)


BaM 0.5893 2.3194 3.936 Townes et al. (1967)
BaM 0.5894 2.321 3.938 Haberey et al. (1977b)
Ba0.sSr0.2M 0.5887 2.318 3.937 Haberey et al. (1977b)
Ba0.6Sr0.4M 0.5890 2.316 3.932 Haberey et al. (1977b)
Ba0.4Sr0.6M 0.5882 2.311 3.929 Haberey et al. (1977b)
Bao.zSr0.sM 0.5884 2.308 3.923 Haberey et al. (1977b)
SrM 0.5884 2.308 3.923 Haberey et al. (1977b)
SrM 0.5885 2.303 3.914 Haberey et al. (1977b)
SrM 0.5887 2.305 3.915 Haberey et al. (1977b)
SrM 0.5864 2.303 3.927 Adelsk61d (1938)
PbM 0.5877 2.302 3.917 Adelsk61d (1938)
PbM 0.5893 2.308 3.916 Klingenberg et al. (1979)

Magnetoplumbite 0.606 2.369 3.91 Aminoff (1925)


Magnetoplumbite 0.588 2.302 3.915 Berry (1951)

Chemical stability
Being oxides, hard ferrites are, by nature, particularly unstable in strong acids, but
are subject to more or less rapid chemical attack even in weak acids, alkalis and in
other chemicals. Qualitative information on this is contained in table 32 (Schiller
et al. 1970, Valvo 1978/79). Quantitative data on corrosion behaviour in different
media will be found in table 33 (Hirschfeld et al. 1963) in which the specific loss of
weight after 14 days of treatment is given. These and other data point to a number
of remarkable aspects:
(1) Fluoric acid and hydrochloric acid are by far the most aggressive media.
(2) An increase in the temperature of the medium of only 30°C can mean a
dramatic increase in corrosion.
(3) The rise in the corrosion rate with increasing temperature is not uniform for
576 H. ST~ilBLEIN

TABLE 32
Chemical stability of BaM-ferrites.

Rather stable in: More or less unstable in:

Ammonia Hydrochloric acid


Acetic acid Hydrofluoric acid
Benzol-trichloroethylene (50 : 50) Nitric acid
Citric acid (5%) Oxalic acid
Citric acid (10%) Phosphoric acid
Cresol Sulfuric acid
Developer
Fixing bath
Hydrogen peroxide (15%)
Hydrogen peroxide (30%)
Petrol
Phenol solution
Potassium hydroxid solution
Sodium chloride solution (30%)
Sodium hydroxide solution
Sodium sulphate solution

various media, so that the consecutive order, in terms of aggressiveness, depends


on temperature.
(4) A comparison between periods of treatment of 1, 4 and 8 days shows that
the specific loss of weight is not constant. For this reason the loss of material
cannot be extrapolated quantitatively to other reaction periods.
(5) The above comparison further indicates that in some media the specimen

TABLE 33
Specific weight loss of non-oriented Ba-hexaferrite specimens after being
treated for 14 days in different aqueous media (Hirschfeld et al. 1963).
Concentration: 15 wt %.

Specific weight loss


Medium in g/m 2 at
20-25°C 50-55°C

Hydrofluoric acid (a) (a)


Hydrochloric acid (a) (a)
Sulfuric acid 57 300
Phosphoric acid 25 650
Potassium hydroxid solution 11 9
Nitric acid 8 610
Tartaric acid 4 60
Acetic acid 4 4
Ammonia 1 (b)
Aqua destillata 2(c) - 15(c)

(a) Specimen completely


dissolved
(b) not determined
(c) means weight increase
HARD FERRITES AND PLASTOFERRITES 577

weight increases first and then decreases, Firmly adherent reaction products
obviously form in the initial stage. These dissolve as the attack progresses.
All in all the corrosion behaviour is very c o m p l e x and it depends to a
considerable extent upon the test parameters. Quantitative predictions therefore
require exactly defined test conditions.

3.5, Comparison with other permanent magnet materials; applications

Since their discovery around 1950 hard ferrites have enjoyed a bigger upswing in
sales than any other p e r m a n e n t magnet material. Figure 77 shows the rate of
increase in output of various hard magnet grades as estimated by various authors.
In spite of considerable fluctuations, a hard ferrite output in the order of 108 kg/a
can be assumed for the beginning of the eighties in the western countries. Since
the average price of hard ferrites is estimated at 10 DM/kg, this represents a value
today of approximately 109DM/a. The value of the other p e r m a n e n t magnet
materials can be expected to be in roughly the same order of magnitude because
the comparably smaller tonnage output is set against a correspondingly higher
average price.
T h e main reason for this large proportion of hard ferrites in the total output of
p e r m a n e n t magnets is their economy (see also section 1.2). The price per unit of
magnetic energy ((BH)m~ value) is much lower for hard ferrites than for the other
magnet materials (Steinort 1973, Rathenau 1974). A m a j o r advantage offered by
hard ferrites is the low-cost and almost inexhaustible supply of raw materials
which opens up excellent prospects for the use of this material in f u t u r e - in spite
of the fact that the energy density values of hard ferrite grades are inferior to
those of other materials as can be seen from figs. 78 and 79.

108

5 +/A/

2- f . .---"
//.-
10z
/
/ t
5 / t
/ /' •
2.

106 I I I I I
1930 1940 1950 1950 19F0 1980year
Fig. 77. Annual mass production of permanent magnets between 1940 and 1980 in the western world:
( ) hard ferrites (upper curve, Cartoceti et al. 1971, Steinort 1973, Andreotti 1973; lower curve,
Van den Broek et al. 1977/78, Rathenau 1974); (0) hard ferrites (Schiller 1973); (---) all permanent
magnets, (..... ) alloys only (Van den Broek et al. 1977/78, Rathenau 1974); (+) all permanent magnets
(Schiller 1973).
578 H. ST~BLEIN

/.5

T
2O
II

/, \
/.0
~,1~'°'~ 2

a5 ~
0,5

-750 kA/m - 500 -250 0


Field s t r e n g t h H
Fig. 78. Typical demagnetization curves of various permanent magnet materials: (1) hard ferrite; (2)
AlNiCo, high B~ grade; (3) AlNiCo, high H~ grade; (4) Mn-AI-C alloy; (5) RECo5 alloy; (6)
RE(Co, Cu, Fe)7 alloy.

QI 1 tO kOe
I~ ~ I I I !!II I, ~ I I I III!I I ! I I
: kh 25
'MGOe
kJ/m3
l I00.
J xt6\
/ 14 ,~x~ 24 xU/"
,,*_L3 _ 9 - "
×23
:~o
- - -
×191"72"',
-7--" %
.... , x22
", , 9 × ,17. 'L*×~2#21
/ x I
," 3/ q8 : ~ -x', "--"
[× / /x 6

10. ,'J~7," x 20
ix
', ',~ ', ', ', ', 1:',II ', ', ', ' , ; I I ' , I ', ',
10 lO0 1000 kA/m
jH c

Fig. 79. Permanent magnet materials compared by (BH)max (static energy) and intrinsic coercivity aHc
(stability) (St~iblein 1972): (1) Co steel; (2) Cu-Ni-Fe, anisotropic; (3) Co-Fe-V-Cr, anisotropic, wire;
(4) Co-Fe-V-Cr, anisotropic, strip; (5)-(7) AlNiCo, isotropic; (8)-(12) AlNiCo, anisotropie; (13)-(16)
AlNiCo, columnar; (17) ESD (elongated single domain), anisotropic; (18) and (19) Cr-Fe-Co,
anisotropic; (20) Hard ferrite, isotropic; (21) Hard ferrite, anisotropie, see table 18; (22) Mn-Bi,
anisotropic; (23) Mn-AI-C, anisotropic; (24) Pt-Co, isotropic; (25) RE-Cos-type alloys, anisotropic;
(26) RE-(Co, Cu, Fe, Mn)7-type alloys, anisotropic.
HARD FERRITES AND PLASTOFERRITES 579

Figure 78 shows the demagnetization curves of typical modern permanent


magnet grades with their (BH)max points. Owing to their relatively low (B/-/')max
values hard ferrites are not particularly suitable for those applications in which the
most important requirement is to keep the volume of the (static) magnet system as
small as possible. As soon as minimum possible mass is required conditions
change, however, as can be seen from the comparison of the columns "volume
efficiency" and "mass efficiency" in table 34. Owing to its low density hard ferrite
is appreciably better in this case.
The same applies to those cases where importance is attached to stability or
reversible magnetic behaviour under alternating magnetic fields, i.e. to coercivity
jHc. For a qualitative comparison of the materials in terms of their static as well as
dynamic characteristics, it is advisable to use the (BI-I)max-jHcdiagram, cf. fig. 79.
This brief description is also used in more recent standardization, cf. section 3.1.
The hard ferrites are thus found to range in stability between the Alnico,
C r - F e - C o and ESD magnets, on the one hand, and the intermetallic compounds
MnBi, PtCo and R E (Co, Cu, Fe, Mn)y, on the other. The (BpH)max value* is a
quantitative measure for the energy conversion capability with dynamic operation
of the permanent magnet, cf. table 34. In these applications hard ferrites are at
least equal to the AlNiCo alloys.
In fig. 78 the ( B H ) m a x points are marked on the demagnetization curves. If flux
density and field strength in the permanent magnet assume the values cor-
responding to the (BH)m~ point, a minimum volume of permanent magnet is
required (with static applications). This can be achieved by an appropriate design
of the magnet system. In this case B/txoH has to assume the values listed in table
34. The smaller this value is the more compact the permanent magnet has to be
designed. Like the R E - C o alloys, hard ferrites therefore as a rule have compact
shapes (small ratio between magnet length and magnet cross-sectional area) in
contrast to AlNiCo alloys.
Another important criterion for actual use is the temperature response of the
magnetic properties, cf. section 3.2. Table 34 shows the temperature coefficients of
the remanence a(Br) and of the intrinsic coercivity a(jHc) of some materials, as
applying for a certain range around room temperature. Hard ferrites can be seen
to exhibit the greatest temperature response of remanence, and this applies, of
course, to the temperature response of the magnetic flux in the permanent magnet
system. Hard ferrites are thus less suitable for applications where functioning
must remain unaffected by temperature.
Special mention has already been made in section 3.2 of the consequences of
the large positive temperature coefficient of jHc of hard ferrites and in particular
of the risk of irreversible losses on cooling below room temperature. Certain
AlNiCo grades show a similar although less pronounced behaviour.
Amenability to shaping and machining is an important criterion in actual

* For the definition see section 3.1, footnote on p. 536. Some authors use half or one quarter of this
value. References: D e s m o n d (1945), Schwabe (1958, 1959), Schiller (1967), St~iblein (1968b), Gould
(1969), Zijlstra (1974).
580 H. S T ~ 3 3 L E I N

¢,q

g~

x
?
~s
A × x
tt%
,4

.d ©

I ,?
?

?
e.-,

,..-i

,.a

tel
,,'5

$
< <
H A R D FERRITES AND PLASTOFERRITES 581

(magnification I/4)
Fig. 80. Assortment of hard ferrite permanent magnets (courtesy of Fried. Krupp GmbH, Krupp
WIDIA, D-4300 Essen).
582 H. STJ~BLEIN

practice. Like AlNiCo* and the RE-Co materials, hard ferrites are produced by
powder metallurgical methods. Being brittle they can only be machined by cutting
processes producing minute chips?. Where the material has to be ductile, e.g., for
making sheet or wire, permanent magnets of Co-Fe-V-Cr, Cr-Fe-Co or Cu-Ni-
Fe alloys have to be used (Zijlstra 1978).
Table 34 further shows some comparative properties which occasionally have to
be observed in the choice of material: the high specific electrical resistance of hard
ferrites in comparison with metallic materials, their relatively low tensile strength
and their relatively low linear thermal expansion.
Hard ferrites are found in a variety of applications, the most important aspect
being that of economy. Magnetically speaking it is the high coercivity and the
resultant high stability and reversibility of the magnetic state ("rigid polarization")
which led to their use especially in motors, generators, holding, attracting,
repulsing, coupling and eddy current devices. It became apparent that they can be
used to advantage in permanent magnet loudspeaker systems owing to their still
adequate (BH)max values. General descriptions and special topics on the ap~
plication of hard ferrites can be found in text books of Hadfield (1962), Ireland
(1968), McCaig (1967, 1977) Moskowitz (1976), Parker et al. (1962), Reichel (1980)
and Schiller et al. (1970). A major disadvantage is the high temperature depen-
dence which impairs their use in measuring instruments and at high or low
temperatures. Hard ferrites for microwave applications were described by Akaiwa
(1973) and by Nicolas (1979). Figure 80 shows a selection of hard ferrite magnets
for various applications.

4. Bonded hard ferrites, plastoferrites

In these composite materials hard ferrite powder is embedded in a non-mag-


netizable matrix. As with all composite materials, the different properties of two
component materials can also be united here in one constructional element,
producing new technological possibilities. Since the volume proportion of the
magnetic phase is smaller than in the compact hard ferrites, reduced magnetic
properties result. The bonded hard ferrites are therefore only used for those
applications where this relatively low magnetic level is acceptable. If this is the
case, then the advantages of the composite material, e.g., formability, low cost and
improved non-magnetic properties, come to the fore. If synthetic or natural
organic materials are used as matrix materials, their wide-ranging technological
possibilities, e.g., in terms of elasticity, strength, resistance to fracture and impact,
dimensional tolerance, lower density, shaping and further processing can be fully
exploited for the magnetic composite materials. Nowadays, only organic materials
are used for the matrix on a large scale. Common materials are rubber, poly-

* Manufacture by melting processes is feasible too and commonly used.


t AlNiCo alloys after special treatment also by turning and drilling (Pant 1977, Pant et al. 1977).
HARD FERRITES AND PLASTOFERRITES 583

vinylchloride (PVC), polyamides (PA), polyolefins (e.g. polyethylene and poly-


propylene), polystyrene, phenol and polyester resins (Casper et al. 1965).
Inorganic materials such as metal and glass have also been proposed but there
seems to be no practical demand for them and their use would be more expensive.
In this connection mention should be made of experiments by Passerone et al.
(1975) on the wettability of hard ferrite with molten metals and tests by Cavallotti
et al. (1976, 1977) and Asti et al. (1976) on the electrochemical and chemical
manufacture of inorganically bonded hard ferrites. For glass-bonded hard ferrites,
cf. section 2.2.3.
The following sections will therefore only deal with the hard ferrites bonded
with an organic matrix, which are also called plastic bonded hard ferrites or,
simply, plastoferrites.

4. i. Manufacturing technologies for plastoferrites

Figure 81 shows a diagrammatic representation of the process for manufacturing


plastoferrites (Caspar et al. 1965, Richter et al. 1968b).
In many cases raw material is used which is obtained in manufacturing compact
magnets anyway, e.g., after reaction sintering or as waste material after final
sintering or in indirect shaping, cf. fig. 19 and sections 2.1.1 to 2.1.4 and 2.1.7. This
material is then processed further in order to meet the special requirements
placed on the desired plastoferrite grade. These requirements relate in particular
to high coercivity, alignability of the powder particles and the desired ferrite
content of the powder. The top grade powders, however, require special
manufacturing techniques tailor-made to meet the requirements of the plastofer-
rites.
As far as coercivity is concerned, it should be high enough so that no
irreversible polarization reversals occur along the B - H demagnetization curve,
i.e., a rigid magnetization is maintained. This is ensured by using a crystallite size
of roughly 1 txm.
Depending on whether anisotropic or isotropic magnets are to be manufac-
tured, the powder particles must be alignable or non-alignable. Alignability is
achieved when the particles are either monocrystalline or made from compact
material with preferred direction. If alignability is not wanted, however, in order
that isotropic magnets are obtained, then the use of polycrystalline particles from
isotropic material is appropriate since monocrystalline particles can be aligned
during the shaping process owing to their platelet-like shape, cf. section 2.1.5. Asti
et al. (1974) have referred to the significance of the particle shape as regards
alignability and to the influence of special additions during the reaction process.
Particle size and particle size distribution are major factors affecting the
maximum attainable ferrite content as coarser powders can be more densely
packed.
The above-mentioned requirements imposed on the crystallite and particle size
are largely fulfilled if suitable reaction and grinding conditions are selected. If
intensive grinding is necessary, the resultant lattice defects must be eliminated by
annealing prior to further processing, cf. section 2.1.4.
584 H. ST~13LEIN

Reacted hexaferrite from raw materials;


or waste from sintered bulk material

Crushing, milling

Annealing

Mixing with matrix material:


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

Thermoplastic Rubber Thermosetting

..... rigid . . . . ~.... ~llsti; . . . . . . . . . ;lai:i~ . . . . . . . . . ;igid- . . . .

Injection extrusion Rolling, Die


moulding calenderinfl pressing
with or with or
without without
magnetic magnetic
field field

Annealing

Cutting, machining I

AssembLing, magnetizing

Fig. 81. Manufacturing technology for plastic bonded hard ferrites (plastoferrites).
HARD FERRITES AND PLASTOFERRITES 585

Depending on the material used, the ferrite powder is mixed with the binding
agent either in the cold or hot state in mixers, mixing extruders, kneaders or
calenders. With increasing proportions of ferrite powder the viscosity of the
paste-like mixture increases drastically in the loading range considered and also,
as a consequence, the power needed for mixing and working. This is reflected in
the form of wear on the machines and in the form of lowered jHc values in the
magnets as a result of plastic deformation of the particles. The optimum percen-
tage of ferrite for anisotropic magnets in order to obtain the desired magnetic
properties lies between 50 and 65% by volume with injection moulding as
otherwise the particles would interfere with one another when being aligned in
the magnetic field. The corresponding figure for rolling is about 70% by volume.
Slightly higher ferrite contents of about 70% by volume are also possible in
injection moulding if no or nonintensive alignment is required. Ferrite contents of
80% by volume or even more can only be attained in die pressing.
The subsequent shaping stage produces finished parts (injection moulding, die
pressing) or semi-finished parts such as ribbons, strips etc. (extrusion, rolling)
which are then punched or cut into their final shape. Special mention should be
made of the dimensional accuracy and the cost-efficiency, especially of injection
moulding, where further machining is generally not required. Mechanical and
magnetic forces can be used to manufacture anisotropic magnets, the former
mainly in extruding and rolling, with the platelet shape of the crystallites being
exploited (cf. section 2.1.5), the latter mainly in injection moulding and die
pressing comparable to the manufacture of compact anisotropic magnets (cf.
section 2.1.5). If rubber is used as binding agent, vulcanization may be carried out
after shaping, thus increasing the strength of the magnet. Thermosetting plastics,
too, require heat treatment where the plastic is irreversibly cured. In injection
moulding and die pressing the shaping of the actual magnetic material can be
combined with the fitting of insert components (axles, pole sheets, etc.). Figure 82
shows an assortment of injection moulded magnets.

4.2. Technical properties and applications of plastoferrites

Before discussing the magnetic characteristics let us deal with the question as to
what values can be attained. Only an approximate answer can be given as it is not
yet known how the demagnetization curves of a compact permanent magnetic
material and a composite material using the same grade as magnetic component can
be accurately transposed from one to the other. Approximate calculations are based
on the conception that magnetic material and matrix, with the total volume
remaining constant, are separate and concentrated in themselves. Here, a difference
must be made between two extreme cases depending on whether permanent
magnetic and matrix materials lie in parallel or in series with the magnetic flux
(Edwards et al. 1975, Joksch 1976). In the former case, the magnetic flux, in
comparison with the compact specimen, is diluted in proportion to the volume
content of the magnetic material, whereas in the latter case shear is present as in a
magnet system with air gap. Here the amount of shear is expressed by the volume
586 H. STi~d3LEIN

Fig. 82. Injection moulded plastoferrites (Ebeling et al. 1978).

ratio between magnetic and matrix material. The demagnetization curves B(H) can
be easily calculated for both extreme cases, assuming that there is no flux leakage a n d
jHc is sufficiently high. The actual curve will lie between the two extreme cases.
The model calculation is particularly simple if the magnetic material exhibits
rigid magnetization with/£rec 1.0, i.e. with complete orientation of all crystallites,
=

as the demagnetization curves B(H) are then identical for both extreme cases.
This gives the values listed in table 35 for the anisotropic specimen. The
corresponding demagnetization curves for bonded Sr hexaferrite magnets are
shown in fig. 83 for various volume percentages xv of the magnetic material. The
curve for Xv = 1.0 corresponds to the Stoner-Wohlfarth curve of the anisotropic
magnet in fig. 59.
When #rec exceeds 1.0 the r e m a n e n c e values for the two extreme cases differ,
the difference increasing as /Xrec rises and the proportion of ferrite Xv falls. The
series model provides slightly lower values than the parallel model. The m a x i m u m
differences are, however, only 8% for /Xrec= 1.171 (isotropic specimen, cf. table
17) and for the technically interesting cases of Xv~> 0.5. Therefore, useful (some-
what too high) approximations result from the parallel model for the isotropic
specimen, too. These are also listed in table 35. Figure 84 shows the respective
demagnetization curves for various values of xv. The curve for Xv = 1.0 cor-
responds to the Stoner-Wohlfarth curve of the isotropic magnet shown in fig. 59.
From studies on elastic bonded hard ferrites Besjasikowa et al. (1977) gave a
H A R D FERRITES A N D PLASTOFERRITES 587

© ©
0 0

0 @

X~ X ~ x

© X

X X x ~ X

N
"m.

© ©

'~, t"q r--

x~x

×x~x
~e

•E =,
N . E

II
"r.
©

©
588 H. ST.Sd3LEIN

Magnetic
Field strength H . flux ~B
-5 kOe -4 -3 -2 -1 _ J .iX v
densit~v_^
, , , .

0¢6) 300 |

-400 kA/m -300 -200 -100 0


Field strength H
Fig. 83. Demagnetization curves B(H) of anisotropic commercial plastoferrites (curves (1) to (3))
compared with ideal coherent rotation model (Stoner-Wohlfarth curve "S-W", 100% ; and fractions of
that). (1) Injection moulded magnet (Drossel et al. 1969); (2) Rolled magnet (Blondot 1973); (3) Die
pressed magnet (catalogue of Fa. Baermann, Bensberg near Cologne, F.R.G.).

Magnetic
Field strength H ~ flux ~B
^. density ~ Xv
-e.s .Oe -2.0 -1.5 -1.0 ,2so !
I ~ t .... 100
~ mT Io/o
-: 2oo_l so

-200 kA/m -150 -100 -50 0


Field strength H =
Fig. 84. As fig. 83 but for i s o t r o p i c plastoferrites: (1) Injection moulded magnet (Ebeling et al. 1978);
(2) Rolled or extruded magnet (Blondot 1973); (3) Die pressed magnet (Ebeling et al. 1978).

s o m e w h a t different d e p e n d e n c e of the characteristics on the ferrite c o n t e n t Xv:

Br = 290x~ 4 m T ; BHc = 135xv k A / m (1.7Xv k O e ) ;

(BH)max = 10.3x 24 kJ/m3(1.3x 24 M G O e ) .

T h e s e s p e c i m e n s had possibly b e c o m e s o m e w h a t anisotropic.


A s regards the coercivity jHc n o effect of t h e packing density is expected as
HARD FERRITES AND PLASTOFERRITES 589

there should be no or only slight interaction. In fact, Shimizu et al. (1972) found
no influence in the range Xv = 0.5 to 0.7 with high-coercivity barium hexaferrite
powders where j H c = 2 4 5 k A / m (3.1kOe). Somewhat different findings were
reported by Hagner (1980) with SrM powders treated in various ways. A special
preparation procedure assured microscopic deagglomeration of the particles. With
increasing packing density xv the coercivity jHc first increased drastically, while no
or only minor changes occurred above Xv-~ 0.5. This is explained by assuming a
positive magnetostatic interaction between the SrM particles. With powder pre-
paration the possible introduction of plastic deformation and its influence on
coercivity has to be considered, see section 2.1.4. Wohlfarth (1959) provides a
survey of a series of theoretical and experimental studies concerning the influence
of the packing density on the coercivity of various materials.
Figures 83 and 84 show the demagnetization curves of some high-quality
commercial grades. Lower quality grades can be easily produced by reducing the
ferrite content and are therefore not included. Isotropic magnets attain 70 to 75%,
anisotropic magnets only 50 to 55% of what is theoretically possible.
This different performance is due to the incomplete orientation of the crystallites
in bulk material compared with monocrystal orientation and to the lower maximum
ferrite content of the anisotropic grades. Similar viewpoints apply here as already
described in section 3.1 for the compact magnets. Samow (1973) considers the
manufacture of anisotropic magnets with Br = 270 mT (2.7 kG) and (BH)max =
14 kJ/m 3 (1.8 M G O e ) commercially feasible, for which a ferrite powder content of
70% by volume with an 80% degree of alignment would be necessary. The
demagnetization curves of anisotropic magnets, measured perpendicular to the
preferred direction, are, of course, lower than the curves of the isotropic grade,
cf. fig. 85. Table 36 contains magnetic characteristics and densities of some
internationally standardized bonded hard ferrite grades (IEC-Document 68 CO

Field strength H
-3 kOe -2 -1
1 I I 3OO
mT

- 200 l

100 ~

0
-300
400 .~

-200

-300
Fig. 85. Demagnetization curves of isotropic (1) and anisotropic plastoferrites (2 and 3: measured
parallel and perpendicular to preferred direction, resp.) (Ebeling et al. 1978).
590 H. S T ~ d 3 L E I N

,"-,
~rd

~b la~ tt~
t-- t--

eag

8t )
tt~

0 ©

,...,

© ..,9.
©

O © .rz

e~

© oJ
O
:=.
HARD FERRITES AND PLASTOFERRITES 591

TABLE 37
Standard designations and trademarks of plastoferrites from
various countries. Some trademarks are of historical interest
only. Some manufacturers use the same trademark (with
different additional designation) for their sintered grades, see
table 18, page 540.

Standard
Country designation Trademarks

Federal Rep. Hartferrit P KOEROX P


Germany Oxidur P
Oxilit
Prox
Sprox
Tromaflex (TX)
Tromalit (OP, OT)

German Democ. Maniperm Maniperm


Rep.

France Ferriflex
Plastoferroxdnre
Plastoferrite

Great Britain Magnadur P, D, Sp

Italy Ferriplast
Plastomag

Netherlands Ferroxdure P, D, Sp

Japan Ferrogum
KPM
MBS
RM
RN
YRM

USA Koroseal
Magnalox
Magnyl
Plastiform (PL)

24, 1980). T h e s e a r e b a s e d on t h e values in D I N 17 410 ( M a y 1977) of t h e F e d e r a l


R e p u b l i c of G e r m a n y . T h e d e s i g n a t i o n s y s t e m i s t h e s a m e as that for c o m p a c t
m a g n e t s , cf. section 3.1. In t h e G e r m a n D e m o c r a t i c R e p u b l i c t h e r e is a n o t h e r
n a t i o n a l s t a n d a r d for i s o t r o p i c a n d a n i s o t r o p i c g r a d e s : T G L 16541/04 (1979).
T a b l e 37 c o n t a i n s a s u m m a r y of t r a d e m a r k s of v a r i o u s countries.
T h e t e m p e r a t u r e d e p e n d e n c e s of t h e m a g n e t i c c h a r a c t e r i s t i c s c o r r e s p o n d to
t h o s e of c o m p a c t m a g n e t s , cf. t a b l e 22.
592 H. STJ~d3LEIN

Electrical, thermal, mechanical and chemical properties largely depend on the


type of matrix material and the ferrite content. General statements can only be
made to a limited extent and inquiries must be made at the manufacturer's in each
individual case.
The automobile industry places special requirements on the matrix material as
regards its resistance to oil, petrol and solvents. The television industry demands a
high degree of flame retardancy. Special grades have been developed for these
purposes.
Plastoferrites are used in a large number of applications, see fig. 82, where the
main criterion, more so than with compact magnets, is the high economy of these
material groups. The following applications deserve particular mention (Caspar et
al. 1965, Hinderaker 1976, Samow 1973, Badner 1978): holding magnets, mainly
for refrigerator catches but also for planning boards, toys etc., correction magnets
for television tubes, fractional-horsepower motors and dynamos, actuating mag-
nets for reed switches, electronic ignition systems and separators (filters) for
cleaning fluids. The application as a magnetic information storage medium is
discussed by Fayling (1979). According to Samow (1973) s o m e 106 kg of bonded
hard ferrites are produced annually in Western Europe alone.

References

Abrams, J.C. and M.G. McLaren, 1976, J. Asti, G. and S. Rinaldi, 1974, Proc. 3rd. Eur.
Amer. Ceram. Soc. 59, 347-350. Conf. on Hard Magn. Mat. Amsterdam, 91-
AdelskGld, V., 1938, Arkiv fGr kemi, mineralogi 93.
och geologi 12A, No. 29, 1-9. See also: Asti, G., P. Cavallotti and R. Roberti, 1976,
Zeitschr. Kristallographie, Erg. Bd. 6, 1941, Proc. 3rd. Cimtec Rimini, paper 21;
Strukturbericht Vol. VI, 1938, 74-75. and 1977, Ceramurgia Internat. 3, no. 2,
Akaiwa, Y., 1973, Jap. J. Appl. Phys. 12, No. 70-77.
i1, 1742-1747. Asti, G., M. Carbucicchio, A. Deriu, E. Luc-
Aminoff, G., 1925, Geol. FGren. FGrhandl. 47, chini and G. Slokar, 1980, J. Magn. Magn. Mat.
No. 3, 283-289. 20, 44-46.
Andreotti, R., 1973, 1. IOS-Kolloquium, Badner, JJ., 1978, J. Appl. Phys. 49 (3), 1788-
Varese, Italien. 1789.
Anonymous, 1967, Machine Design (16, 3, Balek, V., 1970, J. Mater. Sci. 5, 714-718.
1967) p. 12. Barham, D. and A.E. Schwalm, 1974, J. Canad.
Appendino, P. and M. Montorsi, 1973, Ann. di Ceram. Soc. 43, 27-29.
Chimica, 63, 449-456. Batti, P., 1960, Ann. di Chimica, 50, 1461-
Ardelean, I., E. Burzo and I. Pop, 1977, Solid 1478.
State Commun. 23, 211-214. Batti, P., 1961a, Ann. di Chimica, 51, 1318-
Ardelean, I. and E. Burzo, 1980, J. Magn. Magn. 1339.
Mat. 15-18, 1369-1370. Batti, P., 1961b, Univ. Trieste, Fac. Ingegneria
Arendt, R.H., 1973a, J. Appl. Phys. 44, 3300- No. 11.
3305. Batti, P., 1962a, Ann. di Chimica, 52, 941-961.
Arendt, R.H., 1973b, J. Solid State Chem. 8, Batti, P., 1962b, Ann. di Chimica, 52, 1227-
339-347. 1247.
Assayag, P., 1963, Bull. Soc. Franc. Electr. 4, Batti, P., 1976, Ceramurgia, 6, 11-16.
5-23. Batti, P. and G. Sloccari, 1967, Ann. di Chi-
Asti, G., 1976, Ceramurgia 6, No. 1, 3-10. mica, 57, 777-804.
H A R D FERRITES AND PLASTOFERRITES 593

Batti, P. and G. Sloccari, 1968, Ann. di Chimica, Broese van Groenou, A., 1975, IEEE Trans.
58, 213-222. Magn. MAG-11, No. 5, 1446-1451.
Batti, P. and G. Sloccari, 1976, Ceramurgia, 6, Broese Van Groenou, A. and J.D.B. Veldkamp,
No. 3, 136-141. 1979a, Phil. Techn. Rdsch. Nr. 4/5, 109-123.
Bauer, I., 1976, Wahlarbeit W 146 der TH Broese van Groenou, A. and P.E.C. Franken,
Aachen. 1979b, Proc. Brit. Ceram. Soc. No. 28, 243-
Becker, J.J., 1962, Metallurg. Rev. 7, No. 28, 266.
371-432. Broese van Groenou, A., J.D.B. Veldkamp and
Becker, J.J., 1967, J. Appl. Phys. 38, 1015-1017. D. Snip, 1977, J. de Phys. 38, C1-285-289.
Beer, H.B. and G.V. Planer, 1958, Brit. Com- Buessem, W.R. and A. Doff, 1957, Proc. 13.
munications and Electronics, 5, 939-941. Annual Meeting, Metal Powder Association,
Belyanina, N.V., I.N. Ivanova, Yu.G. Sak- Chicago/Ill., Vol. II: Ferrites and Electron.
sonov and A.A. Shvarts, 1977, Russ. J. Inorg. Core Session, 196-204.
Chem. 22 (11), 1718-1719. Bulzan, M. and E. Segal, 1976, Rev. Roumaine
Beretka, J., 1968, Div. Build. Res. CSIRO, de Chim. 21, 651~53.
Report F2-1, Melbourne, Australia. Bungardt, K., P. Kagner and F. Th/immler,
Beretka, J. and M.J. Ridge, 1968, J. Chem. Soc. 1968, DEW-Techn. Ber. 8, 157-187.
(A) 2463-2465. Bunget, I. and M. Rosenberg, 1967, Phys. Sta-
Beretka, J. and T. Brown, 1971, Aust. J. Chem. tus Solidi, 21, K 131-K 133.
24, 237-242. Burzo, E., I. Ardelean and I. Ursu, 1980, J.
Berger, W. and F. Pawlek, 1957, Arch. Eisen- Mat. Sci. 15, 581-593.
hfittenwes. 28, 101-108. Buschow, K.H.J., W. Luiten, P.A. Naastepad
Bergmann, F., 1958, Ber. Arb. Gem. Ferro- and F.F. Westendorp, 1968, Phil. Techn. Rev.
magn. (Dr. Riederer-Verlag, Stuttgart). 29, 336-337.
Berry, L.G., 1951, Amer. Mineralogist, 36, 512- Bye, G.C. and C.R. Howard, 1971, J. Appl.
514. Chem. Biotechnol. 21, 319-323.
Bertaut, E.F., A. Deschamps, R. Pauthenet Bye, G.C. and C.R. Howard, 1972, J. Appl.
and S. Pickart, 1959, J. Phys. Rad. 20, Chem. Biotechnol. 22, 1053-1064.
404-408. Calow, C.A. and R.J. Wakelin, 1968, J. Inst.
Besjasikowa, T.G., B.W. Aisikowitsch, A.G. Met. 96, 147-154.
Alekseew, S.A. Kowatschewa and A.E. Cartoceti, A. and E. Steinort, 1971, La metal-
Kornew, 1977, Elektritschestvo (Moskow) No. lurgia italiana-atti notizie No. 10, 291-
1, 81-83. 295.
Black, D.B., A.E. Schwalm and D. Barham, Cartoceti, A., F. Scansetti and E. Steinort, 1976,
1976, J. Canad. Ceram. Soc. 45, 4%52. Ceramurgia 6, 39-46.
Blix, R., 1937, Geol. F6ren. F6rhandl. 59, No. Casimir, H.B.G., J. Smit, U. Enz, J.F. Fast,
3, 300-302. H.P.J. Wijn, E.W. Gorter, A.J.W. Duy-
Blondot, 1973, Preprints Aimants 1973; vesteyn, J.D. Fast and J.J. de Jong, 1959, J.
Chambre syndicale des producteurs d'aciers Phys. Rad. 20, 360-373.
fins & speciaux, 12, rue de Madrid, Paris 8°; Caspar, H.J. and G. Samow, 1965, Valvo-Ber.
paper Q. 11, H. 5, 136-145.
Bohning, R.G., 1978, Powd. Met. Internat. 10, Cavallotti, P., U. Ducati and R. R0berti, 1976,
37. Ceramurgia, 6, No. 1, 17-20.
Bottoni, G., D. Candolfo, A. Cecchetti, L. Cavallotti, P., R. Roberti, G. Caironi and G.
Giarda and F. Masoli, 1972, Phys. Status Asti, 1977, J. de Phys. 38, Coll. C 1, suppl, au
Solidi A, 32, K47-K50. No. 4, C1-333-C1-336.
Bowman, W.S., Sutarno, N.F.H. Bright and Cavallotti, P., R. Roberti, A. Cartoceti and F.
J.L. Horwood, 1969, J. Canad. Ceram. Soc. Scansetti, 1979, IEEE Trans. Magn. MAG-
38, 1-8. 15, No. 3, 1072-1074.
Brady, L.J., 1973, J. Mat. Sci. 8, 993-999. Cho, K. and K. Kim, 1975a, J. Korean Ceram.
Braun, P.B., 1957, Phil. Res. Rep. 12, 491- Soc. 12, 71-75.
548. Cho, K. and K. Kim, 1975b, J. Korean Ceram.
Brisi, C. and P. Rolando, 1969, Ann. di Chi- Soc. 12, 76-81.
mica, 59, 385-399. Chroust, V., 1972, Fortschritte der Pulver-
594 H. STJ~d3LEIN

metallurgie (Forsch. Inst. in Sumperk, CSSR) Dullenkopf, P. and H.P.J. Wijn, 1969, Z.
No. 2, 3-38. Angew. Phys. 26, 22-29.
Chukalkin, Yu.G., V.V. Petrov and B.N. Durant, B. and J.M. Pfiris, 1980, Ann. Chim.
Goshitskii, 1979, Izvest. Akad. Nauk SSSR, Fr. 5, 589-595.
Neorg. Mater. 15, No. 7, 1307-1308. Durant, B. and J.M. P~ris, 1981, J. Mat. Sci.
Chukalkin, Yu.G., V.V. Petrov and B.N. Lett. 16, 274-275.
Goshchitskii, 1981, Phys. Status Solidi A, 67, Ebeling, R. and H. Krause, 1978, VDI-Beric-
421-426. hte, No. 309, Dfisseldorf.
Clark, A.F., W.M. Haynes, V.A. Deason and Edwards, A. and H.E. Gould, 1975, Report of
R.J. Trapani, 1976, Cryogenics, May, 267- Fa. IOS, Malgesso (Varese), Italy.
270. Efremov, G.L. and I.I. Petrova, 1977 Iz. Akad.
Cocco, A., 1955, Ann. di Chimica, 45, Nauk SSSR, Anorg. Mat. 13, 318-321.
737-753. Eisenhuth, C., 1968, Stahl und Eisen, 88, 264.
Cochardt, A., 1966, J. Appl. Phys. 37, No. 3, 269.
1112-1115. Emberson, C.C., W.C.M. Leung and D. Bar-
Cochardt, A., 1969, DE-OS (Deutsche Offenle- ham, 1978, J. Canad. Ceram. Soc. 47, 1-5.
gungsschrift) 1 911 524. Erchak Jr., M., I. Fankuchen and R. Ward,
Courtel, R., H. Makram and G. Pigeat, 1962, 1946, J. Amer. Chem. Soc. 68, 2085-2093.
Compt. Rend. 254, No. 26, 4447 4449. Erickson, R.H., 1962, US-Patent, 3 155 623.
Curci, T.J., W.R. Bitler and R.C. Bradt, 1978, Erzberger, P., 1975, Firmenschrift der Fa.
Mat. Sci. Res. 11, Proc. Cryst. Ceram. Bayer, Krefeld, Technische Eisenoxide zur
359-368. HersteUung von Ferriten.
Dambier, Th. and K. Ruschmeyer, 1960, AEG- Esper, FJ. and G. Kaiser, 1972, Int. J. Magn. 3,
Mitt. 50, 388-391. 189-195.
Davis, R.T., 1965, thesis Pennsylvania State Esper, F.J. and G. Kaiser, 1974, Bosch Techn.
University. Ber. 4, 308-314, and Proc. 3rd Eur. Conf. on
DeBitetto, D.J., 1964, J. Appl. Phys. 35, 3482- Hard Magn. Mat. (Amsterdam 1974) 106--
3487. 108.
Denes, P.A., 1962, Amer. Ceram. Soc. Bull. 41, Esper, F.J. and G. Kaiser, 1975, Ber. Dr.
509-512. Keram. Ges. 52, 210213.
DE-OS (= Deutsche Offenlegungsschrift) Esper, FJ. and G. Kaiser, 1978, Ber. Dt.
2 110 489 (Priority: Japan, 2, 3, 70). Keram. Ges. 55, 294-295.
Desmond, D.J., 1945, J. Instr. electr. Eng., P. 2. Fagherazzi, G., 1976, Ceramurgia, 6, 26-32.
Power Engng. 92, No. 27, 229-244. Fagherazzi, G. and L. Giarda, 1974, Proc. 3rd
Dietrich, H., 1968, Feinwerktechnik 72, 313. Europ. Conf. on Hard Magn. Mat. (Am-
322 and 425-433. sterdam) 79-82.
Dietrich, H., 1969, Feinwerktechnik 73, 171- Fagherazzi, G., C.M. Maggi and G. Sironi,
180 and No. 5, 199-208. 1972, Ceramurgia, 2, 181-189.
Dietrich, H., 1970, IEEE Trans. Magn. MAG- Fahlenbrach, H., 1953, Elektrotechn. Zeitschr.
6, No. 2, 272-275. A, 388--389.
Dixon, S., M. Weiner and T.R. AuCoin, 1970, Fahlenbrach, H., 1963, Techn. Mitt. Krupp,
J. Appl. Phys. 41, 135%1358. Forsch.-Ber. 21, 113-119.
Dornier System GmbH, 1979, EP 0 011 265 A2 Fahlenbrach, H., 1965, Techn. Mitt. Krupp,
(Europ. Pat.). Forsch.-Ber. 23, 26-35.
Drofenik, M. and D. Kolar, 1970, Ber. Dt. Fahlenbrach, H., 1972, Metall 26, 1230-1234.
Keram. Ges. 47, 666~68. Fahlenbrach, H. and W. Heister, 1953, Arch.
Drossel, E. and G. Samow, 1969, Valvo-Ber. Eisenh/ittenwes. 24, 523-528, and 1954,
15, H. 2, 58--63. Techn. Mitt. Krupp 12, 47-51.
Dtilken, H., 1971, thesis Aachen. Fayling, R.E., 1979, IEEE Trans. Magn. MAG.
Dtilken, H., F. Haberey and H.P.J. Wijn, 1969, 15, No. 6, 156%1569.
Z. Angew. Phys. 26, 29-31. Fischer, E., 1962, Angewandte Meg- und
Dullenkopf, P., 1968, thesis Aachen. Regeltechnik 2, No. 24, a229-a233.
Dullenkopf, P. and H.P.J. Wijn, 1968, Elek- Fischer, E., 1978, Powd. Met. Internat. 10, 30-
troanzeiger, 21, 472-473. 32 and Sprechsaal Nr. 2, 38-42.
HARD FERRITES AND PLASTOFERRITES 595

Flink, G., 1924, Geol. F6ren. F6rhandl. 46, Goto, Y. and K. Takahashi, 1971, J. Jap. Soc.
No. 6/7, 704-709. Powd. Met. 17, 193-197.
Foniok, F. and S. Makolagwa, 1977, J. Magn. Goto, K., M. Ito and T. Sakurai, 1980, Jap. J.
Magn. Mat. 4, 95-104. Appl. Phys. 19, No. 7, 1339-1346.
Frei, E.H., S. Shtrikman and D. Treves, 1959, J. Gould, J.E., 1962, Magnetic Stability, in: Per-
Appl. Phys. 30, 443. manent Magnets and Magnetism, ed. D.
Gadalla, A.M. and H.W. Hennicke, 1973, Hadfield (Iliffe Books Ltd., London) p. 443-
Powd. Met. Internat. 5, 196-200. 472.
Gadalla, A.M. and H.W. Hennicke, 1975, J. Gould, J.E., 1969, IEEE Trans. Magn. MAG-5,
Magn. Magn. Mat. 1, 144-152. No. 4, 812-821.
Gadalla, A.M., H.E. Schfitz and H.W. Hen- Granovskii, I.V., Yu.D. Stepanova, D.Ya.
nicke, 1976, J. Magn. Magn. Mat. 1, 241- Serebro,R.A. Lysyak and E.M. Krisan, 1970,
250. Soy. Powd. Metallurgy and Met. Ceram. No. 6,
Gallagher, P.K., D.W. Johnson Jr., F. Schrey 486-490 and No. 11,895-901.
and D.J. Nitti, 1973, Ceram. Bull. 52, 8,42- Gray, T J . and R.J. Routil, 1972, Sympos.
849. Electr. Magn. Opt. Ceramics, London, 13-14
Gallo, G., 1936, Annali di Chimica, 109-115. Dec. 1972, 91-103.
Gasiorek, S., 1980, Sci. of Ceramics, 10, 311- Haag, R.M., 1969, Joint Fall Meeting, Elec-
319. tronics Div. and New England Section, Am
Gerling, W. and H.P.J. Wijn, 1969, Z. Angew. Ceram. Soc. Boston.
Phys. 27, 77-82. Haag, R.M., 1971, IEEE Trans. Magn. MAG-7,
Gershov, LYu., 1963, Soviet Powd. Metallurgy Sept., 609.
and Metal Ceram. No. 3 (15), 22%234. Haberey, F., 1967, thesis Aachen.
Gershov, I.Yu., 1971, Soy. Powd. Met. 10, 388- Haberey, F., 1969, J. Appl. Phys. 40, 2835-
392. 2837.
Giarda, L., 1976, Ceramurgia, 6, 33-38. Haberey, F., 1978, Ber. Dt. Keram. Ges. 55,
Giarda, L., A. Cattalani and A. Franzosi, 1977, 297-301.
J. de Phys. 38, Coll. C 1, Suppl. au no. 4, Haberey, F. and A. Kockel, 1976, IEEE Trans.
C 1-325-328. Magn. MAG-12, No. 6, 983-985.
Giarda, L., G. Bottoni, D. Candolfo, A. Cec- Haberey, F. and F. Kools, 1980a, Ferrites, Proc.
chetti and F. Masoli, 1978, Ceramurgia Int. ICF 3, 356-361.
4, No. 2, 79-81. Haberey, F. and H.P.J. Wijn, 1968, Phys. Status
Giron, V.S. and R. Pauthenet, 1959, Compt. Solidi, 26, 231-240.
Rend. 248, 943-946. Haberey, F. and H. Wullkopf, 1977a, private
Glass, H.L. and J.H.W. Liaw, 1978, J. Appl. communication.
Phys. 49, 1578-1581. Haberey, F., M. Velicescu and A. Kockel,
Glass, H.L. and F.S. Stearns, 1977, IEEE 1973a, Int. J. Magnetism 5, 161-168.
Trans. Magn. MAG-13, No. 5, 1241-1243. Haberey, F., M. Velicescu and A. Kockel,
Glazacheva, M.V. and L.S. Zevin, 1972, Izves- 1973b, Int. J. Magnetism 5, 161-168; F.
tiya Akademii Nauk SSSR, Anorg. Mat. 8, Haberey, K. Kuncl, M. Velicescu; Linseis-
No. 9, 1638-1640. Journal 1/73, 6-10.
Gmelin, 1959, 8th ed., system No. 59, Iron, Part Haberey, F., A. Kockel and K. Kuncl, 1974,
D, Magnetische Werkstoffe Suppl. Vol. 2, p. Ber. Dt. Keram. Ges. 51, 131-134.
43d d41. Haberey, F., G. Oehlschlegel and K. Sahl, 1977b,
Goldman, A. and A.M. Laing, 1977, J. de Phys. Ber. Dt. Keram. Ges. 54, 373-378.
38, (C1) 297-301. Haberey, F., R. Leckebusch, M. Rosenberg and
Gordes, F., 1973, Wahlarbeit W 138 des In- K. Sahl, 1980b, Mat. Res. Bull. lS, 493-500 and
stituts ffir Werkstotte der Elektrotechnik, TH IEEE Trans. Magn. MAG-16, No. 5, 681-683.
Aachen. Hadfield, D. (ed.), 1962, Permanent Magnets and
Gordon, I., 1956, Ceram. Bull. 35, 173-175. Magnetism (Iliffe Books, London, Wiley, New
Goto, K., 1972, J. Jap. Soc. Powd. Met. 18, York).
209-216. Hagner, J., 1980, Hermsdorfer Techn. Mitt. 20,
Goto, Y., T. Takada, 1960, J. Amer. Ceram. Soc. No. 55, 1764-1771, 20, No. 56, 1810-1815,
43, 150-153. and 1981, 21, No. 57, 1819-1825.
596 H. STJ/G3LEIN

Hamamura, A., 1973, Sumitomo-Sondermetall Hiraga, T., 1970, Ferrites, Proc. Intern. Conf.
Techn. Ber. 1, 37-46. Japan, 179-182.
Hamamura, A., 1977, Techn. Ber. der Fa. Hirotsu, Y., H. Sato, 1978, J. Solid State Chem.
Sumitomo Sondermetalle 3, 29-35. 26, 1-16.
Haneda, K. and H. Kojima, 1973a, J. Appl. Hirschfeld, D. and W. Fischer, 1963, unpub-
Phys. 44, 3760-3762. lished.
Haneda, K. and H. Kojima, 1974a, J. Amer. Hodge, M.H., W.R. Bitler and R.C. Bradt,
Ceram. Soc. 57, No. 2, 68--71. 1973, J. Amer. Ceram. Soc. 56, 49%501.
Haneda, K., Ch. Miyakawa and H. Kojima, Hodge, M.H., W.R. Bitler and R.C. Bradt,
1973b, Bull. Res. Inst. Sci. Meas. Tohoku 1975, in: Deformation of Ceramic Materials,
Univ. 22, No. 1, 67-78. ed., R.C. Bradt and U.R.E. Tressler (Plenum
Haneda, H., Ch. Miyakawa and H. Kojima, Press, New York) p. 483M96.
1974b, J. Amer. Ceram. Soc. 57, No. 8, 354- Horn, E., P. KaBner and H. Dietrich, 1968,
357. DEW-Techn. Ber. 8, 234-242.
Haneda, K., C. Miyakawa and H. Kojima, Hoselitz, K. and R.D. Nolan, 1970, IEEE
1975, AIP Conf. Proc. No. 24, Magn. Magn. Trans. Magn. MAG-6, No. 2, 302.
Mat. 1974, Amer. Inst. Phys. 770771. Ichinose, N. and K. Kurihara, 1963, J. Phys.
Harada, H., 1970, Ferrites, Proc. Int. Conf. Soc. Jap. 18, 1700-1701.
Japan, 279-282. Ichinose,N. and Y. Tanno, 1975, Proc. US-
Harada, H., 1980, Ferrites, Proc. ICF 3, 354. Japan Sem. Sci. Ceram., Hakone, Japan,
355. 207-211.
Harvey, J.W. and D.W. Johnson Jr., 1980, Ireland, J.R., 1959, Appl. Magnetics, Indiana
Ceramic Bull. 59, No. 6, 637~539, 645. Gen. Corp. 7.
Hausknecht, P., 1913, thesis, Strasbourg. Ireland, J.R., 1968, Ceramic permanent-magnet
Hayashi, N., Y. Syono, Y. Nakagawa, K. motors (McGraw-Hill, New York).
Okamura and S. Yajima, 1980, Sci. Rep. Ito, S., Y. Kajinaga, I. Imai and I. Endo, 1974,
RITU, A, 28, No. 2, 164-171. J. Jap. Soc. Powd. Met. 21, No. 5, 132-139.
Heck, C., 1967, Magnetische Werkstoffe und Iwanow, O.A., E.W. Shtolts and J.S. Shut,
ihre technische Anwendung, (Dr. Alfred 1966, Fiz. Metal. i Metalloved. 22, 455-
H/ithig Verlag, Heidelberg) and 1975, 2nd 458.
edition. Iwasa, M., E.C. Liang, R.C. Bradt and Y.
Heidel, M. and W. Schneider, 1977, Hermsdor- Nakamura, 1981, J. Amer. Ceram. Soc. 64,
fer Techn. Mitt. 17, 1572-1574. No. 7, 390-393.
Heimke, G., 1958, Naturwissenschaften, 45, Jfiger, P., 1976, Keram. Zeitschr. 28, Nr. 9,
260-261. 454456; 1978, Keram. Zeitschr. 30, No. 5,
Heimke, G., 1960, Ber. Arb. Gem. Ferromagn. 246-249.
1959, Stahleisen, 213.221. Jahn, L., 1968, thesis Halle; see also 1967, Phys.
Heimke, G., 1962, Ber. Dt. Keram. Ges. 39, Status Solidi, 19, K 75-K 77.
326-33O. Jahn, L. and H.G. Miiller 1969, Phys. Status
Heimke, G., 1963, Z. Angew. Phys. 15, 271-272. Solidi, 35, 723-730.
Heimke, G., 1964, Z. Angew. Phys. 17, 181- Jander, W., 1927, Z. anorg, u. ailg. Chemie,
183. 163, 1-30.
Heimke, G., 1966, Ber. Dt. Keram. Ges. 43, Jaworski, J.M., G.A. Ingham, W.S. Bowman
600.604. and G.E. Alexander, 1969, J. Canad. Ceram.
Heimke, G., 1976, Keramische Magnete, Appl. Soc. 38, 171-175.
Mineralogy 10, (Springer, Wien). John, W., 1973, Farbe + Lack, 79, 537-542.
Hempel, K.-A., P. Grosser, 1964, Z. angew. Johnson Jr., D.W., 1981, Ceram. Bull. 60, No.
Phys. 17, 153.157. 2, 221-224, 243.
Hempel, K.A. and C. Voigt, 1965, Z. angew, Joksch, Ch., 1964, DEW-Techn. Ber. 4, 182-
Phys. 19, 108-112. 188.
Hennig, G., 1966, thesis Berlin; see also IEEE Joksch, Chr., 1976, J. Magn. Magn. Mat. 2,
Trans. Magn. MAG-2, No. 3, 165-166. 303--307.
Hinderaker, P.D., 1976, Machine Design Jonker, G.H. and A.L. Stuijts, 1971, Phil.
(Febr. 12th) 94.98. Techn. Rev. 32, 79-95.
H A R D FERRITES AND PLASTOFERRITES 597

Kanamaru, F., M. Shimada and M. Koizumi, Krijtenburg, G.S., 1970, IEEE Trans. Magn.
1972, J. Phys. Chem. Sol. 33, 1169-1171. MAG-6, No. 2, 303.
Kanamaru, F., K. Oda, T. Yoshio, M. Shimada Krijtenburg, G.S., 1974, Proc. 3rd Eur. Conf.
and K. Takahashi, 1981, J. Jap. Soc. Powd. and on Hard Magn. Mat., Amsterdam, 83-86.
Powd. Met. 28, 70-76. Kronenberg, K.J. and M.A. Bohlmann, 1960, J.
Kantor, P., A. Revcolevschi and R. Collongues, Appl. Phys. Suppl. 31, 82 S-84 S.
1973, J. Mat. Sci. 8, 1359-1361. Krupi6ka, S., 1973, Physik der Ferrite und der
Kaf3ner, P., 1970, Chemie-Ing.-Technik, 42, 48. verwandten magnetischen Oxide (Vieweg,
Klingenberg, R. and K. Sahl, 1979, Ber. Dt. Braunschweig).
Keram. Ges. 56, 75-78. Kuntsevich, S.P., Yu.A. Mamalu i and A.S.
Klug, F.J. and J.S. Reed, 1978, Ceramic Bull. Mil'ner, 1968, Fiz. metal, metalloved. 26, No.
57, No. 12, 1109-1110, 1115. 4, 610-613.
Knight, F., 1962, The Magnetizing, Testing and Kuntsevich, S.P., Yu. A. Mamaluj and V.P.
Demagnetizing of Permanent Magnets, in: Palekhin, 1980, Sov. Phys. Solid State, 22 (7),
Permanent Magnets and Magnetism, ed., D. 1278-1279.
Hadfield (Iliffe Books Ltd., London) p. 401- Lacour, C. and M. Paulus, 1973a, Compt.
442. Rend. Ser. C, 277, 1001-1004.
Kockel, A., 1980, private communication. Lacour, C. and M. Paulus, 1973b, Compt.
Kohatsu, I. and G.W. Brindley, 1968, Z. phys. Rend. Ser. C, 277, 1085-1088.
Chemie, N.F. 60, 79-89. Lacour, C. and M. Paulus, 1975a, Phys. Status
Kohn, J.A., D.W. Eckart and C.F. Cook Jr., Solidi A, 27, 441--456.
1971, Science, 172, 519-525. Lacour, C. and M. Paulus, 1975b, Phys. Status
Kojima, H., 1955a, Sci. Rept. Res. Inst. Tohoku Solidi A, 28, 71-80.
Univ. 7, Ser. A, No. 5, 502-506. Landolt-B6rnstein, 1962, Vol. 1I, Part 9, Mag-
Kojima, H., 1955b, Sci. Rept. Res. Inst. Tohoku netic properties I, Sect. 2924: Hexagonal
Univ. 7, Ser. A, No. 5, 507-514. Ferrites (Springer, Berlin) p. 2-222 to
Kojima, H., 1956, Sci. Rept. Res. Inst. Tohoku 2-236.
Univ. 8, Ser. A, 540546. Landolt-B6rnstein, 1963, 6th ed., Vol. 1V, Part.
Kojima, H., 1958, Sci. Rept. Res. Inst. Tohoku 2, p. 132.
Univ. 10, Ser. A, 175-182. Landolt-B6rnstein, 1970, Vol. 4, Part b, Sect. 7:
Kojima, H., Ch. Miyakawa and N. Nakaigawa, Hexagonal Ferrites (Springer, Berlin) p. 547-
1969, Bull. Res. Inst. Sci. Meas. Tohoku 583.
Univ. 17, 1-12. Landolt-B6rnstein, 1981, Group III, Vol: 12c,
Kojima, H., K. Goto and C. Miyakawa, 1980, in press.
Ferrites, Proc. ICF 3, 335-340. Laville, H. and J.C. Bernier, 1980, J. Mat. Sci.
Kondorsky, E., 1940, J. Phys. USSR 2, 161-181. 15, 73-81.
K6nig, U., 1974, Techn. Mitt. Krupp, Forsch.- Laville, H., J.C. Bernier and J.P. Sanchez, 1978,
Ber. 32, 75-84. Solid State Commun. 27, 259-262.
Kools, F., 1973, Sci. Ceramics, 7, 27-44. Lliboutry, L., 1950, Thesis, Grenoble.
Kools, F., 1974, Proc. 3rd Eur. Conf. on Hard Lotgering, F.K., 1959, J. Inorg. Nucl. Chem. 9,
Magnetic Mat., Amsterdam, 98-101. 113-123.
Kools, F., 1975, Ber. Dt. Keram. Ges. 52, 213- Lotgering, F.K. and M.A.H. Huyberts, 1980,
215. Solid State Commun. 34, 49-50.
Kools, F., 1978a, Ber. Dt. Keram. Ges. 55, 301- Lucchini, E. and G. Sloccari, 1976, Ceramurgia
304. International, 2, 13-17.
Kools, F., 1978b, Ber. Dt. Keram. Ges. 55, Lucchini, E. and G. Slokar, 1980a, J. Mat. Sci.
296-297. 15, 2123-2125.
Kools, F., M. Klerk, P. Franken and F. den Lucchini, E. and G. Slokar, 1980b, J. Magn.
Broeder, 1980, Sci. Ceramics, 10, 349-357. Magn. Mat. 21, 93-96.
Kooy, C., 1958, Phil. Techn. Rev. 19, 286-- Mackintosh, G.H. and P.F. Messer, 1976,
289. Science of Ceramics Vol. 8, Brit. Ceram.
Krijtenburg, G.S., 1965, Vortrag Nr. 2.13 auf Soc., 403-414.
der 1. Europ. Tagung fiber Magnetismus in Mahdy, A.N. and A.M. Gadalla, 1976a, J.
Wien. Magn. Magn. Mat. 1, 326-329.
598 H. ST~d3LEIN

Mahdy, A.N. and A.M. Gadalla, 1976b, J. N6el, L., R. Pauthenet, G. Rimet and V.S.
Magn. Magn. Mat. 1, 330-332. Giron, 1960, J. Appl. Phys., Suppl. 31, 27
Malakhovskij, A.N., 1980, Soy. Powd. Metal- S-29 S.
lurgy Met. Ceram. (SPMCAV) 19, No. 3, Nicolas, J., 1979, Rev. techn. Thomson c.s.f. 11,
201-203. No. 2, 243-258.
Mamaluj, Yu.A., A.A. Murakhovskij and L.P. Nishikawa, T., T. Nishida, K. Inoue, H. Inoue
Ol'khovik, 1975, Inorg. Mater. 11, 1145-1146. and I. Uei, 1974, Yogyo Kyokai Shi, 82, No. 5,
Mansour, N.A., A.M. Gadalla and H.W. Hen- 241-247.
nicke, 1975, Ber. Dt. Keram. Ges. 52, 201- Oda, K., T. Yoshio, K. Takahashi, 1982, J. Jap.
204. Soc. Powd. Met. 29, No. 2, 39-44.
Mason, W.P., 1954, Phys. Rev. 96, 302-310. Odor, F. and A. Mohr, 1977, IEEE Transl
Maurer, Th. and H.G. Richter, 1966, Powder Magn. MAG-13, No. 5, 1161-1162.
Metallurgy 9, Nr. 18, 151-162, and 1966, Okamura, T., H. Kojima and Y. Kamata, 1952,
Sprechsaal ffir Keramik, Glas, Email, Sili- J. Appl. Phys. (Japan) 21, 9-12.
kate, 99, Heft 24, 1084-1089. Okamura, T., H. Kojima and S. Watanabe,
Maurer, Th. and H.G. Richter, 1972, Powd. 1955, Sci. Rept. Res. Inst. Tohoku Univ. 7,
Met. Int. 4, 78-8i. Ser. A, No. 4, 411-417 and 418-424.
McCaig, M., 1967, Attraction and repulsion Okazaki, K. and H. Igarashi, 1970, Ferrites,
(Oliver and Boyd, Edinburgh). Proc. Internat. Conf. Japan, 131-133.
McCaig, M., 1977, Permanent magnets in Oron, M. and P. Ramon, 1975, IEEE Trans.
theory and practice (Pentech Press, London). Magn. MAG-11, No. 5, 1452-1454.
Mee, C.D. and J.C. Jeschke, 1963, J. Appl. Otsuka, T., Y. Yamamichi, Y. Watanabe, K.
Phys. 34, No. 4, part 2, 1271--1272. Kanaya and T. Sasaki, 1973, J. Jap. Soc.
Menashi, W.P. and T.R. AuCoin, J.R. Shap- Powder and Powd. Met. 20, No. 5,
pirio, D.W. Eckart, 1973, J. Cryst. Growth, 126-132.
20, 68-70. Pant, P., 1977, Techn. Mitt. Krupp, Forsch.-
Meriani, S., 1972, Acta Cryst. B 28, 1241-1243. Ber. 35, No. i, 59-64.
Metzer, A. and Ch. Gorin, 1975, Intermag. Pant, P. and H. Stfiblein, 1977, Proc. World
London, Report 34.6. Electrotechn. Congress, Moskow, Section
Metzer, A., E. Basevi, A.M. Baniel and Ch. 3B, Paper No. 10.
Gorin, 1974, US-Patent 3 796 793. Parker, R.S. and R.J. Studders, 1962, Per-
Miller, R.J., 1970, thesis Ohio State University. manent Magnets and Their Applications
Mondin, L.Ya., 1969, Poroshkovaya Metallur- (Wiley, New York, London).
gya, 77, 99-103. Passerone, A., E. Biagini and V. Lorenzelli,
Monteil, B., J.-C. Bernier and A. Revcolevschi, 1975, Ceramurgia Int. 1, No. 1, 23-27.
1977, Mat. Res. Bull. 12, 235-240. Pauthenet, R. and G. Rimet, 1959, Compt.
Monteil, J.B., L. Padel and J.C. Bernier, 1978, Rend. 249, 1875-1877.
J. Solid State Chem. 25, 1-8. Pawlek, F. and K. Reichel, 1957a, Arch.
Moon, D.W., J.M. Aitken, R.K. MacCrone and Eisenhfittenwes. 28, 241-244.
G.S. Cieloszyk, 1975, Phys. Chem. Glasses Pawlek, F. and K. Reichel, 1957b, Naturwiss.
16, 91-102. 44, 390.
Mountvala, A.J. and s.F. Ravitz, 1962, J. Petrdlik, M., V. Rubes, I.D. Radomysel'skij,
Amer. Ceram. Soc. 45, 285-288. A.F. Zhornyak and I.S. Nikishov, 1971, Sov.
Moskowitz, L.R., 1976, Permanent Magnet Powd. Metallurgy and Met. Ceram. 10, 516-
Design and Application Handbook (Cahners 520.
Books International, Boston/Mass.) available Petzi, F., 1971, Powd. Met. Int. 3, No. 4, 199-202.
from the Permanent Magnet Users Associa- Petzi, F., 1974a, Powd. Met. Int. 6, No. 3, 1-3.
tion of the Franklin Institute Research Petzi, F., 1974b, Keram. Zeitschrift, 26, No. 3,
Laboratories, 20th & Parkway, Philadelphia, 1-10.
PA. 19103, USA. Petzi, F., 1975, Ber. Dt. Keram. Ges. 52, No. 7,
Miiller, H.G. and G. Heimke, 1959, Ber. Arb. 249-251.
Gem. Magnetismus 1958 (Dr. Riederer~Ver- Petzi, F., 1980, Powd. Met. Int. 12, No. 1, 32-37.
lag, Stuttgart) 101-104. Pingault, D., 1974, Proc. 3rd Eur. Conf. on
N6el, L., 1951, J. Phys. Radium, 12, 339. Hard Magn. Mat., Amsterdam, 74-78.
H A R D FERRITES AND PLASTOFERRITES 599

Qian, X. and B.J. Evans, 1981, J. Appl. Phys. Ries, H.B., 1975b, Aufbereitungstechnik No.
52 (3), 2523-2525. 12.
Rademakers, A. and H. van Suchtelen, 1957, Roos, W., 1979, Thesis, Aachen.
Matronics, Nr. 12, 205-215. Roos, W., 1980, J. Amer. Ceram. Soc. 63, No.
Rathenau, G.W., 1953, Rev. Mod. Phys. 25, 11-12, 601-603.
29%301. Roos, W., H. Haak, C. Voigt and K.A. Hem
Rathenau, G.W., 1974, Proc. 3rd. Eur. Conf. on pel, 1977, J. de Phys. Colloque C1, Suppl. No.
Hard Magn. Mat,, Amsterdam, %16 4, 38 (C1) 35.37.
Rathenau, G.W., J. Smit and A.L. Stuyts, 1952, Roos, W., C. Voigt, H. Dederichs and K.A.
Z. Phys. 133, 250-260. Hempel, 1980, J. Magn. Magn. Mat. 15-18,
Ratnam, D.V. and W.R. Buessem, 1970, IEEE 1455-1456.
Trans. Magn. MAG-6, No. 3, 610-614. Rosin, P. and E. Rammler, 1933, Zement, 23,
Ratnam, D.V. and W.R. Buessem, 1972, J. 427-433.
Appl. Phys. 43, 1291-1293. Routil, R.J. and D. Barham, 1969, Canad. J.
Reddy, B.P.N. and P.J. Reddy, 1974, Phil. Mag. Chem. 47, 3919-3920.
30, No. 3, 8th series, 55%563. Routil, R.J. and D. Barham, 1971, J. Canad.
Reed, J.S. and R.M. Fulrath, 1973, J. Amer. Ceram. Soc. 40, 1-7.
Ceram. Soc. 56, 20%211. Routil, R.J. and D. Barham, 1974, Canad. J.
Reed, J.S. and F.J. Klug, 1975, Proc. US-Jap. Chem. 52, 3235-3246.
Seminar Bas. Sci. Ceram., Hakone, Japan, Rupprecht, J. and C. Heck, 1959, Ber. Arb.
189-193. Gem. Ferromagn. of 1958, 98-100.
Reichel, K., 1980, Praktikum der Magnettech- Ruthner, M.J., 1977, J. de Phys. 38, Coil, C 1,
nik (Franzis-Verlag, M/inchen). Suppl. au no. 4, C 1-311-315.
Reinboth, H., 1970, Technologie und Anwen- Ruthner, M.J., 1979, Sci. of Sintering, 11, No. 3,
dung magnetischer Werkstoffe, 3rd ed. (VEB 203-214.
Verlag Technik, Berlin). Ruthner, M.J., 1980, Ferrites, Proc. ICF 3, 64-
Remmey Jr., G.B., 1970, Powd. Met. Int. 2, 67.
12%129. Ruthner, M.J., H.G. Richter and I.L. Steiner,
Richter, H., 1962, Diploma work, Clausthal. 1970, Ferrites, Proc. Intern, Conf. Japan, 75-
Richter, H.G., 1968, DEW-Techn. Bet. 8, 192- 78.
208. Sadler, A.G., 1965, J. Canad. Ceram. Soc. 34,
Richter, H.G. and H.E. Dietrich, 1968a, IEEE 155-162.
Trans. Magn. MAG-4, 263-267. Sadler, A.G., W.D. Westwood and D.C. Lewis,
Richter, H. and H. V611er, 1968b, DEW-Techn. 1964, J. Canad. Ceram. Soc. 33, 12%137.
Ber. 8, 214-221. Saito, K., F. Hashimoto, M. Okuda and M.
Ries, H.B., 1959, Ber. Dr. Keram. Ges. 36, Torii, 1981, IEEE Trans. Magn. MAG-17,
223-228. No. 6, 2656-2658.
Ries, H.B., 1963, Angew. Mess- u. Regeltech- Samow, G., 1973, Elektroanzeiger, 26, No. 22,
nik, 3, No. 1, al-a15. 453455.
Ries, H.B., 1966, Interceram. No. 1, 84-92 and Schat, B.R. and H.J. Engel, 1970, Proc. Brit.
No. 2, 177-179. Ceram. Soc., No. 18, 281-292.
Ries, H.B., 1969a, Aufbereitungstechnik No. 1. Schieber, M., 1967, Experimental Magneto-
Ries, H.B., 1969b, Keram. Zeitschr. 21, No. 10., chemistry ed., E.P. Wohlfarth (North-Hol-
664-666. land, Amsterdam) p. 202.
Ries, H.B., 1970, Aufbereitungstechnik No. 3, Schinkmann, A., 1960, Hermsdorfer Technische
5, 10 and 12. Mitteilungen, 1, 42-49.
Ries, H.B., 1971a, Keram. Zeitschr. 23, No. 9, Schnettler, F.J., F.R. Montforte and W.H.
516-518, 520, 533-534 and No. 10, 591-597. Rhodes 1968, Sci. Ceramics 4, 79-90.
Ries, H.B., 1971b, Aufbereitungstechnik No. Sch6ps, W., 1979, Silikattechnik, 30, No. 7,
11. 195-201.
Ries, H.B., 1973, Interceram. No. 3, 207-210 Sch/Sps, W. and H. Beer, 1977, Silikattechnik
and No. 4, 298-304. 28, No. 7, 208-210.
Ries, H.B., 1975a, Keramische Zeitschrift, 27, Sch6ps, W., H. Beer and H. Gottwald, 1976,
No. 1 and 2. Silikattechnik, 235-239.
600 H. STJ~BLEIN

Schiller, K., 1965, DEW-Techn. Ber. 5, 64-73. Stiiblein, H., 1968a, Techn. Mitt. Krupp,
Schiller, K., 1967, Z. Angew. Phys. 22, 481-484. Forsch.-Ber. 26, 81-87.
Schiller, K., 1968, DEW-Techn. Ber. 8, 147-156. St/iblein, H., 1968b, Techn. Mitt. Krupp,
Schiller, K., 1973, Int. J. Magnetism, 5, 249- Forsch.-Ber. 26, No. 1, 1-10.
250. Stiiblein, H., 1970, IEEE Trans. Magn. MAG-6,
Schiller, K. and K. Brinkmann, 1970, Dauer- No. 2, 172-177, and Techn. Mitt. Krupp,
magnete, Werkstoffe und Anwendungen Forsch.-Ber. 28, No. 3/4, 103-116.
(Springer, Berlin). St~iblein, H., 1971, unpublished.
Schtitz, H.E. and H.W. Hennicke, 1978, Ber. St/iblein, H., 1972, AIP Conf. Proc. No. 5, Part
Dt. Keram. Ges. 55, 308--311. 2, New York, 950-969.
Schwabe, E., 1957, Z. Angew. Phys. 9, 183-187. St~iblein, H., 1973, Z. Werkstofftechnik, 4, 133-
Schwabe, E., 1958, Feinwerktechn. 62, 1-8. 142.
Schwabe, E., 1959, in: Ber. Arbeitsgem. Fer- Stilblein, H., 1974, unpublished.
romagn. 1958, Stuttgart, 74-80. St/iblein, H., 1975, Techn. Mitt. Krupp, Forsch.-
Semiletowa, M.W. and D.N. Polubojarinow, Ber. 33, 1-10.
1971, Mosk. Chim.-Technol. Inst. im. D.I. Stiiblein, H., 1978, Ber. Dt. Keram. Ges. 55,
Mendeleewa Trudi, 68, 136-139. 305-307.
Shimizu, S. and K. Fukami, 1972, J. Jap. Soc. Stilblein, H. and J. Willbrand, 1966a, IEEE
Powd. Met. 18, No. 7, 259-265. Trans. Magn. MAG-2, No. 3, 459--463.
Shirk, B.T. and W.R. Buessem, 1970, J. Amer. St~iblein, H. and J. Willbrand, 1966b, Z. angew.
Ceram. Soc. 53, 192-196. Phys. 21, 47-51.
Shtrikman, S. and D. Treves, 1960, J. Appl. Stiiblein, H. and W. May, 1969, Ber. Dt.
Phys. 31, 58 S-66 S. Keram. Gesellsch. 46, 69-74 and 126-128.
Silber, L.M., E. Tsantes and P. Angelo, 1967, J. Stiiblein, H. and J. Willbrand, 1971, Z. angew.
Appl. Phys. 38, 5315-5318. Phys. 32, 70-74.
Sironi, G., G. Fagherazzi, F. Ferrero and G. Stiiblein, H. and J. Willbrand, 1972, Proc. 7.
Parrini, 1972, DE-OS 2 246 204. Int. Symp. React. Sol. (Chapman and Hall) p.
Sixtus, K.J., K.J. Kronenberg and R.K. Tenzer, 589-597.
1956, J. Appl. Phys. 27, 1051-1057. Stiiblein, H. and J. Willbrand, 1973a, Science of
Sloccari, G., 1973, J. Amer. Ceram. Soc. 56, Ceramics, 6, Proc. Int. Conf., Baden-Baden
489-490. 1971, XXXII/1-12.
Sloccari, G. and E. Lucchini, 1977a, Ceramurgia Stiiblein, H. and J. Willbrand, 1973b, Proc. Int.
Int. 3, 10-12. Conf. Magnetism (ICM) Vol. IV, p. 232-236.
Sloccari, G., E. Lucchini and G. Asti, 1977b, Stanley, D.A., L.Y. Sadler, D.R. Brooks and
Ceram. Int. 3, 79-80. M.A. Schwartz, 1974, Ceram. Bull. 53, 813-815
Slokar, G. and E. Lucchini, 1978a, J. Mag. Magn. and 829.
Mat. 8, 232-236. Stearns, F.S. and H.L. Glass, 1975, Mat. Res.
Slokar, G. and E. Lucchini, 1978b, J. Mag. Bull. 10, 1255-1258.
Magn. Mat. 8, 237-239. Steingroever, E., 1966, J. Appl. Phys. 37, 1116--
Smit, J. and H.G. Beljers, 1955, Phil. Res. 1117.
Rep. 10, 113-130. Steinort, E., 1973, 1. IOS-Kolloquiun (Varese,
Smit, J. and H.P.J. Wijn, 1959, Ferrites, Phil. Italy).
Techn. Lab., Eindhoven. Steinort, E., 1974, Proc. 3rd Eur. Conf. on
Snoek, J.L., 1947, New Developments in Fer- Hard Magn. Mat., Amsterdam, 66--69.
romagnetic Materials (Elsevier Publishing Stephens, R.A., 1959, Amer. Ceram. Soc. Bull.
Comp., New York, Amsterdam). 38, 106-109.
Stiiblein, F., 1934, Techn. Mitt. Krupp 127-128. Stickforth, J., 1975, Techn. Mitt. Krupp,
Stfiblein, H., 1957, Techn. Mitt. Krupp 15, 165- Forsch.-Ber. 33, 22-24.
168. Stoner, E.C., E.W. Wohlfarth, 1948, Phil.
St~iblein, H., 1963, Techn. Mitt. Krupp, Trans. Roy. Soc. London, Ser. A, 240,599-642.
Werksber. 21, 171-184. Strijbos, S., 1973, Chemical Engineering
St~iblein, H., 1965, unpublished. Science, 28, 205-213.
Stfiblein, H., 1966, Techn. Mitt. Krupp, Forsch.- Strijbos, S., 1974, Proc. 3rd Eur. Conf. on Hard
Ber. 24, 103-112. Magn. Mat., Amsterdam, 102-105.
H A R D FERRITES AND PLASTOFERRITES 601

Stuijts, A.L., 1956, Trans. Brit. Ceram. Soc. 55, Tenzer, R.K., 1965, J. Appl. Phys. 36, No. 3,
57-74. 1180-1181.
Stuijts, A.L., 1968, Ch. 19 in: Ceramic Micro- Tokar, M., 1969, J. Amer. Ceram. Soc. 52,
structures, eds., R.M. Fulrath and J.A. Pask 302-306.
(Wiley, New York) 443-474. Torii, M., 1981, J. Jap. Soc. Pow. Met. 28, No. 3,
Stuijts, A.L., 1970, Ferrites, Proc. Int. Conf. 83-89.
Kyoto, 108-113. Torii, M., H. Kobayashi, F. Hashimoto and K.
Stuijts, A.L., 1973, Ann. Rev. Mat. Sci. 3, Saito, 1979, IEEE Trans. Magn. MAG-15,
363-395. No. 6, 1864-1866.
Stuijts, A.L., G.W. Rathenau and G.H. Weber, Torii, M., H. Kobayashi and M. Okuda, 1980,
1954, Phil. Techn. Rev. 16, 141-147. Ferrites, Proc. ICF 3, 370-374.
Stuijts, A.L., G.W. Rathenau and G.H. Weber, Townes, W.D., J.H. Fang and A.J. Perrotta,
1955, Phil. Techn. Rdsch. 16, 221-228. 1967, Z. Kristallogr. 125, 437-449.
Suchet, J., 1956, Bull. Soc. Franc. Ceram. 33, Tul'chinskii, L.N. and V.N. Pilyankevich, 1971,
33-43. Sov. Powd. Metallurgy and Met. Ceram. 10,
Sutarno and W.S. Bowman, 1967, Mines 359-361.
Branch Invest. Rep. IR 67-23 (Dept. Energy Ullmanns Encyklop/idie der techn. Chemie,
Mines and Resources, Ottawa, Canada). 1965, 3rd ed., Vol. 16 (Urban and
Sutarno, W.S. Bowman, G.E. Alexander and Schwarzenberg, Mfinchen, Berlin) p. 455.
J.D. Childs, 1969, J. Canad. Ceram. Soc. 38, Ullmanns Encyklop~idie, 1974, 4th ed., Vol. 8,
%13. 301-311.
Sutarno, W.S. Bowman and G.E. Alexander, Underhill (ed.), E.A., 1957, Permanent Magnet
1970a, J. Canad. Ceram. Soc. 39, 33-41. Handbook (Crucible Steel Comp., Pitts-
Sutarno, R.H. Lake and W.S. Bowman, 1970b, burgh, USA).
Mines Branch Res. Rept. R 223 (Dept. Valvo, 1978/1979, Handbuch Permanentmag-
Energy Mines and Resources, Ottawa, nete (FERROXDURE, FXD).
Canada). Van den Broek, C.A.M., 1974, Proc. 3rd Eur.
Sutarno, W.S. Bowman and G.E. Alexan- Conf. on Hard Magn. Mat., Amsterdam, 53-
der, 1970c, J. Canad. Ceram. Soc. 39, 61.
45-50. Van den Broek, C.A.M., 1977, Ceramurgia In-
Sutarno, W.S. Bowman and G.E. Alexander, ternational, 3, 115-121.
1971, J. Canad. Ceram. Soc. 40, 9-14. Van den Broek, C.A.M. and A.L. Stuijts
Syono, Y., A. Ito and O. Horie, 1979, J. Phys. 1977/78, Phil. Techn. Rdsch. 37, Nr. 7, 169-
Soc. Japan, 46, No. 3, 793-801. 188.
Takada, T. and M. Kiyama, 1970a, Proc. Int. Van der Oiessen, A.A., 1970, Klei en Keramik,
Conf. Ferrites, Japan, 69-71. 20, 30-38.
Takada, T., Y. Ikeda, H. Yosbinaga and Y. Van Hook, H.J., 1964, J. Amer. Ceram. Soc.
Bando, 1970a, Proc. Int. Conf., Ferrites, 47, 579-581.
Japan, 275-278. Van Hook, H.J., 1976, Phase Equilibria in
Tanasoiu, C., 1972, IEEE Trans. Magn., Sept. Magnetic Oxide Materials, in: Phase
348-351. Diagrams, Vol. IV, ed., A.M. Alper
Tanasoiu, C., P. Nicolau, C. Miclea and E. (Academic Press, New York, London).
Varzaru, 1976a, J. Magn. Magn. Mat. 3, Van ,Tendeloo, G., D. van Dyck, J. van Lan-
275-280. duyt and S. Amelinckx, 1979, J. Solid State
Tanasoiu, C., P. Nicolau and C. Miclea, 1976b, Chem. 27, 55-70.
IEEE Trans. Magn. 1VIAG-12, No. 6, 980- Van Uitert, L.G., 1957, J. Appl. Phys. 28, 317-
982. 322.
Taylor, R.C. and V. Sadagopan, 1972, Solid Veldkamp, J.D.B. and R.J. Klein Wassink,
State Sci. and Techn. 119, no. 6, 788-790. 1976, Philips Res. Repts. 31, 153-
TDK, 1978, Data Handbook. 189.
Tenzer, R.K., 1957, Conf. Magnetism and Veldkamp, J.D.B. and N. Hattu, 1979, Phil. J.
Magn. Mat. Boston, 203-211. Res. 34, No. 1/2, 1-25.
Tenzer, R.K., 1963, J. Appl. Phys. 34, No. 4, Vogel, E.M., 1979, Ceramic Bull. 58, No. 4,
126%1268. 453-454, 458.
602 H. ST)kt3LEIN

Vogel, R.H. and B.J. Evans, 1979a, J. Mag- (suppl. of Phil. Mag.) 8, No. 30, 87-224.
netism Magn. Mat. 13, 294-300. Wolski, W. and J. Kowalewska, 1970, Jap. J.
Vogel, R. and B.J. Evans, 1979b, J. Phys. Col- Appl. Phys. 9, 711-715.
loqu, 40, No. C-2, pt. 3, C2-277 to C2-279. Wullkopf, H., 1972, Intern. J. Magnetism 3,
Voigt, C., 1969, Z. Angew. Phys. 26, 160-165. 179-187.
Voigt, C. and K.A. Hempel, 1969, Phys. Status Wullkopf, H., 1973, Intern. J. Magnetism 5,
Solidi 33, 249-256. 147-155.
Vollmershaus, E. and K.A. Hempel, 1975, Bet. Wullkopf, H., 1974, thesis, Bochum.
Dt. Keram. Ges. 52, 216-218. Wullkopf, H., 1978, Ber. Dt. Keram. Ges. 55,
Von Aulock, W.H. (ed.), 1965, Handbook of 292-293.
microwave ferrite materials (Academic Press, Yamamoto, H., T. Kawaguchi and M.
New York). Nagakura, 1978a, J. Jap. Soc. Powd. Met.
Von Basel, H.B., 1981, IEEE Trans. Magn. 25, No. 7, 236-241.
MAG-17, No. 6, 2654-2655. Yamamoto, H., T. Kawaguchi, M. Nagakura
Von Basel, H.B. and K.A. Hempel, 1979, Phys. and Y. Kobayashi, 1978b, J. Jap. Soc. Powd.
Status Solidi A, 55, K 183-K 184. Met. 25, No. 7, 242-248.
Went, J.J., G.W. Rathenau, E.W. Gorter and Yamamoto, H., T. Kawaguchi and M.
G.W. van Oosterhout, 1952, Phil. Techn. Nagakura, 1979a, IEEE Trans. Magn. MAG-
Rev. 13, 194-208. 15, No. 3, 1141-1146.
Wickham, D.G., 1970, Ferrites, Proc. Internat. Yamamoto,H. and Y. Kobayashi, 1979b, Res.
Conf. Japan, 105-107. Rept. Fac. Eng. Meiji Univ., Tokyo, No. 36,
Willbrand, J. and U. Wieland, 1975, Techn. 63-72.
Mitt. Krupp, Forsch.-Ber. 33, 15-21. Yamamoto, H., T. Kawaguchi and M.
Wills, D. and J. Masiulanis, 1976, J. Canad. Nagakura, 1980, J. Jap. Soc. Powd. Met. 27.
Ceram. Soc. 45, 15-19. No. 7, 171-177.
Wilson, C.M., G.C. Bye, C.R. Howard, J.H. Zfiv~ta, K., 1963, Phys. Status Solidi, 3, 2111-
Sharp, D.M. Tinsley and S.A. Wentworth- 2118.
Rossi, 1972, React. Sol. 7, Int. Symp. Ziolowski, Z., 1962, Prace Institut Hutniczych,
(Chapman and Hall) 598~o09. 14, 155-163.
Winkler, G., 1965, React. Sol. 5, Int. Symp. Zijlstra, H., 1974, Phil. Techn. Rev. 34, no. 8,
(Elsevier, Amsterdam) 572-582. 193-207.
Wippermann, A., 1968, thesis, Aachen. Zijlstra, H., 1978, IEEE Trans. Magn. MAG-14,
Wohlfarth, E., 1959, Advances in Physics no. 5, 661-664.
chapter 8

SULPHOSPINELS

R.P. VAN S T A P E L E
Philips Research Laboratories
5600 JA Eindhoven
The Netherlands

Ferromagnetic Materials, Vol. 3


Edited by E.P. Wohlfarth
© North-Holland Publishing Company, 1982

603
CONTENTS

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . 607
2. Crystal chemistry . . . . . . . . . . . . . . . . . . . . . . . . . 608
3. L o c a l i z e d a n d d e l o c a l i z e d s t a t e s in s u l p h o s p i n e l s . . . . . . . . . . . . . 616
4. Sulphospinels containing copper . . . . . . . . . . . . . . . . . . 618
4.1. V a l e n c y of t h e c o p p e r i o n s . . . . . . . . . . . . . . . . . . 618
4.2. CuTi2S4 . . . . . . . . . . . . . . . . . . . . . . . . . 62O
4.3. CuVzS4 . . . . . . . . . . . . . . . . . . . . . . . . . 622
4.4. CuCo2S4 a n d C u C o T i S 4 . . . . . . . . . . . . . . . . . . . 624
4.5. CuRh2S4 a n d CuRh2_xCoxS4 . . . . . . . . . . . . . . . . . . 626
4.6. C u R h 2 S e 4 a n d C u R h 2 - x S n x S e 4 . . . . . . . . . . . . . . . . . 627
4.7. CuCr2S4, CuCr2-xTixS4, C u C r 2 - x S n x S 4 a n d C u C r 2 - x V x S 4 . . . . . . . . 630
4.8. CuCr2Se4, C u C r 2 - x R h x S e 4 a n d CuCr0.3Rhl.7-xSnxSe4 . . . . . . . . . . 636
4.9. C u C r 2 T e 4 a n d Cul+~Cr2Te4 . . . . . . . . . . . . . . . . . . . 641
4.10. C u C r 2 ( X , X')4 w i t h X, X ' = S, Se a n d T e . . . . . . . . . . . . . . 643
4.11. CuCr2X4-xYx w i t h X = S, S e o r T e a n d Y = C1, B r a n d I . . . . . . . . 644
5. F e r r o m a g n e t i c a n d a n t i f e r r o m a g n e t i c s e m i c o n d u c t o r s . . . . . . . . . . . 647
5.1. G e n e r a l a s p e c t s . . . . . . . . . . . . . . . . . . . . . . 647
5.2. Z n f r 2 S 4 . . . . . . . . . . . . . . . . . . . . . . . . . 653
5.3. CdCrzS4 . . . . . . . . . . . . . . . . . . . . . . . . . 654
5.4. HgCr2S4 . . . . . . . . . . . . . . . . . . . . . . . . . 666
5.5. Z n C r 2 S e 4 . . . . . . . . . . . . . . . . . . . . . . . . . 669
5.6. CdCr2Se4 . . . . . . . . . . . . . . . . . . . . . . . . . 675
5.7. H g C r z S e 4 . . . . . . . . . . . . . . . . . . . . . . . . . 691
5.8. M i x e d c r y s t a l s b e t w e e n t h e c o m p o u n d s Z n C r z X 4 , CdCr2X4 a n d H g C r 2 X 4
w i t h X = S, S e . . . . . . . . . . . . . . . . . . . . . . . 694
5.9. M i x e d c r y s t a l s A172A3~2CrzX4 w i t h X = S, Se a n d d i a m a g n e t i c i o n s A . . . . 698
6. F e r r i m a g n e t i c s e m i c o n d u c t o r s . . . . . . . . . . . . . . . . . . . 701
6.1. I n t r o d u c t i o n . . . . . . . . . . . . . . . . . . . . . . . 701
6.2. MnCrzS4 . . . . . . . . . . . . . . . . . . . . . . . . . 701
6.3. FeCr2S4 . . . . . . . . . . . . . . . . . . . . . . . . . 706
6.4. CoCr2S4 . . . . . . . . . . . . . . . . . . . . . . . . . 711
6.5. T h e m i x e d c r y s t a l s Fea-xCoxCrzS4, Fel-x(CUl/zlnl/2)xCrzS4, Fel-xCdxCrzS4,
Col-xfdxfr284 and Col-x(CumFem)xCr2S4 . . . . . . . . . . . . . 714
6.6. T h e m i x e d c r y s t a l s M l - x C u x C r 2 S 4 w i t h M = M n , F e a n d C o . . . . . . . 718
6.7. T h e m i x e d c r y s t a l s MI-xNi~Cr2S4 w i t h M = M n , Fe, C o , C u a n d Z n . . . . 721
6.8. T h e m i x e d c r y s t a l s MCr2-xInxS4, w i t h M = M n , F e , C o a n d Ni . . . . . . 722
6.9. T h e m i x e d c r y s t a l s M n C r z - x V ~ S 4 . . . . . . . . . . . . . . . . 725

604
6.10. The mixed crystals FeCr2-xFe~S4 . . . . . . . . . . . . . . . . 726
6.11. The mixed crystals MCr2S4-xSex with M = Mn, Fe, Co or CUl/2Fem . . . . 726
7. Some rhodium and cobalt spinels . . . . . . . . . . . . . . . . . . 728
7.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . 728
7.2. CoRh2S4, Col-xFexRh2S4 and FeRh2S4 . . . . . . . . . . . . . . . 728
7.3. The mixed crystals FeRh2-xCrxS4, CoRh2-xCr~S4 and NiRh2-xCrxS4 . . . . . 730
7.4. The mixed crystals Fel-xCuxRh2S4 and C01-xCuxRh2S4 . . . . . . . . . 732
7.5. Co3S4 and NiCo2S4 . . . . . . . . . . . . . . . . . . . . . . 736
Notes added in proof . . . . . . . . . . . . . . . . . . . . . . . 737
References . . . . . . . . . . . . . . . . . . . . . . . . . . . 737

605
1. Introduction

The denomination of "sulphospinels" will be used for materials with compositions


(M, M', M")3X4 where M, M', M " . . . are metals and X = S, Se or Te. In contrast to
the "oxyspinels" (X = O) the sulphospinels have not found applications. They
exhibit however a much larger variety of physical properties, which makes them
interesting from a scientific point of view. Whereas the oxyspinels are in general
semiconductors with antiferromagnetic interactions, the sulphospine!s exhibit
metallic conduction and superconduction as well as semiconductivity, and ferro-
magnetic as well as antiferromagnetic interactions.
At the time that the first magnetic measurements on sulphospinels were carried
out (Lotgering 1956) the oxyspinels had already been investigated for about ten
years in many laboratories. The stimulation for an intensive study of sulphospinels
came later, however, with the observation that CuCr2X4 (X = S, Se or Te) are
metallic ferromagnets with Curie temperatures above room temperature (Lotger-
ing 1964a, b). The subsequent discovery that MCr2X4 (M = Cd or Hg, X = S or Se)
are semiconducting ferromagnets (Baltzer et al. 1965, Menyuk et al. 1966) with
Curie temperatures not far below room temperature started numerous in-
vestigations of the physical and chemical properties, which resulted in an increas-
ing number of articles on sulphospinels. After peak years in the early seventies,
interest has been fading away. It may therefore be useful to review the results that
have been obtained. The properties of magnetic semiconductors have earlier been
summarized by Haas (1970), Methfessel and Mattis (1968), Nagaev (1975), and
Wojtowicz (1969).
Among the results of the study of sulphospinels the following are worth
mentioning in advance:
(1) The fact that superexchange interactions between transition metal ions can
be ferromagnetic in semiconductors had been predicted theoretically and CrTe
was considered to be an example (Anderson 1950). However, this compound,
which was the only example known at that time, shows metallic conduction. The
first indication of ferromagnetic superexchange in a real semiconductor was found
in a sulphospinel, namely in MnCr2S4 (Lotgering 1956).
(2) In magnetic semiconductors electrical and optical properties were observed
that strongly depend on the magnetic state. These included an anomalous
maximum in the resistivity near the Curie temperature, first observed in FeCr2S4

607
608 R.P. VAN STAPELE

(Lotgering 1956); a negative magnetoresistance effect near the Curie temperature,


first observed in n-type CdCr2Se4 (Lehmann and Harbeke 1967), and an
anomalous shift to the red of the optical absorption edge of CdCr2Se4 (Busch et al.
1966, Harbeke and Pinch 1966).
(3) As expected theoretically, an overlap of the d-orbitals of two magnetic ions
may contribute to the exchange interaction (Wollan 1960, Goodenough 1960).
Experimental evidence of this direct exchange was obtained from the asymptotic
Curie temperature of ZnCr2X4 with X = O, S or Se (Lotgering 1964a, b).
(4) Long-distance superexchange was observed in sulphospinels. A strong
interaction via two sulphur ions occurs in CoRh2S4 (Blasse and Schipper 1964) and
a fairly strong interaction via four sulphur ions occurs in Fel/2Cul/2Rh2S4 (Plumier
and Lotgering 1970).
(5) It was discovered that CuCr2X4 (X = S, Se or Te) are metallic ferromagnets
in which the chromium spins are coupled ferromagnetically via interaction with
delocalized conduction electrons. This provides an experimental confirmation of
Zener's basic condition for strong ferromagnetism (Zener 1951) in simple, highly
symmetric compounds.
The chapter will be organized as follows. After a section (2) on the crystal
structure and a section (3) on localized and delocalized electronic states, we will
discuss in section 4 the metallic sulphospinels containing copper. In section 5 we
deal with the ferromagnetic and antiferromagnetic semiconductors, in section 6
with the ferrimagnetic semiconductors and finally, in section 7 we consider some
rhodium and cobalt spinels.

2. Crystal chemistry

Compounds with the general formula AB2X4, where A and B are metal ions and
X = S, Se or Te, crystallize in a large number of crystal structures. It is not
possible to calculate the relative stability of the various structures for a given
compound, but it has been found possible to define parameters that place
different crystal structures in different regions of the parameter space. For that
purpose Kugimiya and Steinfink (1968) used the radius ratio rA/rB of the cations A
and B and a bond-stretching force constant KAB that is proportional to the
product of the cation electronegativities and the inverse of the square of a suitably
defined equilibrium distance. In the K ~ versus rA/rB plot the observed crystal
structures separate nicely, with the exception of the spinel, the Cr3Se4 and the
Ag2Hgh structure (Iglesias and Steinfink 1973). The difficulty in distinguishing
between the spinel and the Cr3Se4 structure is reflected in the occurrence of both
structures in one compound at different temperatures and pressures. High-
pressure polymorphism in spinel compounds was first observed by Albers and
Rooymans (1965), who succeeded in changing the spinel structure of FeCr2S4 into
a structure related to Cr3Se4. Other examples have been found among the
sulphides (Bouchard 1967, Tressler and Stubican 1968, Tressler et al. 1968), but
not unambiguously among the selenides and the tellurides. Being concerned with
SULPHOSPINELS 609

the spinel structure only, we will not discuss these matters here. We conclude this
passage with the remark that at room temperature and under normal pressures
the number of AB2X4 compounds with the spinel structure decreases strongly in
the sequence S, Se and Te. The only tellurides reported are CuCrzTe4 (Hahn and
Schr6der 1952) and ZnMnzTe4 (Matsumoto et al. 1966).
Since a description of the spinel structure has been given in chapter 4 by
Krupi6ka and Novfik, we shall confine ourselves here to the main details of this
structure (an extensive description was given by Gorter (1954)). The space group
of the spinel structure is Fd3m. The chalcogenide anions approximate to a
close-packed cubic lattice. The cations occupy twice as many octahedral sites B as
tetrahedral sites A. The A sites form a diamond lattice and can be divided into
two fcc Bravais lattices. Their local symmetry is purely tetrahedral (point group
Td). The B sites are divided into four Bravais lattices. Their point group is D3a.
The smallest unit cell is rhombohedral and contains two molecules AB2X4 (fig. 1).
More commonly used is the cubic unit cell (fig. 2) which contains eight molecules.
The structure is completely described by the cubic cell edge a and one parameter
u that fixes the X positions. A deviation from the ideal close packing is caused by
a variable A - X distance in one of the four [111] directions. The parameter u is
defined by the shift a6X/3 with 6 = u - 3 of X from the ideal position (drawn in fig.
2) away from the A site. The distances A - X and B - X are then:
(A-X) = aV'5(6 + ~),
(B-X) = a ( ½ - ½6 + 362) m . (1)

Each X belongs to the (perfect) tetrahedron of one A. With u # 3 two distances


(X-X)1 for X of one tetrahedron and (X-X)2 for X of two neighbouring tetrahedra
exist:
(X-X)1 = 2aX/2(-~+ 6); (X-X)2 = 2aX/2(~- ~). (2)

Fig. 1. T h e primitive rhombohedral unit cell which contains two AB2X4 units in the cubic cell.
610 R.P. V A N S T A P E L E

Fig. 2. The cubic unit cell of the spinel structure with cell edge a. The structure can be described using
two types of cubic octants (edge = a/2) that alternate like Na and C1 in rocksalt. Shaded circles A lying
on the corner and in the centre of an octant. Black circles B and white circles X (for u = ~) lying on the
body diagonals of the octants at ~ of its length.

With increasing u at constant a, A - X increases and B - X decreases. This gives


an accommodation of the lattice to the radii of the ions. For the ideal close
packing (u = 3) B - X is 16% larger than A - X , which is compensated by u > 3 in all
spinels. Large values of u occur for large A ions like, for instance, Cd 2+.
The octahedral sites are denoted by brackets. For two metals M and N the
distributions M[Nz]X4 and N[NM]X4 among A and B sites are called "normal"
and "inverse", respectively. All sulphospinels with composition MN2X4 are
normal. On the octahedral sites we will encounter the ions Ti, V, Cr, Co, In, Rh
and Sn, and on the tetrahedral sites the ions Mn, Fe, Co, Cu, Zn, Cd and Hg.
Table i lists u and a. Using these data a plot of u versus a (Raccah et al. 1966,

TABLE 1
Lattice parameters at room temperature as determined by X-ray diffraction ((n) denotes results of
neutron diffraction).

Compound a (A) u References

Cu/Ti2/S4 9.88 ± 0.008 0.382 Hahn and Harder (1956)


9.994 ± 0.002 Bouchard et al. (1965)
9.994 0.381 Riedel and Horvath (1973a)
10.002 ± 0.003 (0.3805) Le Nagard et al. (1975)
Cu/V2/S4 9.824 ± 0.008 0.384 Hahn et al. (1956)
9.808 ± 0.002 Bouchard et al. (1965)
9.803 0.382 Riedel and Horvath (1973a)
9.805 ± 0.005 Le Nagard et al. (1979)
9.800 ± 0.001 ibid.
SULPHOSPINELS 611

T A B L E 1 (continued)

Compound a (A) References

Mn/Cr2/S4 10.129 Lotgering (1956)


10.110 ± 0.002 Bouchard et al. (1965)
10.110 0.3876 (n) Menyuk et al. (1965)
10.110 0.3863 Raccah et al. (1966)
10.107 ± 0.002 Tressler and Stubican (1968)
10.11 Robbins et al. (1974)
10.108 Darcy et al. (1968)

Fe/Crz/S4 9.97 ± 0.01 Hahn (1951)


9.998 Lotgering (1956)
9.995 ± 0.002 Bouchard et al. (1965)
9.995 0.384 ± 0.002 (n) Shirane et al. (1964)
9.995 0.3850 Raccah et al. (1966)
9.983 ± 0.001 Tressler and Stubican (1968)
(9.97) 0.3858 (n) Broq. Colominas et al. (1964)
9.995 Robbins et al. (1970b, 1974)
9.9893 ± 0.0008 Shick and Von Neida (1969)

Co/Cr2/S4 9.91±0.01 Hahn (1951)


9.934 Lotgering (1956)
I 9.923 ± 0.002 Bouchard et al. (1965)
9.923 0.3821 Raccah et al. (1966)
0.3830(n) Raccah et al. (1966)
9.923 ±0.001 Tressler and Stubican (1968)
9.9213 ± 0.0007 Carnall et al. (1972)
9.936 Robbins et al. (1974)
9.923 Lutz and Becker (1973)
9.918 Lisnyak and Lichter (1969)
9.9158± 0.0007 Shick and Von Neida (1969)

Cu/Cr2/84 9.63±0.006 0.381 Hahn et al. (1956)


9.822 Lotgering (1964b)
9.814 ± 0.002 Bouchard et al. (1965)
9.810 0.383 ± 0.001 Riedel and Horvath (1973a)
9.814 0.3841 Raccah et al. (1966)
9.833 ± 0.019 Kanomata et al. (1970)
9.820 Lutz and Becker (1973)
9.820 Robbins et al. (1970a)
9.813 Belov et al. (1973)
0.384 Ohbayashi et al. (1968)
0.385 Riedel and Horvath (1973b)

Zn/Cr2/S4 9.983 Lotgering (1956)


9.988 0.385 Baltzer et al. (1965)
9.986 ± 0.002 Bouchard et al. (1965)
9.986 0.3842 Raccah et al. (1966)
9.986± 0.001 Von Neida and Shick (1969)
9.983 0.3854 ± 0.0005 Riede! and Horvath (1973a)
9.983 0.3869 Riedel and Horvath (1969)
612 R.P. V A N S T A P E L E

T A B L E 1 (continued)

Compound a (~,) u References

Cd/Cr2/S4 10.244 0.390 Baltzer et al. (1965, 1966)


10.242 Busch et al. (1966)
10.239 +- 0.002 Von Neida and Shick (1969)
10.238 0.3901 + 0.0007 Riedel and Horvath (1973a)
10.238 0.3943 Riedel and Horvath (1969)
Hg/Cr2/S4 10.206-+ 0.007 Hahn (1951)
10.237 0.390 Baltzer et al. (1965)
Co/Coz/S4 9.416 Lotgering (1956)
9.399 + 0.002 Bouchard et al. (1965)
9.391 -+ 0.004 Heidelberg et al. (1966)
9.405 Knop et al. (1968)
Ni/Co2/S4 9.392 Lotgering (1956)
9.384 + 0.002 Bouchard et al. (1965)
9.387 Knop et al. (1968)
Cu/Co2/S4 9.482 Lotgering (1956)
9.461 _+0.002 Bouchard et al. (1965)
9.464 0.382 Riedel and Horvath (1973a)
9.478 -+_0.001 0.388 - 0.001 Williamson and Grimes (1974)
Co/Rh2/S4 9.72 Koerts (1963)
9.74 Blasse (1965)
9.805 0.383 Kondo (1976)
9.76 Blasse and Schipper (1964)
Fe/Rh2/S4 9.86 Koerts (1%5)
9.902 0.385 Kondo (1976)
Ni/Rh2/S4 9.64 Koerts (1963)
9.701 -+ 0.001 Tressler et al. (1968)
9.702 Itoh (1979)
Cu/Rh2/S4 9.72 Blasse and Schipper (1964)
9.78 Van Maaren et al. (1967)
9.786 0.3802 -+ 0.0004 Riedel and Horvath (1973a)
9.792-9.787 Shelton et al. (1976)
9.790-+ 0.001 Schaeffer and Van Maaren (1968)
9.7877 -+ 0.0005 0.384-+ 0.001 Dawes and Grimes (1975)
9.790 Robbins et al. (1967a)
9.788 0.384 Riedel et al. (1976)
9.73 Koerts (1963)
Cu/Cr2/Se4 10.356-+ 0.006 0.380 Hahn et al. (1956)
10.337 Lotgering (1964b)
10.334 0.384 (n) Robbins et al. (1967b)
0.3826 -+ 0.0003 (n) Colominas (1967)
10.321 -+ 0.014 Kanomata et al. (1970)
10.334 Belov et al. (1973)
0.383 Ohbayashi et al. (1968)
0.385 Riedel and Horvath (1973b)
SULPHOSPINELS 613

T A B L E 1 (continued)

Compound a (.~) u References

Zn/Cr2/Se4 10.443 _+0.008 0.376 ~< u ~<0.380 Hahn and Schr6der (1952)
10.500 Lotgering (1964b)
10.44 0.3843 ± 0.0006 (n) Plumier (1965)
10.443 Baltzer et al. (1965)
10.4973 Kleinberger and de Kouchkovsky
(1966)
10.440 0.3849 Raccah et al. (1966)
10.495 Busch et al. (1966)
10.494 ± 0.001 Von Neida and Schick (1969)
10.493 0.3843 Riedel and Horvath (1969)
Cd/Cr2/Se4 10.721 ± 0.008 0.383 Hahn and Schr6der (1952)
10.755 0.390 Baltzer et al. (1965)
10.740 0.3894 Raccah et al. (1966)
10.744 Busch et al. (1966)
10.72 ± 0.01 V o n Neida and Shick (1969)
10.741 0.3889 Riedel and Horvath (1969)
Hg/Cr2/Se4 10.753 0.390 Baltzer et al. (1965)

Cu/Rhz/Se4 10.259 (n) Robbins et al. (1967b)


10.34 Van Maaren et al. (1967)
10.260 Shelton et al. (1976)
10.2603 ± 0.0004 0.384 ± 0.001 D a w e s and Grimes (1975)
10.263 Robbins et al. (1967a)
10.264 0.379 Riedel et al. (1976)
Cu/Cr2/Te4 11.137 Lotgering (1964b)
0.3811 _+0.0003 (n) Colominas (1967)
11.127 Kanomata et al. (1970)
11.049 ± 0.007 0.379 Hahn et al. (1956)
0.376 Riedel and Horvath (1973b)
11.134 Robbins et al. (1968)

Riedel and Horvath 1973a) for MCr2S4 is given in fig. 3. Setting aside M = Cu, M has
the sequence of increasing radius of the divalent ions. The figure illustrates the
increase of u (see above) and of a with the M 2+ radius. This also appears directly
from the increase of the distance M - S calculated from u and a using eq. (1) (table
2).
Figure 3 shows that the Cu c o m p o u n d behaves differently from the others. This
appears also from the Cr-S = B - X distances (table 2), which are 2.40 to 2.42
and are practically independent of M. For M = Cu, however, Cr-S is slightly
smaller (2.38A). A n anomalously small Cr-Se distance occurs also in the
selenides M[Cr2]Se4 for M = Cu: Cr-Se = 2.51 to 2.54 A for M = Zn, Cd or Hg,
whereas Cr-Se = 2.49 to 2.50 A for M = Cu.
S o m e sulphospinels, to which CuCr2S4 belongs, have an anomalously small a
(Lotgering 1956). (This is the reason for the anomalous position of M = Cu in fig.
614 R.P. VAN STAPELE

Cd/
0 39o
/Hg
0388 /
0386 Zn /0Mo
0.384

0382 /l co
/
0.380 r / i i I
97 98 " 99 10 10.1 10.2 10.3
a (~}
Fig. 3. Lattice parameter a versus u parameter of the compounds MCr2S4.

3.) Using eq. (2) one finds am = 7.47 and 9.84 A for the cell edges of ideal spinel
(u = 3) with X - X distances equal to the sum of the ionic X 2- radii of 02_ (1.32 A)
and S 2- (1.74 A), respectively. According to the definition of this radius, a ~> am
for ionic compounds, and this is indeed observed for oxyspinels (the smallest
value a =-7.94 A occurs for LiomAls/204 (Blasse 1964). I n sulphides with spinel
structure, a runs from 10.24A for CdCr2S4 to 9 . 3 9 A for NiCo2S4 (table 1).
a < 9.84,~ means that the mean S-S distance is smaller than the shortest $2--S 2-
distance, i.e., the normal ionic S 2- state does not occur. We will see later that this is
reflected in the electric and magnetic properties, which are essentially different from
those of oxyspinels (and other oxides) for a < 9.84 A, but not for a > 9.84,~. For
example (see table 1 and section 5.2) Zner204 and ZnCr2S4 (a = 9.99 A) are
semiconductors and exhibit Curie constants corresponding to the metal ions (Zn 2+,
Cr3+). This holds also for C o 3 0 4 (Wagner 1935, Cossee 1958) but C o 3 S 4 is metallic and
the Curie constant does not agree at all with ionic C02+Co~+X2-, in contrast to CosO4.
Since ions are not rigid spheres, the ionic S 2- radius and thus am are not determined
exactly, so that the rule does not hold for a near oto am (e.g. CoRh2S4 with a = 9.80
has normal ionic properties). CuTi2S4 (a = 9.99 A), which is Pauli paramagnetic and
metallic, is the only clear exception. Consequently the metallic conduction has
another origin than in the o t h e r sulphospinels (section 4.2).
A comparison of the A - X distances in the compounds CuM2S4 of table o 2
demonstrates that this distance is almost constant with an average value of 2.25 A,
close to the value 2.26 A obtained from a more extensive analysis by Riedel o
and
H o r v a t h (1973a). In the selenospinets CuM2Se4
o
the average value is 2.38 A and in
the tellurospinel the C u - T e distance is 2.53 A. These distances compare well with
the C u - X distances in other chalcogenides (Lotgering and Van Stapele 1968b).
T h e A - X distances in MCr2X4 with M = Zn, Cd and Hg and X = S and Se
compare well with the tetrahedral covalent radii as derived by Pauling (1960) from
the sphalerite and wurtzite structures of the compounds M X (table 3).
TABLE 2
Cation-anion distances

A-X B-X
Compound (~) (~) References

CoCr2S4 2.27 2A1 Raccah et al. (1966)


FeCr2S4 2.32 2.41 Shirane et al. (1964)
2.34 2.40 Raccah et al. (1966)
ZnCr2S4 2.34 2.40 Baltzer et al. (1965)
2.32 2.41 Raccah et al. (1966)
2.34 2.40 Riedel and Horvath (1973a, 1969)
MnCrzS4 2.39 2.42 Raccah et al. (1966)
2.41 2.41 Menyuk et al. (1965)
CdCr2S4 2.48 2.42 Baltzer et al. (1965)
2.48 2.42 Riedel and Horvath (1973a, 1969)
HgCrzS4 2.48 2.42 Baltzer et al. (1%5)
CoRh2S 4 2.26 2.38 Kondo (1976)
CuCr2S4 2.26 2.38 Riedel and Horvath (1973a)
2.28 2.37 Raccah et al. (1966)
CuTizS4 2.27 2.44 Riedel and Horvath (1973a)
2.26 2.44 Le Nagard et al. (1975)
CuVzS4 2.24 2.38 Riedel and Horvath (1973a)
CuCo2S4 2.16 2.30 Riedel and Horvath (1973a)
2.26 2.25 Williamson and Grimes (1974)
CuRh2S4 2.27 2.36 Dawes and Grimes (1975)
2.23 2.39 Riedel et al. (1976)
CuCr2Se4 2.38 2.50 Riedel and Horvath (1973b)
2.40 2.49 Robbins et al. (1967b)
ZnCr2Se4 2.43 2.52 Plumier (1965)
2.44 2.51 Raccah et al. (1966)
2.46 2.52 Riedel and Horvath (1969)
CdCrzSe4 2.61 2.54 Baltzer et al. (1965)
2.59 2.54 Raccah et al. (1966)
2.61 2.53 Riedel and Horvath (1969)
HgCr2Se4 2.61 2.54 Baltzer et al. (1965)
CuRhzSe4 2.38 2.48 Dawes and Grimes (1975)
2.34 2.50 Riedel et al. (1976)
CuCr2Te4 2.53 2.72 Riedel and Horvath (1973b)

TABLE 3
Cation-anion distance on tetrahedral sites in the
compounds MCr2X4, compared with the distance in
the compounds M X

A-X M-X
Compound (A) compound M X (A)

ZnCr2S4 2.33 ZnS 2.35


CdCrzS4 2.48 CdS 2.53
HgCr2S4 2.48 HgS 2.52
ZnCr2Se4 2.44 ZnSe 2.45
CdCr2Se4 2.60 CdSe 2.63
HgCr2Se4 2.61 HgSe 2.63

615
616 R.P. V A N S T A P E L E

3. Localized and delocalized states in sulphospinels

Many metallic conducting chalcogenides of transition metals are known. In most


of them the magnetic moments do not agree with the spin-only value of magnetic
ions. For example, CrTe and MnSb, which might be compounds of Cr z+ and Mn 3+,
are metallic ferromagnets. The experimental ferromagnetic moments of 2.4 and
3.5 txB and the paramagnetic moments of 4.0 and 4.1 #B, respectively (Lotgering
and G o r t e r 1957) deviate strongly however from the spin-only values 4/xB and
4.9/XB of the ferromagnetic and paramagnetic moment of the 3@ ions Cr z+ and
Mn 3+. Metallic conduction and a non-integer number of d electrons per metal
atom point to delocalized electrons in energy bands. In our opinion, an unam-
biguous interpretation of such properties has never been given (see, e.g., Bouwma
and Haas 1973).
A few metallic conducting chalcogenides exhibit magnetic moments that are in
reasonable agreement with the presence of transition metal ions having the
expected valency. This has been found clearly in Mel_~LaxMnO3 with Me = Ca, Sr
or Ba (Jonker and Van Santen 1950), and in CoS2 (Benoit 1955, Ohsawa et al.
1976). Such "ionic" moments are also found in the non-oxidic spinels CuCr2X4,
which makes an interpretation easier than in the case of the MnSb compounds.
The metallic conduction of these spinels is to be attributed to delocalized
electrons, their magnetic behaviour to electrons in localized 3d states of the Cr
ions (section 4.8).
We will now give a short review of the electronic structure of transition-metal
chalcogenides*. This structure is characterized by two broad bands (with a width of
5 to 10 eV) separated by an energy gap, and narrow d bands or ionic levels of
cations M n+ with a d m configuration. The broad bands arise from the strong
overlap of the occupied outermost s and p states of the anions (e.g. 2s and 2p of
02-) with the first empty s states of the cations (e.g. 4s of Mn 2+) which gives rise to
a "bonding" valence band and an "anti-bonding" conduction band. For a cation d
orbital the overlap with the anion orbitals is much smaller. This overlap tends to
break the highly correlated state of the d electrons in the cations Mn+(d m) and to
delocalize the d electrons in a narrow d band. This delocalization is counteracted
by the energy required for the excitation 2Mn+(dm)--->M~"+l)+(dm-l) + M("-l)+(dm+l).
Delocalization or localization occurs if the effect of the overlap or the excitation
energy dominates (Mott 1949). With an increasing ratio of the two factors, a sharp
transition from a localized to a delocalized state is expected (Mott transition).
The electrical and magnetic properties will be discussed with the help of
schemes (fig. 4) which represent the relative energies of electrons in the broad
valence and conduction bands and in d states, localized or otherwise. On the
right-hand side of such a scheme the one-electron energy versus the density of
states of the bands is sketched. The bands are occupied below and empty above a
Fermi level EF. On the left-hand side of each scheme we place the ionic levels of
cations M n+ with a d m electron configuration. Such an ionic level denotes the
relative energy (with respect to the broad bands) of the localized mth electron.

* See notes added in proof (a) on p. 737.


SULPHOSPINELS 617

E J
MIo-~
--E F

M~*__
~ density
of states
(3 b c

Mn*~ F Mn* F
d e f
Fig. 4. Schematic band structure of chalcogenide spinels. O n the right-hand side the schemes show the
conduction band and at a lower energy the valence band. T h e left-hand side shows the valency states
of a transition metal ion. UF denotes the Fermi energy.

Each level can accommodate one electron only. If occupied (situated below Ev) it
represents an ion with valency n, if empty (situated above UF) an ion with
valency (n + 1). For a given metal M, the M (n-l)÷ level lies above the M n÷ level.
The energy difference is the energy needed to excite an electron from Mn+(d m) to
M(n-1)+(dm+l)), i.e., the transition 2 M " + ~ Mtn+~)+ + M (n ~)+. E F coinciding with M "+
levels indicates a mixed valency state (a mixture of M "+ and M (n+l)+) that gives an
electrical conduction with a high concentration of charge carriers with a low
mobility (~<lcm2/Vs). Electrons or holes in a broad band give n- or p-type
metallic conduction, which decreases weakly with increasing temperature, and
they have a much higher mobility (>1 cm2/Vs).
In oxides the valence and conduction bands are separated by a wide gap in
which successive valency states are situated (fig. 4(a)). It has been proposed
(Albers and Haas 1964, Albers et al. 1965) that the properties of chalcogenides
and pnictides may be explained by cation levels that fall in one of the broad
bands. The various possibilities are drawn in figs. 4(b), (c) and (d), Which represent
semiconductors with conduction of electrons in a narrow band or holes in a broad
band (b), of electrons in a broad band or holes in a narrow band (c) and of
electrons or holes in a broad band (d). In figs. 4(e) and (f) there is metallic
conduction of either holes in the valence band (e) or electrons in the conduction
band (f).
A narrow d band is expected to consist of two separated or overlapping
branches arising from the cubic crystal field splitting (Goodenough 1969). In our
energy schemes narrow bands are drawn schematically without such details,
because experimental information is lacking.
618 R.P. VAN STAPELE

4. Sulphospinels containing copper

4.1. Valency of the copper ions

The Cu oxides CuCr204 and CuFe204 with spinel structure are semiconductors
and exhibit ferrimagnetic and paramagnetic properties that are consistent with the
valencies c,,
. . . .2+g-~3+g-~2-
2 ,-,4 and t~ u 2+12"
TM
, e 23 + g,-,4.
'~2- In contrast with this, all non-oxidic
spinels CuM2X4 show metallic conduction and essentially different magnetic
properties. For M = Cr they are ferromagnets with Curie temperatures above
room temperature, and the others ( M # Cr) show a temperature independent
paramagnetism. CuV2S4, CuRh2S4 and CuRh2Se4 become superconducting at low
temperature. A survey of the electrical and magnetic properties of the compounds
CuM2X4 is given in table 4. Detailed data will be given later, when each of the
compounds is discussed separately.
The striking difference in properties between the non-oxidic spinels CuM2X4
and the corresponding oxyspinels CuM204 points to an essential difference in
electronic structure. Two explanations have been proposed. The first is based on
the assumption that copper ions in the sulphospinels are monovalent, i.e., in the
3d 1° state, in contrast to their divalent state in the oxyspinels (Lotgering 1964a, b,
Lotgering and Van Stapele 1968a). The second is based on the assumption that
the copper ions are formally divalent, but that their 3d electrons are delocalized in
a band formed by the copper 3d states (Goodenough 1965, 1967 and 1969). The
crucial question in these matters is the valency of the tetrahedrally coordinated
copper ions in non-oxidic chalcogenides. The following arguments and experi-
mental results lead us to adopt the monovalent state:
(a) Divalent Cu ions are known to have a great instability in a sulphur lattice
(Akerstrom 1959), in agreement with the fact that no sulphides, selenides or
tellurides are known with tetrahedrally coordinated Cu ions, exhibiting the
properties of Cu e+, as CuCr204 does, for example.
(b) The C u - X distances in the copper sulpho-, seleno- and tellurospinels do not
differ significantly from the C u - X distances in other chalcogenides containing
monovalent Cu (Lotgering and Van Stapele 1968b) (see section 2 and Sleight
1967).
(c) A zero moment is observed at the A'sites in CuCr2Se4 and CuCr2We4 at
temperatures down to 4 K by means of neutron diffraction (Colominas 1967,
Robbins et al. 1967b) in contrast with 1/xB expected for a spin-ordered state at
T = 0 with Cu 2+ on the tetrahedral sites (see also section 4.8).
(d) The X-ray photo-electron spectrum of the copper 2p energy levels in
CuCr2Se4 closely resembles the spectra of CuCrSe2, CuCrS2 and CuA1S2, in which
compounds Cu is monovalent. All the spectra have the narrow peaks typical of
Cu + at the same energy, while satellites due to the simultaneous excitation of an
electron into empty states of the 3d shell are absent (Hollander et al. 1974).
(e) The chemical Shift of the K absorption edge of copper in CuCr2S4, Cufr2Se4
and CuCrzTe4 with respect to Cu metal is small and indicative of Cu + (Ballal and
Mande 1976) (table 5).
SULPHOSPINELS 619

...-t

O~

E
=
.=

II
,,,..¢
vh

+-

I "~

~x x Q
x , x ~ " i ,.- I I
',R. oo,~ "T'

b,-

~.~. , ~,, ~, ,~
d~

e~
3~
+ + ÷
.,,-

8 ii
~...'

E
x~ x×× ×~

d
620 R.P. VAN STAPELE

TABLE 5
Wavelengths of the K discontinuity of copper in various
compounds, according to Ballal and Mande (1976).

A (x.u.)
Compound Valency (-+0.09X.U.) AA

Cu metal 1377.67 -
CuO 2 1376.76 0.91
CuSO4.SH20 2 1376.42 1.25
CuC12.2H20 2 1376.44 1.23
CuCO3 2 1376.65 1.02

CuC1 1 1377.10 0.57


Cu20 1 1377.54 0.13
Cu2Se 1 1377.41 0.26

CuCr204 1376.80 0.87


CuCr2S4 1377.36 0.31
CuCr2Se4 1377.46 0.21
CuCraTe4 1377.56 0.11

Especially the last three experiments provide a direct establishment of the


monovalent state of copper in non-oxidic CuCrzX4. We adopt Cu ÷ in all sul-
phospinels, which is represented by a completely occupied Cu + level sufficiently
far below the top of the valence band. The experimental data do not give
information about the precise position of the Cu + state, so that this level will not
be drawn in the energy level schemes.

4.2. CuTi2S4

CuTi2S4 is a normal spinel (Hahn and Harder 1956) with lattice parameters as
given in table 1. Le Nagard et al. (1975) reported the existence of strongly
Cu-deficient CUl-xTi2S4 compositions with 0 ~< x ~<0.44. The cell edge varies linearly
with x, leading to a = 10.002-+ 0.003 A for x = 0. X-ray diffraction on a single
crystal with a = 9.985 A and a corresponding composition Cu0.92Ti284 gave u =
0.3805.
The compound has a nearly temperature-independent paramagnetic suscep-
tibility, decreasing slightly with temperature (fig. 5). Le Nagard et al. (1975)
measured on powders with a composition CuTi2S4 a room temperature suscep-
tibility of 5 x 10-4cm3/mol after correction for the diamagnetic susceptibility.
This agrees well with the value ( 4 . 5 + 0 . 1 ) x 10-4 cm3/mol reported by Lotgering
and Van Stapele (1968a).
The transport properties of powder samples have been measured by Bouchard
et al. (1965). The electrical conduction is metallic with a room temperature
resistivity of 4.8 x 10 -4 ~ c m (fig. 6). The Seebeck voltage is -11.8 ~xV/deg. These
results are in accordance with those of Le Nagard et al. (1975), who measured on
a single crystal a resistivity with a positive temperature coefficient and a value of
SULPHOSPINELS 621

"~mol l
10 xlO -~'

4 b

6;0 860 o'oo


T(K)
Fig. 5. Molar susceptibility of CuTizS4 corrected for the diamagnetic susceptibility. (a) According to Le
Nagard et al. (1975). (b) According to Lotgering and Van Stapele (1968a).

?(f?crn]~ L
7 - x lO-

1
100 200 3(?0 a T(K)
Fig. 6. The electrical resistivity of CuTi2S4and CuV2S4. (a) of a sintered powder of CuTi2S4, according to
Bouchard et al. (1965). (b) of a single crystal with composition Cu0.92Ti2S4,according to Le Nagard et al.
(1975). (c) of a sintered powder of CuV2S4, according to Bouchard et al. (1965).

4 . 2 x 10 4 ~ c m at r o o m temperature. T h e Hall constant was negative and in-


d e p e n d e n t of the t e m p e r a t u r e , with a value of 6.2 x 10 -4 cm3/C. Finally, V a n
M a a r e n et al. (1967) searched for superconductivity. D o w n to 0.05 K they did not
find a transition.
T h e negative sign of the Seebeck and Hall voltages is indicative of conduction of
electrons in a band. T h e extreme possibilities are a Ti 3÷ level well a b o v e the
b o t t o m of the b r o a d conduction b a n d (as in fig. 4(f)) or well below it (as in fig.
4(c)). In the second case the Fermi level will fall in a partly filled and b r o a d e n e d
Ti 3÷ band, since half of the Ti ions has d o n a t e d its d electron to the Cu ions,
c o r r e s p o n d i n g to the valency distribution Cu+Ti3+Ti4+S42-. It is h o w e v e r question-
able w h e t h e r the correlation b e t w e e n the remaining d electrons is strong e n o u g h
to m a k e this f o r m u l a realistic in the sense that Ti 3+ and Ti 4+ exist, exchanging
their valencies rapidly. G o o d e n o u g h (1969) argues that a 3d b a n d will be f o r m e d
that is filled with half an electron per Ti ion. In the first case, w h e r e half of the
Ti 3+ has already lost its 3d electron to the Cu ions, the o t h e r half of the Ti ions
622 R.P. VAN STAPELE

donates its 3d electron to the conduction band, which consequently contains one
electron per molecule. However, the temperature-independent susceptibility of
5 × 10 -4 cm3/mole is much too large to be attributed to the Pauli paramagnetism of
electrons in a broad band (Lotgering and Van Stapele 1968a). It can better be
reconciled with a narrow 3d band, as is also the case with the rather large
magnitude of the Seebeck coefficient. In this situation, where there is one type of
charge carrier, the number of charge carriers is 1.3 per molecule as calculated
from the value of the Hall constant. This agrees more or less with the expected
number of about one electron per molecule. We therefore draw the energy bands
as in fig. 7.

E~ n band

EF Ti3dband
~ valenceband
~g(E)
Fig. 7. Energy bands in CuTi2S4.

The magnetic moment x H induced by an applied magnetic field H in the 3d


band of the Ti ions with a Pauli susceptibility X gives a hyperfine field Hh~ = K H
on the nuclear spin of the Cu ions. The Knight shift K has been measured by
Locher (1968). It decreases slightly with increasing temperature and is propor-
tional to X (fig- 8), The ratio a = K/X = 5 can be compared with the corresponding
ratio between the hyperfine field of 55 kOe on the Cu nuclear spin originating in
the 6/~B of ordered C r s+ ions in semiconducting sulphospinels like CuuzInl/zCrzS4
(section 5.9). In such cases o~ = 1.63 (Locher and Van Stapele 1970). We can
conclude that the strength of the interaction between the 3d moment and the Cu
nuclear spin in CuTi2S4 is of the same order of magnitude as in semiconducting
spinels with well localized 3d electrons on the octahedral sites.

4.3. CuVzS4

CHV2S 4 is a normal spinel (Hahn et al. 1956) with lattice parameters as given in
table 1.
Little is known about the physical properties of the compound. This is due to
the difficulty of preparing sufficiently pure samples. As far as the susceptibility has
been measured, a temperature-independent susceptibility with values of 13.5 x
10-4cm3/mol (Blasse, private communication) or 9.1x 10 4cm3/mol (Lotgering,
private communication) has been found. At lower temperatures the susceptibility
behaves irregularly or shows signs of a weak permanent moment. The Knight shift
SULPHoSPINELS
623
0.5 CuV2S4
%
/

~o.3 0 CuTi 2 S~
._ Cu - metal
0.2

03
:3

I non- metallic compounds


$4
- 0 , ro

-0.2~
• CuRhzSe
" Cu Rh2 $44
-o.#_
0
100
200 3 0 - - ~ U '--d
T(K)
Fig, 8. The Cu Knight shift with respect to CulnSea, in the COmpoundsindicated (/,ocher 1968).
of the Cu nmr line (fig. 8) is positive and decreases with decreasing temperature
([,ocher 1968). The Cu nmr line suddenly broadens near 90 K and disappears at
lower temperatures. This remarkable phenomenon, which is possibly due to a
phase transformation, has not been investigated any further (Locher 1968). Both
the Knight shift and the SUsceptibility corrected for the diamagnetic SUsceptibility
of the sample Used in the nmr measurements (10.9 x I0-4) are about twice as large
as in CuTi2S4. This means that the interaction between the Cu nuclear spin and
the 3d mOmen!s are of COmparable strength in both COmpounds,
(+5~.2e~V~d~1~hYyversus temperature cu ,,,
1965). The ~'" ~,w me occurrenceof~,_,.a~e (fig: 6/c)) and Seeh,~..,- •
t. ~yp~ metallic Cond,.~,;-_ ,L--~'r" voltage
range 4.45 to material
3.95K exhibits superconductivity
measured on samples with transition temperatures
-~,.,,un (t~oncbardinetthe
al.
(Van Maaren et al. 1967). contaminated with 5 to 10% Cu3Vs4
The high Value of the Paul/ Susceptibility points to COnduction in a rather
narrow band. It is reasonable to assume that this is the t2g band of the V ions. This
is also in agreement with the Knight shift, which has a positive sign as in CuTi, S4.
However, this cannot be the Whole Story. The t2g band o
electrons per molecule. With Cu in the 3dI0 can contain twelve
statetheonly
molecule will occupy the band. One therefore expects three electrons
conduction to be thatper
of
electrons in a narrow band, with a negative and rather large Seebeck coefficient.
EXperimentally, the Seebeck voltage is positive and this means that the COnduc_
tion is partly due to holes in another band. Because the t2g band of the V ions is
expected to have a lower energy than that of the Ti ions, we assume that it
verlaps with the broad band structure of the valence band, as sketched in fig. 9.
624 R.P. VAN STAPELE

E,
conduction band

EF ~ n v 3d band
d
~"~/~ ~- g (E)
Fig. 9. Energy bands in CuV2S4.

This picture is essentially that given by Goodenough (1969). The simultaneous


occurrence of itinerant electrons and itinerant holes coincides with the occurrence
of superconductivity. The possibility of such a coincidence has been suggested in
the interpretation of the behaviour of the superconducting transition temperature
in the series CuRh>xSnxSe4 (Van Maaren and Harland 1969).
Recent results of measurements on powder samples and single crystals of
CuV2S4 confirm the earlier observations (Le Nagard et al. 1979). The electrical
resistivity, the magnetic susceptibility, and the nuclear magnetic resonance of the
nuclei 51V, 63Cu and 65Cu show a more or less sudden change below 100 K.
However, X-ray diffraction fails to show a structural phase transition. Down to o
4 K the c o m p o u n d r e m a i n s a cubic spinel with a cell edge that varies from 9.810 A
at 300 K to 9.782 A at 4.2 K*.

4.4. CuC02S4 and CuCoTiS4

CuCo2S4, which occurs in nature as the minerals carolyte and synchnodymite, is a


normal spinel (De Jong and Hoog 1928). The lattice parameters are given in table
1. The compound exhibits a p-type metallic conduction, as appears from the
resistivity versus temperature curve (fig. 10(a)) and the Seebeck voltage of
+12.71xV/deg (Bouchard et al. 1965). CuCo2S4 has a temperature-independent
paramagnetic susceptibility with a value of 3.7 to 3.8 × 10-4cm3/mO1 (Lotgering
and Van Stapele 1968a). The part of the specific heat that depends linearly on the
temperature is large (y = 24.4 mJmol-lK-2), which means a high density of states at
the Fermi energy (Van Maaren, private communication).
Replacement of Co by Ti removes the metallic conduction. This is shown by
the properties of CuCoTiS4 (Lotgering and Van Stapele 1968a, Lotgering 1968b).
The observed Seebeck voltage of +110 ixV/deg is much too high for metallic
conduction and proves that the compound is a semiconductor. The resistivity
versus temperature curve (fig. 10(b)) corresponds to a behaviour between that of a
metal and a semiconductor. This is found in several semiconducting sulphospinels
and has to be attributed to a small deviation from the stoichiometric composition,
which gives a small number of charge carriers with a high mobility. The nearly
temperature-independent magnetic susceptibility, corrected for the diamagnetic

* See notes added in proof (b) on p. 737.


SULPHOSPINELS 625

?(#cm) 9(9cm)

8x1()'
7 ! " T(K)
160 260 300
6
5
4
3
2
1
1(]0 '
200 '
3.00 ~ T(K)

Fig. 10. The electrical resistivity of: (a) CuCo2S4, according to Bouchard et al. (1965), (b) CuTiCoS4,
according to Lotgering (private communication).

susceptibility ( - 1 . 8 x 10-4cm3/mol), was found to be 2.8 to 2.9x 10-4cm3/mol.


The properties of CufoTiS4 can easily be understood on the basis of valencies,
represented by Cu+fo3+Ti4+S4,with the Co 3+ ions in the low spin t6g state, as in
oxyspinels. The empty 3d states of the Ti 4+ ions and the completely filled 3d states
of the Cu 1+ ions explain the semiconduction. The temperature-independent sus-
ceptibility arises from the Van Vleck susceptibility of the Co 3+ ions (Lotgering and
Van Stapele 1968a).
On the assumption of monovalent copper ions the p-type metallic conduction of
CuCo2S4 has been attributed to the conduction of holes in the valence band. This
hole conduction can be compensated by the substitution of Ti for Co, since the Ti
3d states have a higher energy (fig. 7), which explains the semiconductivity of
CuCoTiS4. If all the Co ions in CuCo2S4 were in the low spin t6g state, the Van
Vleck susceptibility would amount to some 5.7x !0-4cm3/mol, about the
measured value corrected for the diamagnetic susceptibility. This has led to the
conclusion that all Co ions are trivalent and in the low spin state, and that the
magnetic susceptibility of the conducting holes is small (Lotgering and Van
Stapele 1968a). However, the specific heat, mentioned above, indicates a high
density of states at the Fermi energy (10.3 states/eV molecule), which would mean
a Pauli susceptibility of 3.3 x 10 -4 cm3/mol. This is of the same order of mag-
nitude as the observed susceptibility. Such a poor agreement between magnetic
and specific heat data has also been noted in the compounds CuRh2S4 (section 4.5)
and CuRh2Se4 (section 4.6). In the last case the specific heat was found to be
strongly enhanced by the electron-phonon interaction.
An unexplained question is why CuCo2S4 behaves electrically and magnetically
as CuRh2S4 and CuRh2Se4 but does not become superconducting as the two Rh
compounds do.
Locher (1968) has measured the Cu and Co nmr in CuCo2S4. The Knight shift of
the Cu resonance is negative in this p-type conducting material, whereas it was
positive in CuTi2S4 and CuV2S4, in which compounds the magnetic susceptibility is
due to the Pauli paramagnetism of the 3d bands of Ti and V (fig. 8). The
quadrupole splitting in the 59Co nmr spectrum, deduced from the complicated nmr
626 R.P. VAN STAPELE

powder spectrum, and confirmed by zero field nuclear quadrupole resonance, is


much larger in CuC02S4 than in oxides such a s Z n C 0 2 0 4 and Co304 (Locher 1968).

4.5. CuRh2S4 and CuRh2-xCoxS4

The normal spinel CuRh2S4 (Blasse and Schipper 1964) exhibits essentially the
same electrical and magnetic properties as CuRh2Se4, which compound will be
discussed in the next section (4.6). We will therefore review experimental data
without much comment.
The resistivity at room temperature is 1.6x 10-3~cm (Lotgering and Van
Stapele 1968a). The Hall coefficient is negative (Van Maaren, private com-
munication). The compound becomes superconducting between 4 and 5 K. The
transition temperature and width depend on the method of preparation (Van
Maaren et al. 1967, Robbins et al. 1967a, Schaeffer and Van Maaren 1968, Dawes
and Grimes 1975, Shelton et al. 1976). The reason for this dependence, which has
been observed only in the case of CuRh2S4, is not known. The pressure depen-
dence of the transition temperature To has been studied on samples with To
between 3.8 and 4.8 K (Shelton et al. 1976). The pressure derivatives increase with
increasing To from 2.73 × 10-5 K/bar at To = 3.81 K to 4.95 x 10-s K/bar at To =
4.76 K. The increase of To with pressure is ascribed to the increase of the Debye
temperature, which primarily should determine To in compounds like
CuRh2S4. It is doubtful, however, whether this conclusion is correct. In the series
of mixed compounds CuRh2_xCoxS4, the lattice parameter varies linearly from
9.790 A at x = 0 to 9.498 A at x = 2 and the Debye temperature increases from
230 K at x = 0 to 310 K at x = 1 and 334 K at x = 2. However, To decreases fast
for small increasing x and is about 1 K for CuRhCoS4 (Van Maaren, private
communication). Above 50mK, CuCo2S4 has not been found superconducting
(Van Maaren et al. 1967).
At room temperature CuRh2S4 has a magnetic susceptibility of 1.6×
10 .4 cm3/mol (Blasse and Schipper 1964, Lotgering and Van Stapele 1968a). On
the basis of Cu + ions, Rh > ions in the low spin t 6 state and hole conduction in the
valence band the susceptibility can be explained as the sum of the diamagnetic
susceptibility (-2.1 x 10 .4 cm3/mol), the Van Vleck susceptibility of the Rh 3+ ions
(3 x 10-4 cm3/mol) and a Pauli susceptibility of 0.7 x 10 .4 cm3/mol (Lotgering and
Van Stapele 1968a). As in CuCo2S4 (section 4.4) and in CuRh2Se4 (section 4.6),
such a small Pauli susceptibility does not agree with the large linear temperature
coefficient of the specific heat (y = 30 mJmol-lK -2) (Schaetter and Van Maaren
1968), which corresponds to a Pauli susceptibility of 4 x 10 .4 cm3/mol.
The Knight shift of the Cu nmr lines measured on CuRh2S4 (Locher 1968) is
negative and has nearly the same value as in CuRh2Se4 (fig. 8).
A complete solubility has been observed in the series CuRh2(Sl_xSex)4 (Riedel
et al. 1976). The lattice constant varies linearly with x. The anion sublattice shows
a random distribution of chalcogen atoms with a mean value for the anion
parameter u of 0.381.
SULPHOSPINELS 627

4.6. CuRh2Se4 and CuRh2-xSnxSe4

CuRh2Se4 is a normal spinel with lattice parameters as given in table 1.


The c o m p o u n d has a p-type metallic conduction (Lotgering and Van Stapele
1968a) with at r o o m t e m p e r a t u r e a Seebeck coefficient of +7.3 p.V/deg and a
resistivity of 1.3 × 10 -3 ~ c m . The magnetic susceptibility depends weakly on the
temperature. At r o o m t e m p e r a t u r e the value is 1.1 × 10 -4 cm3/mol.
Based on the assumption of monovalent copper ions, the properties of
CuRh2Se4 have been attributed to Rh 3+ ions and hole conduction in the valence
band, as sketched in fig. 4e, with the Rh 3+ states below the Fermi level (Lotgering
and Van Stapele 1968a).
At low temperatures the c o m p o u n d becomes superconducting. Van Maaren et
al. (1967) found a transition at 3.47 K with a width of 0.04 K, which agrees with
the findings of Robbins et al. (1967a) who observed the transition at 3.46 K with a
width of 0.09 K, and those of Dawes and Grimes (1975), who obtained 3.50_+
0.05 K. Shelton et al. (1976) measured the influence of a hydrostatic pressure up to
22 kbar on To. On samples with Tc -- 3.49 K and 3.38 K the pressure derivative has
a value of 1.53 and 1.44 × 10 -5 K / b a r respectively. These m e a s u r e m e n t s were done
to confirm the suggestion that in spinel compounds the transition to superconduc-
tivity occurs at a t e m p e r a t u r e determined by the D e b y e temperature, the strength
of the electron-phonon interaction being roughly constant within the class of
superconductive spinel compounds. However, a study of the properties of the
system CuRh2_xSnxSe4 with 0~<x ~ 1 (Van Maaren and Harland 1969, Van
Maaren et al. 1970a, b) has shown that the occurrence of superconductivity in
CuRh2Se4 is determined critically by the details of the band structure. In this
investigation use has been made of the possibility to change the n u m b e r of charge
carriers drastically by substitution of Sn for Rh 3+. The Sn ions are expected to be
tetravalent and to decrease the n u m b e r of holes in the band, which is responsible
for the p-type metallic conduction. This turns out to be the case. T h e lattice
p a r a m e t e r changes linearly with the composition from 10.26 ,~ at x = 0 to 10.62
at x = 1. T h e conduction changes from metallic for 0 ~< x ~< 1 to semiconductive at
x = 1, at which composition the valencies are given by Cu+Rh3+Sn4+Se]-. The
Seebeck voltage is positive in the complete range o f compositions, whereas the
Hall coefficient RH is positive for 0.15 ~< x ~ 1 and negative for x ~< 0.15 (fig. 11).
For 0.5 < x ~< 1, R~I1 at r o o m t e m p e r a t u r e is about proportional to the n u m b e r
( 1 - x) of charge carriers per molecule C u +Rh2_xSn 3+ 4+x Se4 (fig. 11). For x < 0.5,
however, R n behaves anomalously. The value of R ~ 1 increases strongly with
decreasing x and R u changes sign at x -~ 0.15. The dependence of the transition to
superconductivity on the Sn concentration is strongly correlated to that of RH. To
decreases with x and no superconductivity above 0.05 K has been observed for
x > 0.5 (fig. 11). Van Maaren and Harland (1969) conclude from these m e a s u r e -
ments that conduction takes place in at least two bands, one of which shows
n-type behaviour for x < 0 . 5 . This can be reconciled with conduction i n the
valence band, which at the top is split into a single narrow upper branch and a
628 R.P. VAN STAPELE

T × 300 K
i
o 77 K
A 4.2K
D
u

O
E
I
J
I
4-1 lh- p-type
I
I
.1 .3 / .6 .8
X

,92

Fig. 11. Superconducting critical temperature To and reciprocal Hall constant R h 1 vs x of


CuRh2-xSnxSe4 according to Van Maaren and Harland (1969).

broader two-fold degenerate lower branch (Rehwald 1967). The additional


assumption which had to be made is that in the rigid band model the upper
branch will show p-type conduction for 0 ~<x < 1, whereas the lower branch
changes sign of conduction near x = 0.5. For x < 0.5 holes occur in the upper
branch and electrons in the lower branch. The simultaneous presence of holes and
electrons is apparently responsible for the appearance of superconductivity.
At this point it should be noted that the transport properties have also been
measured by Robbins et al. (1967b) and that they obtained a positive Hall
coefficient at room temperature. However, the sample used in the measurements
did not show superconductivity (Van Maaren and Harland 1969).
Schaeffer and Van Maaren (1968) have measured the heat capacity of CuRh2Se 4
and Van Maaren et al. (1970a, b) extended the measurements to the series
CuRh2-xSnxSe4 with 0 ~ x ~< 1. The Debye temperature, as calculated from the
lattice contribution, increases linearly with x from 200 K at x = 0 to 267 K at
x = 1. For x ) 0 . 6 the constant T in the linear term of the heat capacity is
proportional to ( 1 - x) 1/3, the electronic density of states expected for a parabolic
band with (1 - x) charge carriers per molecule. At smaller values of x, for which
compositions the compounds are superconductive, y vs ( 1 - x ) 1/3 is again linear,
but with a steeper slope. However, after correction of T for the enhancement of T
by the electron-phonon interaction, y vs ( 1 - x) ~/3 is linear in the full interval
0 <~ x ~< 1 (fig. 12). This is in accordance with the two-band model, in which it is
assumed that the density of states in the upper branch is much higher than in the
lower branch.
SULPHOSPINELS 629

I \,
[\
I

o ,,2,,~ .s,, 9 ,? ,9s ---~'~


1.0 .5 ~
Fig. 12. The constant y in the linear term of the heat capacity of CuRh2-xSnxSe4 vs x according to Van
Maaren et al. (1970a, b); (a) without and (b) with correction for electron-phonon interaction.

T h e d e n s i t y of states of CuRh2Se4 o b t a i n e d f r o m t h e c o r r e c t e d v a l u e of y
(12 m J / m o l K 2) is 5.1 s t a t e s / e V m o l e c u l e . T h i s w o u l d give a Pauli s u s c e p t i b i l i t y of
1.6 x 10 -4 cm3/mol.
T h e m a g n e t i c s u s c e p t i b i l i t y of t h e m i x e d c o m p o u n d s CuRh2_xSnxSe4 has b e e n
measured by Van Maaren and Harland (quoted by Van Stapele and Lotgering
1970) (fig. 13). T h e m a g n e t i c s u s c e p t i b i l i t y Xm is t h e s u m of t h e d i a m a g n e t i c 0(d),

I0£X~

T
Q •

i
0 0.5 ~ × 1.0
Fig. 13. Molar susceptibility at room temperature of CuRh2_xSnxSe4, according to Van Maaren,
quoted in Van Stapele and Lotgering (1970).
630 R.P. VAN STAPELE

the Van Vleck 0(~) and the Pauli susceptibility 0(p). Since Xd, which arises mainly
from the Se ions, is independent of x and X~, which is due to the Rh 3+ ions, is
linear in x, the curvature is caused by Xp. Taking Xd = - 2 . 5 x 10 -4 cm3/mole, X~v
can be estimated from Xm at x = 1 because Xp = 0 in the semiconductor CuRh-
SnSe4. This gives Xvv = 1.5 × 10 -4 cm3/mol Rh. From the susceptibility of CuRh2Se4
the Pauli susceptibility of this c o m p o u n d is then estimated to be 0.6×
10 _4 cm3/mol, which is smaller than, but of the same order of magnitude as, the
value estimated from the specific heat data.
The Knight shift of the Cu nmr lines in CuRh2Se4 (fig. 8) is negative and has
nearly the same value as in CuRhzS4 (Locher 1968).

4.7. CuCr2S4, CuCr2-xTixS4, CuCr2-xgnxS4 and CuCr2-xVxS4

The normal spinels CuCr2X4 and X = S, Se and Te, which were first prepared by
H a h n et al. (1956) have a metallic conduction and are ferromagnetic with a Curie
t e m p e r a t u r e above r o o m t e m p e r a t u r e (Lotgering 1964a). Because much of our
understanding of these interesting properties is based on the behaviour of mixed
compounds like CuCrz-xTixS4, we will also discuss the properties of these com-
pounds.
C o m p a r e d with CuCrzSe4 and CuCrzTe4, the study of CuCr2S4 has the disad-
vantage that the preparation of pure samples is m o r e difficult. Samples of CuCrzS4
always contain some antiferromagnetic CuCrS2 (Lotgering 1964b), which makes
magnetic data unreliable. Saturation magnetizations of 4.8 ~xB/molecule (Robbins
et al. 1970a) and 4.85/xB/molecule (Van der Steen, private communication) were
measured on the best samples, which also showed a lower Curie t e m p e r a t u r e than
the samples first measured (fig. 14 and table 6). A b o v e the Curie t e m p e r a t u r e the
magnetic susceptibility follows a Curie-Weiss law (fig. 15 and table 6).

TABLE 6
Magnetic data of the compounds CuCr2X4

0 Tc Saturationmagnetization
Compound Cm (K) (K) (/xB/molecule) Reference

CuCr2S4 2.40 390 420 4.6 Lotgering (1964b)


398 4.8 Robbins et al. (1970a)
390 4.85 Van der Steen
(private communication).
2.0 365 364 3.9 Kanomata et al. (1970)

CuCr2Se4 2.50 465 460 4.94 Lotgering (1964b)


433 4.77 Robbins et al. (1967b)
2.3 460 414 4.5 Kanomata et. al (1970)
2.55 441 430 5.07 Nakatani et al. (1977)

CuCr2Te4 2.90 400 365 4.93 Lotgering (1964b)


3.0 367 344 4.2 Kanomata et al. (1970)
SULPHOSPINELS 631

G(gauss cm3/g)

4.58.1J e
I t
80!
Cr? S~ -

6C - - 4.9gJJ B - - ~.
"~"'*'-*-""~'~"~._ ~ CuCr2Se~
4.93.1J B ~%'~'--~
. . . . . . .

gu Cr2 Tez , " ~

o I
0 100 200 300 400 500 °K
Fig. 14. The magnetization o- per gram as a function of temperature in H = 20 kOe of CuCr2X4
(X = S, Se and Te) (Lotgering 1964b).

T h e C u r i e t e m p e r a t u r e d e c r e a s e s with i n c r e a s i n g h y d r o s t a t i c p r e s s u r e .
K a n o m a t a et al. (1970) a n d K a m i g a k i et al. (1970) m e a s u r e d a r e l a t i v e p r e s s u r e
d e r i v a t i v e d In TJdp = - 3 . 1 x 10 -6 kg -1 c m 2.
T h e o c c u r r e n c e of p - t y p e m e t a l l i c c o n d u c t i o n a p p e a r s f r o m a low resistivity
(9 x 10 .4 ~ c m at r o o m t e m p e r a t u r e ) which i n c r e a s e s with i n c r e a s i n g t e m p e r a t u r e
a n d a p o s i t i v e S e e b e c k v o l t a g e of + 16.0 ixV/deg ( B o u c h a r d et al. 1965).

Cm--2.40.e = 39o K /' ,,/

j j " ~';//'-.LCuCr2 Se~


~~*~Cm=2.S0, e 465 K--
I ~ Cucrl2 Te~
c =z9o,e:,.oo K I

0 200 400 600 800 1000 K


Fig. 15. Reciprocal molar susceptibility of CuCr2X4 (X = S, Se and Te) (Lotgering 1964b).
632 R.P. VAN STAPELE

The hyperfine fields on the nuclear spins of the ~'6~Cu and the S3Cr nuclei have
been measured at various temperatures below the Curie t e m p e r a t u r e by Le Dang
Khoi (1968) (table 7). In an applied magnetic field the nmr frequency of the Cu
nuclear spins increases, while that of the Cr nuclear spins decreases. This means
that the hyperfine field on the Cu nuclear spins is positive (parallel to the applied
field) but that the hyperfine field on the Cr nuclear spins is negative. The
magnitude of the hyperfine fields measured by Berger et al. (1971) agrees well
with the results of Le Dang Khoi (table 7). The nmr spectrum measured at 4.2 K
by Kovtun et al. (1978) was m o r e complicated. The relaxation of the Cu nuclear
spins has been studied by Enokiya et al. (1977). In an applied magnetic field the
spin-lattice relaxation time T1 increases with the field strength up to 6 k O e and
remains constant in higher fields. The constant value, which can be considered to
be the relaxation time of nuclei in Weiss domains, was measured as a function of
temperature. The t e m p e r a t u r e dependence of T1 could be expressed as T ; 1=
a T + b T 35 with a = 0.38 and b = 4.8 × 10 -7. The first term, which dominates at low
temperature, is due to a process involving itinerant electrons, the second term has
been attributed to a three-magnon process.

TABLE 7
Hyperfine fields in CuCr2S4; (a)according to Le Dang Khoi (1968);
co/according to Berger et al. (1971)

300 K (a) 77 K (") 4 K co}

v MHz kG v MHz kG v MHz kG

63Cu 65.6 100.6 102.0


+58.1 +89.1 +90.3
65Cu 70.35 107.8 109.2
53Cr 27 -112 38.9 -162 39.8 -165

Recently, Ballal and Mande (1976) measured the chemical shift of the K
absorption edge of Cu in CuCr2S4. T h e shift with respect to the position of the
edge in Cu metal is small and typical of the 3d 1° electron configuration of the Cu +
ion (table 5).
The magnetic and electrical properties of CuCrzS4 are very similar to those of
CuCrzSe4, which will be discussed in the next section. For the m o m e n t we mention
that the valency distribution can be symbolized by Cu+Cr3+~Cr4-+sS4 with a small
n u m b e r of holes (~ per molecule) in the valence band and that the high Curie
t e m p e r a t u r e of CuCr2S4 is associated with the metallic conduction.
The correlation between the electrical and the magnetic behaviour is clearly
demonstrated in the series CuCr2_xTixS4 (Lotgering and Van Stapele 1968a). If we
adopt Ti4++ Cr 3+ to be stable with respect to Ti 3+ + Cr 4+ (which holds in oxides)
substitution of Ti for Cr compensates the p-type conduction. For x = 1 a semi-
conductor with valencies Cu+Cr3+Ti4+S24- is expected, which is experimentally
confirmed by a relatively high resistivity ( 0 . 6 ~ c m ) and Seebeck voltage
SULPHOSPINELS 633

( - 2 5 0 txV/deg), and a c h a n g e of sign of the latter quantity (fig. 16). T h e Curie


constant Cm = 1.84 is close to the spin-only value 1.88 of C r 3+. T h e Curie
t e m p e r a t u r e decreases almost linearly with x to a low value (0 = 4 K) at x = 1 (fig.
17). This decrease, which is m u c h stronger than the reduction expected for
dilution with n o n - m a g n e t i c Ti e+ ions, m e a n s that the strong ferromagnetic inter-
action is a b o u t p r o p o r t i o n a l to the n u m b e r of charge carriers giving metallic
conduction.
A further r e p l a c e m e n t of C r 3+ by Ti 3+ in Cu+Cr3+Ti4+S 2- formally gives a
mixture of Ti 3+ and Ti 4+, so that n-type metallic conduction occurs for 1 < x ~< 2
(fig. 16) as in CuTi2S4 (section 4.2). T h e c h a n g e of sign of the Seebeck voltage at
x = 1 (fig. 16) corresponds to the change f r o m p-type metallic conduction in the

Seebeck coeff (j.JVldeg)

100

-100
S
- 200

- 300
0
~X
Fig. 16. Seebeck coefficient of CuCr2-xTixS4, according to Lotgering and Van Stapele (1968a).

LO0

T30£
x-"

-~ 200
B Cu Cr2_xTix S L
¢-,

E 100
,e

0 0.5 1 1.5 2
~X
Fig. 17. Curie temperatures (0) and asymptotic Curie temperatures (x) of CuCr2 xTixS4(Lotgering and
Van Stapele 1968a) and CuCr2-xVxS4(Robbins et al. 1970a).
634 R.P. VAN STAPELE

valence band to n-type metallic conduction in a Ti 3+ band (fig. 7). With increasing
n-type conduction the asymptotic Curie t e m p e r a t u r e also increases (fig. 17), which
shows the ferromagnetic interaction between the Cr ions to be enhanced by the
conduction electrons. In the interval 1 ~< x ~< 2 the Curie constant remains close to
that of Cr 3+ (fig. 18), which confirms the assumption that Ti 4+ + C P + is stable with
respect to Ti 3+ + Cr 4+ or, in other words, that the C1e+ levels have a lower energy
than the Ti 3d band in fig. 7. The Curie constants of the compounds CuCr2_xTi~S4
with x < 1 gradually change from Ca = 1.88 at x = 1 to Ca = 2.40 for CuCr2S4 (fig.
18). The latter value is smaller than 2.88 expected for CuCr3+Cr4+S4 with spin-only
values of 1.88 and 1.0 for Cr 3+ and Cr 4+ respectively. This lack of agreement also
exists in the case of CuCr2Se4 (section 4.8).

Cm

~2.88
•-, .,.... Cq3. Cry) x

ol
o 1
~X
Fig. 18. Molar Curie constant of CuCr2 xTixS4 (0) (Lotgering and Van Stapele 1968a) compared with
the spin-only value of Cr3++ (1 - x) Cr 4+ for x < 1 and of (2- x) Cr3+ for x > 1.

The substitution of Sn for Cr in CuCr2S 4 has the same effect as the substitution
of Ti. The strongly ferromagnetic interactions between the Cr ions have disap-
peared in CuCrSnS4, which is paramagnetic down to 4 K with an asymptotic Curie
t e m p e r a t u r e of - 2 0 K (Sekizawa et al. 1973). The ion distribution has been
determined by Strick et al. (1968). Cu occupies the tetrahedral sites; the Cr and Sn
ions occupy the octahedral sites. As shown in fig. 19, the lattice p a r a m e t e r
increases and the Curie t e m p e r a t u r e decreases linearly with x in the series
CuCr2_xSnxS4 in the interval 0~<x <~ 1 (Sekizawa et al. 1973). The electrical
resistivity increases and the c o m p o u n d behaves like a semiconductor at x = 1. All
these properties are understandable with tetravalent Sn ions replacing the Cr ions,
as in CuCrz_xTixS4. Sekizawa et al. (1973) also measured the UgSn M6ssbauer
spectra. At a low Sn concentration, as in CuCrl.9Sn0.1S4, they measured a hyperfine
field of 580 k G at 80K. This agrees with the value of +530 k G measured by
Lyubutin and Dmitrieva (1975) in CuCr0.95Sn0.05S4. The positive sign means a
hyperfine field parallel to the Cr magnetization. This field is mainly due to the
supertransferred hyperfine interaction between the Sn nuclear spin and the Cr
spin density, which is partly transferred to the Sn 5s orbital. T h e hyperfine field
due to one 5s electron is estimated by extrapolation of the experimental data of
SULPHOSPINELS 635

- 10.2
@t
I"" Q(,~)
Tc(K
10.0 l
T 40( \
t1¢"
9.8

300 \
xTc
oN
20C
N
oNN
10C No
N
N\
0 I "~:CTN)
0 0.5

Fig. 19. Cell edge and ferromagnetic Curie temperature of CuCr2-xSnxS4 according to Sekizawa et al.
(1973).

Cd (Kelly and Sutherland 1956) and In (Campbell and Davis 1939) to be 16 MG.
The observed hyperfine field o n ll9Sn thus corresponds to a spin density of 3.3%
in the Sn 5s orbital.
An interesting series of compounds is CuCr2-xVxS4, which has been investigated
magnetically by Robbins et al. (1970a). The cell edge behaves anomalously; after
an initial rise for x ~<0.375 it decreases to the smaller cell edge of CuV2S4. This
and the initially slow decrease of the magnetic moment (see fig. 20) has been used

.IJf lJJB ) g.825


al,&)

9.820 T

o
9.815

9.810

9.805

9.800

0.15 110 ll5 2.0


~X
Fig. 20. Lattice parameter a and magnetic moment per formula unit at 1.5 K and 15 kOe of
CuCrz-xgxS4, according to Robbins et al. (1970a).
636 R.P. VAN STAPELE

by Robbins et al. (1970a) as arguments for a localized 3d 2 configuration of V 3+ for


vanadium concentrations smaller than 0.375. The magnetic moment of the V ~+
ions is assumed to couple ferromagnetically with the Cr magnetization. The
results of an nmr study of vanadium-substituted CuCr2S4, however, point in a
different direction. Berger et al. (1971) interpret the splitting of the Cu nmr lines,
observed in a sample with x = 0.1, and the shift of the broadening Cu lines at
higher vanadium concentrations, as being due to non-magnetic V ions. The
replacment of Cr 3+ ions by non-magnetic vanadium ions also splits the 53Cr nmr
line. The magnitude of the splitting corresponds to a supertransferred hyperfine
field of +87 k G exerted by the six Cr neighbours on a 53Cr nuclear spin. With a
total field of - 1 6 5 k G (table 7) the sum of the fields from the core polarization
and the band polarization is then estimated to be about - 2 5 2 kG. If the presence
of holes in the valence band does not influence the supertransferred hyperfine
interactions between the Cr ions, the field from the band polarization can be
estimated by comparison of CuCr2S4 with the semiconductor CdCr2S4. The
hyperfine field on the Cr nuclear spin in the last compound is - 1 9 0 k G (see
section 5.3), which gives +25 k G from the band. The field from the core polariza-
tion is then - 2 7 7 kG, an acceptable 8% below the value of 300 kG measured in Cr
ions in oxides (Geschwind 1967).
The rapid decrease of the Curie temperature of CuCr2-xVxS4 with the vanadium
concentration (Robbins et al. 1970a) parallels the behaviour of CuCr2_xTixS4 as
shown in fig. 17. This suggests that the substitution of vanadium causes a rapid
decrease of the number of holes in the band of Cufr2S4. Because T ~ 0 for
x ~ 0.5 this might mean that the vanadium ions start to donate two electrons to
the band, twice as many as the Ti ions in CuCr2_xTixS4. Such an occurrence of V 5+
(3d°) ions at low vanadium concentrations would also explain the non-magnetic
character of the vanadium ions observed by Berger et al. (1971) and the slow
decrease of the ferromagnetic moment, which is expected to decrease from 5/x~
per molecule to 4.5/zB in the hypothetical semiconductor C u ('~3+~tl'5+c~._.ll.5v0.4.5o However,
as Robbins et al. (1970a) argue, the V ions start to form a partly filled 3d band
from a vanadium concentration as low as 0.375. This changes the picture com-
pletely, giving rise to metallic conduction and ferromagnetism with a low Curie
temperature and a low ferromagnetic moment that is difficult to saturate.

4.8. CuCr2Se4, CuCr2-xRhxSe4 and CuCro.3Rhl.7_xSnxSe4

CuCr2Se4 is a normal spinel (Hahn et al. 1956) with lattice parameters as given in
table 1. The compound is, like CuCr2S4, a metallic ferromagnet (table 6) with
Tc = 460 K, /xf = 4.94/xB/molecule (fig. 14), 0 = 465 K and Cm = 2.50 (fig. 15)
(Lotgering 1964b). Since the material could be prepared in a pure state, the
saturation value is considered to be reliable. Tc = 433 K has been found from
measurements in a much lower field (50 Oe instead of 10 kOe), but on a sample
not completely pure as seen from /xf = 4.77/xB/molecule (Robbins et al. 1967b).
Measurements under hydrostatic pressure give d In Tc/dp = - 1 . 0 × 10 -6 kg icm2
(Kanomata et al. 1970, Kamigaki et al. 1970). The magnetic transition at Tc for
SULPHOSPINELS 637

increasing temperature at 50 Oe is accompanied by a large volume increase and a


sudden change in magnetization, which points to a first-order transition (Belov et
al, 1975). Magnetoscrystalline anisotropy constants Ka and /£2 of - 6 . 9 and
- 0 . 9 x 10s erg cm -3 have been measured on a CuCr2Se4 single crystal. The aniso-
tropy changed upon annealing in vacuum or in a Se atmosphere (Nakatani et al.
1977).
Metallic p-type conduction occurs down to 1 K. A resistivity of 2 x 10 -4 ~'~cm at
room temperature and a Hall constant of (1.0_+ 0.5)10 -l° Vcm-IG-1A -1 at 492 K
were measured on a single crystal (Lotgering 1964b). The value of (1.5 +
0.5)10 -1° Vcm-IG-1A -1 given in the paper is incorrect. This Hall constant, which
was measured in the paramagnetic state, gives 0.1 charge carrier per molecule.
For the Seebeck voltage the values +20.5 txV/deg (Robbins et al. 1967b) and
+40 p.V/deg (Lotgering and Van Stapele 1968a) are reported.
As we have discussed in section 4.7, a striking correlation exists between
metallic conductivity and strong ferromagnetism in the non-oxidic compounds
CuCr2X4. Such ferromagnetism was interpreted by Zener (1951) as being due to
an indirect coupling of localized incomplete d shells via the conducting electrons.
A positive or negative exchange coupling between the localized spins and the
spins of the conduction electrons results in a ferromagnetic coupling between the
localized spins and in a positive or negative partial spin polarization of the
conducting electrons. Goodenough (1965, 1967) assumed that in the compounds
CuCrzX4 (X = S, Se and Te) the copper ions are formally divalent and that the
conduction takes place in a narrow band of copper 3d states. The net mag-
netization of about 5 b~B/molecule was explained as the sum of 6 #B/molecule of
the ferromagnetically coupled Cr 3+ spins and -1/xB/molecule of a fully polarized
copper band. Lotgering and Van Stapele (1967, 1968a) assumed that the copper
ions are monovalent and attributed the p-type conduction to a small number of 6
holes/molecule in the valence band. These holes are brought on by the presence
of formally tetravalent Cr ions, i.e. holes in the Cr 3+ states, which were assumed
to be situated below the top of the valence band, as sketched in fig. 21(a) for
CuCr2Se4. The corresponding valence distribution in this compound can be
symbolized by Cu +{Crl+~Cra-a}{Se4
3+ 4+ 2 - aSea}
- with 6 ~ 0.1, as determined by the Hall
constant. In this model the net magnetization results from (5 + 6)/xB/molecule of the
Cr ions, opposed by a spin polarization of the valence band.
As discussed in section 4.1, measurements of the X-ray photoelectron spectrum
of C u f r 2 S e 4 (Hollander et al. 1974) and of the shift of the K absorption edge
(table 5, Ballal and Mande 1976) confirmed the monovalency of the copper ions.
Neutron diffraction experiments on C u f r 2 S e 4 (Robbins et al. 1967b, Colominas
1967) and on CuCr2Te4 (Colominas 1967) showed the absence of a magnetic
moment on the copper ions. The moment localized on the Cr ions was found to be
close to 3/xR. However, the values of the magnetic moments derived from neutron
diffraction experiments depend critically on the form factors used. In a recent
study of the diffraction of polarized neutrons by a C u f r 2 S e 4 single crystal,
Yamashita et al. (1979a) used measured form factors of the copper and the
chromium ions. These authors determined a moment of -0.07___ 0.02/xB on the
638 R.P. VAN STAPELE

EI
I

(1.8)CrL.÷T...~5: 0.1 hole/molecule


~~//~- EF

~ g ( E )
o) Cu Cr2SeA
A
E
[1_8)Cr~,~ / 6 = 0.1hole/molecule

v/////////////'A ~ g [E)
b) CuCrRhSe4
E
r ~ 6 = 0.7 holeI molecule
03 C r ~ ~ - E~ g(E)

C) Cu Cro3Rh17Sel,
Fig. 21. Energy bands and localized Cr states in CuCr2 xRhxSe4, according to Lotgering and Van
Stapele (1968a).

copper ions and of 2.64_+ 0.04/zB on the Cr ions. T h e difference c o m p a r e d with


the measured value of the saturation magnetization (5.01/zB/molecule) was attri-
buted to a uniform polarization of - 0 . 2 0 _ 0.11/zB/molecule. A p a r t from the small
negative m o m e n t on the copper ions, these results agree satisfactorily with the
model of Lotgering and Van Stapele with 6 = 0.2 holes/molecule in the valence
band.
The crucial point in the Lotgering and Van Stapele model is the tetravaient
state of Cr, which is known to exist in oxides (CrO2). Detailed experiments have
been done to verify its occurrence in selenides. Because Rh 3+ + Cr 4+ is expected to
be more stable than Rh 4+ + Cr 3+, a replacement of Cr > in CuCrzSe4 by Rh 3+ will
ultimately give a material that contains exclusively Cr 4+. This has been in-
vestigated in the system CuCrz_xRhxSe4 (Lotgering and Van Stapele 1968a)*. It
turned out that there is p-type metallic conduction throughout the series, the
Seebeck coefficient being small and positive for all values of x (fig. 22). The
asymptotic Curie t e m p e r a t u r e decreases gradually with increasing x (fig. 23),
while ferromagnetic ordering remains down at x = 1.7. As fig. 23 shows, the
magnetic m o m e n t in the ferromagnetic saturation is close to 2/xB per Cr ion for
x > 1, while the paramagnetic Curie constant per g r a m a t o m Cr is close to 1.0, the
spin-only value of Cr 4÷ (fig. 24). This confirms remarkably well the tetravalent
state of the Cr ions in CuCr2_xRhxSe4 with x > 1.
T h e behaviour of the system CuCr2_xRhxSe4 can qualitatively be understood on

* See notes added in proof (c) on p. 737.


SULPHOSPINELS 639

seebeckcoeff.(jJV/deg)

2C

10

I
0 1 2
~X
Fig. 22. Seebeck coefficient of CuCrz_xRhxSe4, according to Lotgering and Van Stapele (1968a).

e(K)
%\
T/*00 '~ N% -Mf(JJB}
300 xx~'~ l
20C ~\ 5

100 3
1
r
0 1 2

Fig. 23. Asymptotic Curie temperature ( 0 ) and ferromagnetic moment /x~ (O) of CuCr2_xRhxSe4
(Lotgering and Van Stapele 1968a, Van Stapele and Lotgering 1970). The solid lines represent/xf vs x
for Cr 4+ (a) and Cr 3+ (b).

Cm

I
1
~X
Fig. 24. Molar Curie constant of CuCr2_xRhxSe4 (Lotgering and Van Stapele 1968a, Van Stapele and
Lotgering 1970). The solid lines represent Cm vs x for Cr 3+ (a) and Cr 4+ (b).
640 R.P. V A N S T A P E L E

the basis of the energy level diagram for CuCr2Se4 (fig. 21(a)). Substitution of Rh
replaces Cr 3+ levels by deeper Rh 3+ levels. This lowers the Cr 3+ concentration,
leaving the number of holes in the valence band unchanged as long as the Fermi
energy falls in the Cr 3+ levels. Figure 21(b) illustrates the situation in CuCrRhSe4 with
6Cr 3+ ions, (1 - 6)Cr 4+ ions and ~ holes in the valence band. If the resonance width of
the Cr 3+ levels is sufficiently small, a further increase of the Rh concentration will
finally increase the number of holes in the valence band, leaving the remaining Cr
ions in the tetravalent state. This is sketched in fig. 21(c) for CuCr0.3Rhl.7Se4 with 0.3
Cr 4+ and 0.7 holes in the valence band.
The consequence is that if the Fermi level is shifted to higher energies at a
constant Cr concentration, the valency of the Cr ions should also change from
C r 4+ to Cr 3+, This has been accomplished in CuCr0.3Rhl.ySe4 by substituting
Sn 4+ for Rh 3+ (Van Stapele and Lotgering 1970). As we have discussed in
section 4.6, this substitution reduces the number of holes in the band of CuRhzSe4
to zero in the semiconductor CuRhSnSe4. In the series of compounds
CuRhl.7_xSnxCr0.3Se4 the properties change in the same way. As shown in fig.
25, the Seebeck coefficient is positive and increases strongly if x approaches 1,
indicating that CuRh0.7SnCr0.3Se4 is a semiconductor. The asymptotic Curie tem-
perature decreases strongly and, most important, the Curie constant per grama-
tom Cr changes gradually from the spin-only value 1.0 of Cr 4+ to the spin-only
value 1.87 of Cr 3+. The change of valency cannot be detected from the magnetic
moment in the ferromagnetic state, since for x >~0.3 the magnetization of
CuRhl.y_xSnxCr0.3Se4 is difficult to saturate.
It is difficult to account for the paramagnetic moment observed in CuCrzSe4 in
terms of paramagnetic moments of Cr 3+ and Cr 4+ (Lotgering and Van Stapele
1968a, Yamashita et al. 1979b). The observed value of the molar Curie constant
(Cm = 2.50) is lower than the spin-only values 2.97 and 3.06, which are expected

Cm/0.3 e (K) cdJaV,/deg) a(A)


A
t
2.0 Cr 3+ 60'- i6oo 10.8

40 10.6
1.5

o 20 IOL
o o
1.0~ n CFt~+
I 0 10.2
05 0 O5 0 05
--I.- X

Fig. 25. Curie constant per gram atom chromium Cm/0.3, paramagnetic Curie temperature 0, Seebeck
coefficient a and cell edge a of CuCr0.3Rhl.7 ,SnxSe4 as a function of the composition (Van Stapele
and Lotgering 1970).
SULPHOSPINELS 641

for (1.1 C r 3+ + 0.9 Cr 4+) and (1.2 Cr 3+ + 0.8 Cr 4+) respectively. This means that in the
paramagnetic state, too, the negative interaction between the Cr spins and the
spins of the conducting electrons cannot be left out of account.
Hyperfine fields in CuCrzSe4 have been measured on the nuclear spins of Cu, Cr
and Se ions. Locher (1967) found the Knight shift K of the Cu nmr to be
proportional to the magnetic susceptibility: K = c~X with c~ = 1.17x 103gcm -3.
Extrapolation to the magnetic moment in the ferromagnetic saturation gave a
hyperfine field of +68 kOe on the Cu nuclear spin. This agrees well with the
hyperfine field actually measured by Yokoyama et al. (1967a, b) and Locher
(1967) in the ferromagnetic state (table8).

TABLE 8
Hyperfinefieldsin CuCr2Se4;(a) accordingto Locher (1967);(b) according
to Yokoyama et al. (1967a, b).

77 K Extrapol. to 0 K 4.2 K Ref.

63'65Cu +70.4 kOe +72.1 kOe (a)


+70.6 kOe +72.6 kOe (b)
53Cr - 159 kOe - 161 kOe (b)
77Se 71.4 kOe 73 kOe (b)

The positive hyperfine field on the Cu nuclear spin is larger than the positive
field measured in semiconducting selenides: +72 kOe as compared to +32 kOe in
CUl/2Inl/zCr2Se4. This means that the negative polarization of the band in CuCr2Se4
gives a hyperfine field of the opposite sign on the Cu nuclear spin. This sign agrees
with the observations in the Pauli paramagnetic CuRhzS4 and CuRhzSe4, where a
negative Knight shift of the Cu nmr has been observed (fig. 8).
The hyperfine field on the 53Cr nuclear spin ( - 1 6 1 k O e of table 7) differs by
22 kOe from the - 183 kOe measured in the semiconducting CdCr2Se4 (see section
5.6). With Cr 3+ in both compounds this hyperfine field must be attributable to the
(negative) polarization of the polarized band. The hyperfine field due to the Cr
neighbours has been measured on the 119Sn nuclear spin in CuCrl.9Sn0.1Se4
(Lyubutin and Dmitrieva 1975) where it has the value of +490_+ 10 kOe. This
compares well with the values found in CuCr2S4 (section 4.7).

4.9. CuCr2Te4 and Cul+xCrzTe4

CuCr2Te4, which was first prepared by Hahn et al. (1956), resembles CuCr2S4 and
CuCrzSe4. It is a normal spinel (Colominas 1967), which can be prepared in a pure
state like CuCrzSe4. The lattice parameters are given in table 1.
CuCr2Te4 is a ferromagnet (Lotgering 1964b) with a Curie temperature Tc =
365 K and a saturation m o m e n t / ~ = 4.93/~B/molecule (fig. 14 and table 6). Above
the Curie temperature the paramagnetic susceptibility follows a Curie-Weiss law
with Cm = 2.90 and 0 = 400 K (fig. 15). The Curie temperature is not affected by
hydrostatic pressure (Kanomata et al. 1970, Kamigaki et al. 1970).
642 R.P. V A N S T A P E L E

The small shift of the K absorption edge of the Cu ions, as measured by Ballal
and Mande (1976), indicates these ions to be monovalent (table 5).
Hyperfine fields have been measured on the nuclear spins of 63Cu, 65Cu, 53Cr
and tZ~Te (see table 9). Locher (1967) measured the nmr of 63'65Cu in the
paramagnetic state and found the Knight shift K to be proportional to the gram
susceptibility gg: K = t~Xg with o~ = 0.80 x 103 gcm -3. Extrapolation to the ferro-
magnetic saturation moment gives a field of 32.5 kOe, in good agreement with the
values actually measured.

TABLE 9
Hyperfine fields in CuCr2Te4: (a) according to Yokoyama et al.
(1967a, b); (b) according to Berger et al. (1968b); (c) according to
Ullrich and Vincent (1967); (d) according to Frankel et al. (1968).

77 K Extrapol. to 0 K 1.4 K Ref.

63'65Cu +38.0 kOe +39.9 kOe (a)


39.9 kOe (b)
53Cr -145 kOe -151 kOe (a)
151.1 kOe (b)
lZSTe -181 kOe -187.5 kOe (b)
_+148 kOe (c)
_+160 kOe (d)

The Cu content can be strongly enlarged and single-phase CUl+xCr2Te4 with


0 ~< x ~< 1 has been obtained (Lotgering and Van der Steen 1971c). These materials
are metallic and ferromagnetic. Figure 26 gives a, Tc and/xf (4.2 K) as a function
of x. The crystallographic and magnetic properties can be attributed to a spinel
lattice with the excess Cu occupying a part of the tetrahedral sites normally not
occupied.
QtA) _ ,mc(k) ~f

11.30 " ' ~ ~ ~ ~ - u f

o - 400 4

11.2C
200 2

11.10 I 0 0
0 0.5 1.0
Y
Fig. 26. Cell edge a, ferromagnetic Curie temperature T~ and ferromagnetic moment /zf of
CUl+yCrzTe4, according to Lotgering and Van der Steen (1971c).
SULPHOSPINELS 643

4.10. CuCr2(X ~St)4 with X, X ' = S, Se and Te

O h b a y a s h i et al. (1968) o b s e r v e d c o m p l e t e solubility in the series CuCraS4_xSex.


Their data agree well with the results of later investigations. T h e cell edge varies
linearly with x, but the f e r r o m a g n e t i c Curie t e m p e r a t u r e does not, having a
m i n i m u m value at x ~ 1 (fig. 27). T h e magnetic m o m e n t also has an a n o m a l o u s l y
low value at x--~ 1 (fig. 28) (Belov et al. 1973, O b a y a s h i et al. 1968), the actual
magnetic b e h a v i o u r d e p e n d i n g to a great extent on the firing conditions (Obayashi

I oIAI
!
,0L

..~o 10.2
O ~o ~'
Sg ~
Tc(K) . ..,.~.-" 10.0

T LLO
"~

af'
9.8

400 Tc : i ""

360

320

I I I
1 2 3
~X
Fig. 27. Lattice parameter and ferromagnetic Curie temperature of CuCr2S4-xSex. (n) Data of
Ohbayashi et al. (1968); (×) Data of Riedel and Horvath (1973b); (O) Data of Belov et al. (1973).

~3
:a,
2

0 I
0 1 2 3
=X
Fig. 28. Ferromagnetic moment at 77 K and 12 kG of CufreS4-xSex, according to Below et al. (1973).
644 R.P. VAN STAPELE

et al. 1968). The anomalies in the X-ray intensities of CuCr2S3Se, observed by


Obayashi et al. (1968), were not seen by Riedel and Horvath (1973b), who
concluded that no deviations from a statistical distribution of the anions occur.
Data on the electrical behaviour have not been published, except by Belov et al.
(1973) who reported that the sample with x ~ 1 had a resistivity with a negative
temperature coefficient, while the samples with a different composition had a
resistivity with a positive temperature coefficient.
A satisfactory explanation of the properties of CuCr2S4_xSex has not been
given. However, the electronegativity of Se is smaller than that of the S ions. It is
possible that the Se ions, substituted for S in CuCr2S4, start to act as traps for the
holes in the valence band that are responsible for the metallic conduction and the
strong ferromagnetic interaction. It at least suggests that in the case of an
extremely strong binding of the holes to the Se ions, one can expect the
compound Cu+Cr3+S~-Se - to be a semiconductor with a much lower Curie
temperature than CuCr2S4 and CuCr2Se4.
A broad miscibility gap has been observed in the series CuCr2Se4_xTe~ (Riedel
and Horvath 1973b). Solid solutions are found only in the Te-rich samples with
x > 2.8, where the cell edge increases approximately linearly with x.

4.11. CuCr2X4-xYx w i t h X = S, S e or T e a n d Y = Cl, B r or I

The anions of sulphospinels can be replaced by halogen ions (Robbins et al. 1968,
Miyatani et al. 1968). Starting from the electronic structure of CuCr2X4 discussed
in section 4.8, one expects that the electrons produced by a replacement of X 2- by
Y- occupy the holes in the valence band and the C r 3+ levels (fig. 21(a)). A
semiconductor Cu+Cr~+X3Z-Y- is then expected to be the end of the series of
mixed crystals. The Curie temperature of CuCr2X4-xYx is expected to decrease
with increasing x, because the strength of the ferromagnetic interaction via the
conducting electrons will decrease with the decreasing number of holes in the
valence band. Although this behaviour has actually been observed, complications
have arisen. A difficulty encountered in the study of these materials is the
preparation of chemically pure and stable samples. The following experimental
results have been reported.
The series CuCr2S¢_xClx shows an increase of the cell edge and the saturation
magnetization with x, and a decrease of the ferromagnetic Curie temperature
(Sleight and Jarrett 1968). A sample with x ~ 1 and T c ~ 2 1 0 K exhibits an
electrical behaviour typical of an impure semiconductor.
Miyatani e t al. (1971a) prepared CuCr2Se4-~Clx with 0~<x ~<0.8. The lattice
constant increases slowly with x. The Curie temperature decreases with x, but the
magnetic moment remains constant at 5/xB/molecule. A sample of CuCr2Se3C1
was found to be ferromagnetic at room temperature but the material was found to
be very hygroscopic, which prevented further measurements (Robbins et al. 1968).
Recently, the magnetic behaviour of the series CuCr2Se4_xClx with x < 0.6 has
been reinvestigated by Yamashita et al. (1979b). Their data confirm the decrease
of the Curie temperature. However, the saturation magnetization was found to
SULPHOSPINELS 645

increase with x, which does not agree with the observations of Miyatani et al.
(1971a).
The series CuCr2Te4_xIx behaves anomalously. The cell edge passes through a
maximum as a function of x and the ferromagnetic Curie temperature varies
irregularly (Robbins et al. 1968). A M6ssbauer spectrum of i29I in a sample of
nominal composition CuCr2Te3I has been reported by Granot (1973).
The series CuCr2Se4_xBrx (Robbins et al. 1968) shows the best reproducible
properties and is therefore the most extensively investigated system among the
halogenide-substituted sulphospinels. The saturation magnetization /xf, the fer-
romagnetic Curie temperature Tc and the cell edge a are given as a function of x
in figs. 29, 30 and 31. The differences in the results of measurements of /xf are
considerable, which have to be attributed to impurities. From neutron diffraction
experiments, a magnetic moment of 3+0.18/XB was derived for Cr ions in
CuCr2Se3Br (White and Robbins 1968).
In early papers on CuCr2Se4_xBrx the transition from the metallic to the
semiconducting state is reported to take place at x ~ 0.5 in polycrystalline samples
(Robbins et al. 1968) or at x ~ 0.8 in single crystals (Sleight and Jarrett 1968).
More recent work on polycrystalline samples and single crystals indicates that the
transition takes place at x = 0.98 and is accompanied by a sharp decrease of Tc
(fig. 30) (Miyatani et al. 1971a). This illustrates again the close connection
between the strong ferromagnetism and the metallic conduction. In a recent
analysis of the asymptotic Curie temperature and the temperature dependence of
the magnetization, Yamashita et al. (1979b) conclude that the first-neighbour
C r - C r exchange interaction is only weakly influenced by the metallic conduction,
whereas the second-neighbour interaction changes from strongly positive in
CuCr2Se4 to weakly negative in CuCr2Se3Br.
At the outset it was found that substitutions with x > 1 are not possible.
However, more recent work has shown that materials with x > 1 can still be
obtained. A single crystal with a = 10.444 A, corresponding to x = 1.06 (fig. 31),
was found to be a semiconductor with Tc = 110 K and a band gap of 0.9 eV, as
deduced from optical properties (Lee et al. 1973). A series of polycrystalline
samples and single crystals from x = 0 to x - 2 could be prepared (Pinko et al.
1974). Figures 30 and 31 give a and Tc. In this series the largest a (10.444 A) and

61- ~'1
jdf / +
(.uB)~ ///+'1
b I

0 02 0.Z+ 0.6 08 1
~X
Fig. 29. ]Ferromagnetic moment of CuCr2Se4 ~Br~; (a) in 10 kOe at 4 K, Robbins et al. (1968); (b) for
0.1 ~< x ~< 0.5, /xs = (5.0 + x) ~B, Sleight and Jarrett (1968); (c) Miyatani et al. (1971a).
646 R.P. VAN STAPELE

i o

&O0

Tc(K) ~.+

O ~+~"
3O0
\
x~
I
I
I
4-

200

100 x
, , , , , , , , , x
0 02_ 0.4 0.6 08. 10 1.2 1.4 16 1.8 2.0
=X
Fig. 30. Ferromagnetic Curie temperature of CuCr2Se4-xBrx (symbols as in fig. 31).

10.46

x
10.44 x

10.42 [] o
x
x x
x ,,F x
"E 10.40 t-I + +e
G~
4-
10 [3

10.38 +
O
£.3

1036t uu~
10.34]_~ u , , , , , , ,

o o12 d4'o:6'o18' ~ 1'.2 114 1'.6 118 2


P-X
Fig. 31. Cell edge of CuCr2Se4_xBrx according to: (0) Robbins et al. (1968); (C)) White and Robbins
(1968); ([~) Sleight and Jarrett (1968); (+) Miyatani et al. (1971a) and (x) Pink et al. (1974).
SULPHOSPINELS 647

lowest Tc (84 K) of the halogen-substituted sulphospinels have been measured on


a chemically analyzed sample with the composition Cu0.99CrzSe2.mBr2.02. Deviations
from the stoichiometry according to CuyCr2Se4_zBr, with 0.8 ~< y ~< 1.2, 0 ~< x ~<2
and 0 ~ ( z - x ) ~ 0 . 2 have been found upon variation of the preparation con-
ditions. Not all spinels are stable in air. Those with y < 1 are, but the spinels with
y > 1 are not. The unstable spinels could be stabilized by means of a treatment
with NH4OH, which mainly extracts the excess Cu. The Cu content of stable
spinels depends on the Br content. It is always less than 1, with a minimum of
y = 0.78 at x ~ 1 (fig. 32). The origin of this behaviour is not known. Since
electridal and magnetic measurements have not been carried out, it is not possible to
discuss the valencies in these complicated materials.
Magnetostriction (Unger et al. 1974) and hysteresis loops (Unger 1975) have
been measured on some CuCr2Se4_xBr~ single crystals.

0 1 ~×
Fig. 32. Composition of stable compounds CuyCr2Se4-zBrx,according to Pink et al. (1974).

5. Ferromagnetic and antiferromagnetic semiconductors

5.1. General aspects

In this section we will treat the normal spinels ACr2X4 (X = S or Se) with
diamagnetic Zn 2+, Cd 2+ or Hg 2+ on A sites and magnetic Cr 3+ on B sites. The
existence of this kind of compounds was discovered by Hahn (1951). The spinels
CdCrzS4 and CdCrzSe4 exhibit the rare combination of quite strong ferro-
magnetism and semiconductivity (Baltzer et al. 1965, Menuyk et al. 1966). For this
reason they have been investigated more than any other of the sulphospinels. In
particular the influence of the magnetic ordering on the electrical and optical
properties is the subject of many papers. Before treating the compounds
separately we will discuss some general aspects.
The first spinels of this type to be investigated were ZnCrzS4 and ZnCrzSe4.
They are antiferromagnets and are anomalous in that they have a positive
asymptotic Curie temperature 0. The corresponding oxyspinels ZnCr204 or
648 R.P. VAN STAPELE

MgCrzO4 exhibit a normal antiferromagnetism and have a strongly negative 0.


The anomalous behaviour of the non-oxidic compounds has been attributed to the
combination of a ferromagnetic nearest-neighbour interaction between the Cr 3+
ions and antiferromagnetic superexchange CrXXCr interactions at larger dis-
tances (Lotgering 1964b). The difference in sign of 0 for X = O and X = S or Se
has been attributed (Lotgering 1964b) to a superposition of a ferromagnetic 90 °
Cr3+X2-Cr3+ superexchange interaction via the X 2- ion (Kanamori 1959) and an
antiferromagnetic exchange interaction caused by the direct overlap of the d
orbitals of the two Cr 3+ ions. The possibility of the latter interaction mechanism
was anticipated (Kanamori 1959, Goodenough 1960, Wollan 1960) before a
clear-cut example of it was known. In the oxyspinels the CrZ+-Cr3+ distance is
small (2.94A in ZnCr204) so that a strong d - d overlap gives a dominating
negative interaction. As a consequence of the much larger distance in the
sulphospinels (3.53 A for X = S and 3.71 A for X = Se) the negative interaction is
weakened so that the positive superexchange dominates. This is demonstrated in a
plot of 0 versus a (fig. 33). The increase of 0 with increasing lattice parameter or
Cr-Cr distance fits with the behaviour of a much wider class of Cr compounds
with a dominating 90 ° C r - C r superexchange interaction (Rtidorf and Stegemann
1943, Bongers 1957, Lotgering 1964b, Baltzer et al. 1966, Menyuk et al. 1966,
Bongers and V a n Meurs 1967, Motida and Miyahara 1970).
The strength of the various terms that contribute to the 90 ° Cr3+-Cr3+ inter-
action has been determined in a detailed analysis of the optical spectrum of Cr 3+
pairs in Cr-doped ZnGa204 spinel (Van G o r k o m et al. 1973). The interaction
- J S a " Su between the spins $ of two nearest-neighbour Cr ions a and b is the
result of nine interactions -J~jsi • sj between the spins si of the three electrons in

M:Cd,Hg
200
100

-- 0
,/ ,t'×--s
o
*--0~ x~ / 10 of-A) 11
100

200

300 /x=0
Z,00
Fig. 33. Asymptotic Curie temperature versus lattice parameter of the compounds MCr2X4.
SULPHOSPINELS 649

the d states day, d~z and dax of ion a and the spins sj of the three electrons in the d
states dby, dbyz and d~x of ion b. Due to the symmetry of a pair of nearest-
neighbour Cr ions there are four independent interactions among the nine £j: J~,
the direct interaction between spins of electrons in the dxy orbits; J,~, the
superexchange interaction between electrons in d states overlapping with the
same ligand p state; J', the superexchange interaction between electrons in d
states overlapping with two orthogonal p states of the same anion and Jc, the
interaction between electrons in d states overlapping with p states on different
anions (fig. 34). The direct interaction was indeed found to be by far the strongest
(Jd = -561 cm-1), while J= = - 1 0 5 c m -1, J'c = +117 cm -1 and Jc = +39 cm -1.
The existence of antiferromagnetic interactions MXXM via two anions X
appears from the antiferromagnetic ordering in many layer structures, in which
layers of magnetic ions M are separated by double anion layers. The sulphospinels
under considerations are ferromagnetic or antiferromagnetic and the kind of
ordering is determined by the relative strength of the positive MXM and negative
M X X M interactions. The occurrence of both ferromagnets and antiferromagnets
shows that the positive and negative interactions in sulphospinels are in equili-
brium. From this it can be concluded that the M X X M interactions are about ten
times weaker than the MXM interactions (Lotgering 1964b).
It is difficult to derive the strength of the exchange interactions from the
observed magnetic properties. The attempts that have been made, all except one,
have been based on simplifying assumptions with regard to the more distant
interactions. Lotgering (1965) analyzed the data of ZnCr2Se4, calculating the

dby ~ ~ ,-y

° ~ ~~'-y dyz P~ dbz


l o b
Jo:dxy-dxy Jc dyz -Pz ond dyz-Pz
o b

- c / I d;x
x L.-"I J.~ : dzx-Pz- y~ J~ d~×-p=± Px-dby
Fig. 34. (a) The negative interactions in the 90 ° Cr3+-Cr3+ superexchange. (b) The positive interactions
in the 90 ° Cr3+-Cr3+ superexchange.
650 R.P. VAN STAPELE

interactions J1, J2, J3 and J4 from the asymptotic Curie temperature, the pitch of
the observed helix a n d the magnetic susceptibility on the assumption that J3 = J4.
(We use the notation of table 10 and fig. 35.) Plumier (1966a, b) used the same
data to calculate J1, J2 and J3. Baltzer et al. (1966) used the interactions J1, J2, J4
and J5 and neglected J3 in a high temperature expansion of the magnetic
susceptibility to calculate the Curie temperature. Assuming all the more distant
interactions to be equally strong (e.g. K =-/2 = J4 = Js) they calculated J1 and K
from the Curie temperature and the asymptotic Curie temperature for the
ferromagnets CdCr2S4, CdCr2Se4 and HgCr2Se4 and the metamagnet HgCr2S4. A
similar approximation has been made in a multi-sublattice molecular field analysis
of these compounds (Holland and Brown 1972). However, the impressive analysis
given by Dwight and Menyuk (1967) showed that the magnetic properties are very
sensitive to the strength of J2 . . . . ./6 and that the validity of results obtained using
simplifying assumptions is doubtful. They analyzed the stability of the various
classical spin ground states, minimizing the Heisenberg exchange energy, and
found that in the specific case of ZnCr2Se4 a spiral ground state with the observed
properties can exist if the values of the interactions J~ . . . J6 fall inside a limited

TABLE 10
Notation of pairs of octahedral ions in the spinel lattice (fig. 35) and of
exchange constants.

Notation According to

BoB1 BoB2 BoB3 BoBa BoB5 BoB6 Baltzeret al. (1966)


Wo W1 W2 W3 W4 - Lotgering (1965)
J WJ UJ U'J. VJ U2J Dwightand
Menyuk (1967)
J1 J2 J3 J4 J5 J6 this work

133

, , A

/ I

© e.
[30 136
Fig. 35. Octahedral sites and anions in the spinel lattice.
SULPHOSPINELS 651

region. In a similar analysis the stability at T = 0 of various possible spin


configurations has been systematically investigated by Akino and Motizuki (1971)
with the restriction that J3 = J4 and J6 = 0.
Applying the Goodenough-Kanamori rules to the more distant exchange
interactions Dwight and Menyuk (1968) expected a negative sign for J3, J4 and J6,
a positive sign for Js, whereas the sign of J2 remained uncertain. However, a direct
determination of the strength of the more distant exchange interactions between
Cr 3+ ions in Cr-doped ZnGa204 by means of electron spin resonance revealed all
interactions to be negative. The constants J~ of the exchange interactions -J~so" si
were found to be: J 2 = - 0 . 9 4 c m -1, J 3 = - l . 2 2 c m -1, Y4=-0.80cm -I, Js =
-0.45 cm -1 and J6 = -0.55 cm -1 (Henning 1980).
The compounds under discussion are semiconductors, the behaviour of which is
usually not intrinsic but determined by small deviations of the stoichiometry or
the presence of impurities in small concentrations. The most striking property is a
strong influence of the magnetic ordering on the resistivity and magnetoresistance
in certain n-type compounds, e.g., in suitably doped CdCr2S4 and CdCr2Se4.
Measured as a function of temperature these quantities show a maximum close to
the ferromagnetic Curie temperature. These phenomena are due to the inter-
action between the spins of the charge carriers and the spins of the Cr 3+ ions. A
generally accepted mechanism has not been found, and which of the proposed
models applies depends on the type of conduction and the strength of the
interaction between the spin of the Cr 3+ ions and that of the charge carriers. If the
conduction is in a broad band and if the interaction between the spin of the
charge carriers and the Cr 3+ spins is relatively weak, so that the influence of the
charge carriers on the magnetic behaviour of the Cr 3+ spin system can be
neglected, the theory of Haas (1968) applies. It describes the dependence of the
number of charge carriers on the temperature and the strength of an applied
magnetic field as it is due to the splitting of the conduction band and donor or
acceptor levels in the exchange field from the Cr magnetization (Bongers et al.
1969). The scattering of the charge carriers at the spin disorder in the Cr spin
system gives the carriers a mobility that can also be strongly dependent on the
temperature and the magnetic field strength (Haas 1968, Patil and Krishnamurthy
1978, Aers et al. 1975). Both effects can give rise to a peak in the resistivity at the
Curie temperature and to a negative magnetoresistance with a maximum at the
Curie temperature. The critical behaviour of the resistivity has been studied
theoretically by Alexander et al. (1976) and Balberg and Helman (1978).
If the interaction between the spin of the charge carriers in a broad band and
the Cr spins is sufficiently strong, the magnetic state of the Cr spin system changes
and a magnetic polaron can exist. In this state the charge carrier is surrounded by
a cloud of magnetic polarization of the Cr spins. Yanase et al. (1970) and Yanase
(1972) have discussed the conditions for the stability of the magnetic polaron
state. They find the magnetic polaron to be stable in a region around the Curie
temperature, where it has a much smaller mobility than the original charge carrier
would have had. Qualitatively, this can explain the observed anomalies in the
electrical resistivity. The entropy of the polarized Cr spins gives rise to an
652 R.P. V A N S T A P E L E

anomalous thermoelectric power and, as Yanase (1971) has pointed out, this can
explain the observation by Amith and Gunsalus (1969) that a maximum in the
resistivity of Cdl-xInxCr2Se4 coincides with a minimum in the thermoelectric
power.
In the case of an impurity conduction influenced by the magnetic ordering,
Yanase and Kasuya (1968a, b) and Kasuya and Yanase (1968) proposed the model
of the magnetic impurity state. In this model the spins of the Cr ions, which are
neighbours of an occupied impurity state, are polarized by the exchange inter-
action between the spin of the impurity electron and the Cr spins. This stabilizes
the occupied impurity state and the stabilization energy will increase towards the
Curie temperature, where the Cr spin system is highly susceptible. This leads to
an activation energy that is maximum at the Curie temperature. Application of a
magnetic field will decrease the activation energy. The model of the magnetic
impurity state explains in this way a maximum in the resistivity and a maximum
negative magnetoresistance at the Curie temperature.

T A B L E 11
Magnetic data of the semiconductors MCr2X4: the asymptotic Curie tem-
perature 0, the ferromagnetic Curie temperature Tc or the antiferromagnetic
N6el temperature TN, the molar Curie constant Cm and the ferromagnetic
moment /zr. References: (1) Menyuk et al. (1966), (2) Lotgering (1956), (3)
Plumier et al. (1975), (4) Lotgering (1964b), (5) Von Neida and Shick (1969), (6)
Baltzer et al. (1966), (7) Srivastava (1969), (8) Hastings and Corliss (1968), (9) Le
Craw et al. (1967), (10) Eastman and Shafer (1967).

0 Tc TN m
Compound (K) (K) (K) Cm (/xB/molecule) Ref.

ZnCr2S4 18 _+8 3.34 (2)


18 (1)
CdCr2S4 152 84.5 3.70 5.15/./ (6)
156 86 3.8 5.55 (,/ (1)
6.02 co) (5)
84 (7)
HgCr2S4 142 36.0 3.62 5.35 (a~ (6)
- 6 0 (c~ (8)
36.1 (7)
ZnCr2Se4 115 -20 3.54 (4)
20 (1)
105 (dl (3)
CdCr2Se4 204 129.5 3.82 5.62 (a) (6)
210 130 3.66 5.6 (,/ (1)
172 140 4.48 5.98 _+0.4 ~e~ (9)
127.7 (7)
129_+2 5.94 -+ 0.04 (10)
HgCr2Se4 200 106 3.79 5.64 (a) (6)
105.5 (7)
(a) At 4.2 K in 10 kOe. co)At 1.5 K. (c)A spiral spin configuration in zero field.
(d) A spiral spin configuration at T < 21 K, antiferromagnetic microdomains at
T > 21 K. (e)At 1.5 K in 15 kOe.
SULPHOSPINELS 653

Effects of the magnetic ordering have also been seen in optical properties. The
generally observed p h e n o m e n o n is an anomalous shift of the optical absorption
edge (Busch et al. 1966, Harbeke and Pinch 1966) to a higher energy, as in
CdCrzS4, or to a lower energy, as in CdCrzSe4. The interpretation of these
observations depends heavily on the nature of the absorption at the edge. If a
broad conduction band plays a role in the transition, a red shift of the absorption
edge can be explained by the spin splitting of the band due to the exchange
interaction between the Cr spins and the spin of the excited electron. Rys et al.
(1967), Haas (1968) and Kambara and Tanabe (1970) have discussed this
mechanism in detail. The critical behaviour of the band gap in this case has been
treated by Helman et al. (1975) and Alexander et al. (1976). Helman et al. (1975)
included the influence of an applied magnetic field. Callen (1968) has suggested
another mechanism in which the shift of the absorption edge arises from a change
of the band energy due to exchange striction, which distorts the lattice. The shift
of the edge can have either sign, but it is estimated to be too small in all cases of
interest, as we will discuss later on. White (1969) has pointed out that the
absorption edge due to an indirect transition between the valence band and the
conduction band can show a shift to the blue below the Curie temperature.
Nagaev (1977) argued that a blue shift can arise from interband s-d exchange.
Since the magnetic, the electrical and the optical properties and their mutual
relation depend strongly on the detailed composition of the compounds, we will
review in the following each of the compounds separately. The magnetic data are
summarized in table 11".

5.2. ZnCr2S4
ZnCr2S4 is a normal (Hahn 1951) spinel (Natta and Passerini 1931) with lattice
parameters, as given in table 1.
In the paramagnetic region Lotgering (1956) observed above 100 K a Curie-
Weiss behaviour with 0 = 18 _+8 K and a molar Curie constant Cm = 3 . 3 4 - 0.06.
According to Menyuk et al. (1966) the Cr spins order antiferromagnetically below
TN = 18K. A b o v e TN Stickler and Zeiger (1968) observed a paramagnetic
resonance with g ~ 2 and an antiferromagnetic resonance at lower magnetic fields
below Ty. The zero-field antiferromagnetic resonance frequency follows an S = 3
Brillouin function fairly well. The antiferromagnetic spin structure is not known.
A flat spiral, as observed in ZnCrzSe4, does not seem to fit the neutron diffraction
spectrum of ZnCr2S4 (Stickler and Zeiger 1968).
Bouchard et al. (1965) found the compound to be a semiconductor with p
(300 K) = 5 x 10 l° f~cm and an activation energy q = 0.59_+ 0.03 eV at higher tem-
peratures. The samples measured by Albers et al. (1965) had a much smaller
resistivity and a small positive Seebeck coefficient. The cold-pressed samples
studied by Lutz and Grendel (1965) had a p-type conduction with acceptor states
0.5 to 0.8 eV above the valence band. Doping with W for Cr or with In for Zn can
compensate the p-type conduction. Above 200°C the activation energy in undoped
samples is 1.5 eV, which is considered to be the band gap energy.
* S e e n o t e s a d d e d in p r o o f (d) o n p. 737.
654 R.P. V A N S T A P E L E

5.3. CdCr2S4

CdCr2S4 is a normal (quoted by Hahn 1951) spinel (Passerini and Baccaredda


1931). Crystallographic data are given in table 1. As shown in fig. 36, the lattice
constant a decreases in a normal way down to 100 K, but increases slightly below
that temperature (Martin et al. 1969, Bindloss 1971 and G6bel 1976). More
striking is the anomalous increase of the width of the X-ray diffraction lines (to
A a / a ~ 10-3) observed by G6bel (1976) in CdCr2S4 and other sulphospinels below
200 K. This broadening is not correlated with the magnetic behaviour of the
substances, and G6bel (1976) concludes that the spinel structure becomes rather
soft at lower temperatures so that weak random stresses can give rise to
remarkable random strains.
The ferromagnetism of CdCr2S4 was discovered by Baltzer et al. (1965) and
Menyuk et al. (1966). In the paramagnetic state the Curie constants are close to
the value 3.75 for trivalent Cr ions (table 11). However, the moment in the
saturated ferromagnetic state is often found to be less than the expected 6 txs per
molecule (table 11). The observed moments depend strongly on the stoichiometry
and the purity of the samples. For example at 1.5 K the single crystals grown by
Von Neida and Shick (1969) have 3.01/xB per Cr ion in the ferromagnetic
saturation.
The asymptotic Curie temperature 0 is appreciably higher than the ferro-
magnetic Curie temperature (table 11). Baltzer et al. (1966) have analyzed this on
the basis of a simplified model for the interactions between the Cr 3+ moments (see

10.28

10.27

*d 10.26
113
g 10.2s
u

.~ 10.24
O
J

10.23 I
I iiii i i
10.22
/

I I I I I I I
0 100 200 300 400
500 600 700
T (K)
Fig. 36. Lattice constant of CdCr2S4 versus temperature, according to Martin et al. (1969) (I), Bindloss
(1971) (--) and G6bel (1976) (O).
SULPHOSPINELS 655

section 5.1). They find J/k = 11.8K for the nearest-neighbour interaction
- 2 J S i ' N and K / k = - O . 3 3 K for the next-nearest-neighbour interactions
- 2 K 8 i . ~. These values should be regarded with some caution since the results
are expected to be very sensitive to the suppositions about the more-distant-
neighbour interactions (Dwight and Menyuk 1967).
Under hydrostatic pressure the Curie temperature decreases at a rate dTJdP =
-0.58 K/kbar, which corresponds to the value dTJda = +46 K / A for the rate of
increase of T~ with the lattice parameter a (Srivastava 1969).
The dependence of the magnetic moment on the details of the composition has
been investigated by Pinch and Berger (1968). Annealing a polycrystalline sample
with a magnetic moment of 5.79/xB (at 4.2 K and 10 kOe) for 46 h between 800
and 900°C in a sulphur pressure between 30 and 60 atm, they observed a small
increase to a magnetic moment of 5.84/XB, while a reduction treatment in
hydrogen at 800°C for 72h resulted in a smaller decrease to 5.73#B. They
attribute these variations to Cr 2+ ions that charge-compensate the sulphur
deficiency and that couple their magnetic moments of 4~B antiparallel to the
moments of the Cr 3+ ions. They obtained larger variations by replacing Cd z+ with
In 3+. At 4.2 K and 10 kOe polycrystalline samples of Cdl-xInxCr2S4 with x up to
0.15 showed a magnetic moment per molecule that decreased at a rate close to
--7XtXB, corresponding to the replacement of 3 tXB of Cr 3+ parallel to the mag-
netization by 4 txB of C1a+ antiparallel to the magnetization (fig. 37). Pinch and
Berger (1968) observed that the rate of approach to saturation was smaller the
larger the In concentration, which they attributed to an increasing magnetic

-5
o
-'X.
-6
E \\'~ ^Cdl_ x InxCr 2 $4
~5
\ \
Z \ o\
E \
0
~E \
\
\
CdCr2-xInxS~ \
\
\
3 \
\
=F i I I [
0 0.1 0.2
X
Fig. 37. Magnetic moment of Cdl xInxCr2S4 (Pinch and Berger 1968) and of CdCr2-xInxS4 (I.otgering
and Van der Steen 1971a) at 4.2 K and 10 kOe.
656 R.P. VAN STAPELE

anisotropy, expected for Cr 2+. These results were in keeping with the existence of
Cr 2+ in the inverse spinels CrIn2S4 and CrAlaS4 (Flahaut et al. 1961). However,
Lotgering and Van der Steen (1969, 1971a, b) in a study of the paramagnetic
properties of the latter compounds, found that the Curie constant corresponded to
Cr 3+ instead of Cr 2+ and showed that the single phase compounds were the
metal-deficient spinels Crs/9M16/9[[]l/3S4
3+ 3+ (M = A1 or In). Moreover, they discovered
that the magnetization of CdCrzS4 decreased surprisingly fast if they replaced
magnetic Cr 3+ ions by diamagnetic In 3+ ions in CdCr2_xInx
3+ 3+$4 (fig. 37). It is seen
that the magnetic behaviour is comparable to that in Cdl_xInxCr2S4, though no
Cr 2+ is present. The decrease of the ferromagnetic moment with x has been
attributed by Lotgering and Van der Steen (1971a, b) to spin canting around the
In ions due to negative exchange interactions between next-nearest Cr neighbours
(see section 5.1)*.
Two conclusions can be drawn from these experiments. The first is that the
divalent state of chromium is not stable in CdCrz84. The second is that the
ferromagnetic ordering in CdCr2S4 is easily disturbed. This last property makes it
dangerous to use the magnetization of substituted CdCr2S4 as a measure of the
magnetic moment of the substituted ions, as Robbins et al. (1969) had done.
These authors prepared the spinels Cd0.sIn0.zCrl.80M0.2S4 with M = Co or Ni and
CdCrl.sM0.2S4 with M = Ti and V. From the change in magnetization they con-
cluded that the coupling between the spins of Ni 2+ and Co 2+ and the Cr 3+ spins is
negative, while Ti 3+ and V 3+ appear to have no magnetic moment.
Not only the magnetic moment, but also the ferromagnetic resonance spectra
have been found to depend on the sample (Berger and Pinch 1967). Pinch and
Berger (1968) studied the 9.49 G H z ferromagnetic resonance spectra of vapour-
grown single crystals at 4.4 K. In the "as-grown" state the angular dependence o f
the field for resonance was anomalous with sharp peaks in the [111] directions,
while the line width was also anisotropic with a sharp maximum in the same
direction. The angular variation could not be described by the usual cubic
anisotropy constants K1 and /(2. Heating of the crystals in a sulphur p r e s s u r e
reduced the anisotropy and produced samples with a small K~ = 3.8 x 103 erg/cm 3
and K2 = 1.3 x 10 3 erg/cm 3 and an isotropic line width. On a single crystal, exposed
to air for 1½ years, Arai et al. (1972) measured at 9.5 G H z and 4.2 K an angular
variation of the resonance field, corresponding to K1 = 1.6 x 10 4 erg/cm 3, and from
the shift of the resonance fields while applying a uniaxial stress along the [110]
direction they measured magnetostriction constants Am = - 2 . 9 x 10 -5 and h~oo=
- 4 . 7 x 10-5. The observed variations in the magnetic anisotropy were attributed to
the action of Cr 2+ ions, which could be present in non-stoichiometric crystals.
However, Hoekstra and Van Stapele (1973) recognized the similarity between the
angular variation of the anomalous resonance spectra and their spectra of Fe 2+ doped
CdCr2S4 (Hoekstra et al. 1972, Hoekstra 1973). They showed that a small amount of
7 x 10 .4 tetrahedrally coordinated Fe 2+ per formula unit can explain the resonance
spectra of the "as-grown" crystals of Pinch and Berger (1968). Hoekstra and Van
Stapele (1973) also showed that Cr a+ ions on octahedral sites in the spinel lattice are

* See notes added in proof (e) on p. 737.


SULPHOSPINELS 657

expected to give rise to peaks in the resonance field and the line width in the [110] and
[112] directions, unless a static or a dynamic Jahn-Teller effect quenches most of the
orbital moment. Anomalies in the [110] and [112] directions are not, however,
specific to Cr2+; they can also be due to ions, such as Fe z+ on octahedral sites. In the
ferromagnetic resonance spectrum at 9.5 G H z and helium temperatures of an
"as-grown" single crystal of CdCr2S4 Hoekstra and Van Stapele (1973) actually
observed small peaks in the [110] and [112] directions, additional to strong anomalies
in the [111] directions, due to 7 x 10-5 tetrahedral Fe 2+ ions per formula unit. If the
small peaks are due to Cr 2+ ions, their concentration is low: 2 x 10 -5 ions per formula
unit.
Single crystals of CdCr2S4, prepared from very pure starting materials, such that
the Fe concentration was too small to be detected spectrochemically (<0.0006%
by weight), had a weak anisotropy (K1 = +1 x 103 erg/cm 3) and Alll = - 2 × 10 -5,
whereas A100 was found to be negative and one order of magnitude smaller
(Hoekstra 1974). These values are in reasonable agreement with theoretical
estimates of the magnetic properties of octahedral Cr 3+ ions (Hoekstra 1974).
Studies of the paramagnetic resonance have been reported for polycrystalline
samples (Samokhvalov et al. 1973, Stasz 1973, Shumulkina 1975, Shumilkina and
Obraztsov 1975) and for single crystals (Krawczyk et al. 1973, Kaczmarska and
Chelkowski 1977, Zheru et al. 1978). The g factor was observed to have a
temperature-independent value close to 1.99 (Samokhvalov et al. 1973, Krawczyk
et al. 1973, Kaczmarska and Chelkowski 1977). With the temperature decreasing
towards the ferromagnetic Curie temperature the line width starts to decrease and
passes a sample-dependent minimum, after which the resonance line broadens as
the temperature approaches the Curie temperature. The increase of the line width
at higher temperatures is attributed to Raman processes involving local phonons
(Krawczyk et ak 1973).
CdCrzS4 is a semiconductor. The cold-pressed samples studied by Lutz and
Grendel (1965) had a p-type conduction, whereas the high density (99.6%)
polycrystalline samples used by Lehmann and Robbins (1966) had an n-type
conduction with a Seebeck coefficient of -601xV/K and a room temperature
conductivity of 5 x 10-4 (Ocm) -1. The log o- versus T -~ plot was slightly curved, so
that an activation energy could n o t be given, and did not show discontinuities
around the Curie temperature. Like the samples of Lehmann and Robbins the
polycrystalline hot-pressed samples studied by Larsen and Voermans (1973) did
not show discontinuities in the conductivity around the Curie temperature either.
These authors also measured the magnetoresistance and found only a slight
decrease of the resistivity (less than 2 percent in a magnetic field of 7 kOe) at
temperatures around the Curie temperature. Heat treatment of the samples in a
high sulphur pressure led to high resistivities, whereas heat treatment in a low
sulphur pressure decreased the resistivity and lowered the activation energyl
which can be understood in terms of annihilation and formation of sulphur
vacancies, that give rise to shallow donors.
Samples of Cd0.98Ga0.02Cr254 have an n-type conduction and show in the ferro-
magnetic state an anomalous decrease of the resistivity with decreasing tem-
perature (Bongers et al. 1969, Larsen and Voermans 1973). The maximum in the
658 R.P. V A N S T A P E L E

resistivity around the Curie temperature, which combines with a large and
negative magnetoresistance in the same temperature region, has been ascribed by
Bongers et al. (1969)to scattering of electrons in a broad band by a disorder in the
Cr spin system and to a change in the concentration of charge carriers due to spin
splitting of the conduction band (see section 5.1). However, Larsen and Voermans
(1973) showed that the temperature at which the resistivity and the mag-
netoresistance are maximum depends on the heat treatment of the sample (figs. 38
and 39) and they concluded that the conduction in the ferromagnetic state is
determined by electrons in an impurity band formed by Ga 3+ levels, whereas at
high temperatures the conduction is dominated by electrons ionized in a broad
conduction band. The impurity conductivity decreases with increasing tem-
perature, because of the increasing stabilization of the magnetic Ga impurity state
(see section 5.1), and the resistivity will be maximum at the temperature where

1012 ~ b

1010
103/Tc
(3

108 J
u

> 106
to
0)
y
Or"

10 ~

102 C

100
Z, 6 8 10 12 1/-, 16
103/T (K-1)
Fig. 38. Resistivity of CdCr2S4 (a) annealed in a low sulphur pressure and of Cd0.98Ga0.02Cr2S4 (b) a
highly compensated sample after annealing in a low sulphur pressure and (c) another sample without
annealing, according to Larsen and Voermans (1973).
SULPHOSPINELS 659

1.0

0.8

~- 06 e

0.1.,

0.2

I I t I I I
0 60 80 100 120 1L0 160
T(K)
Fig. 39. Magnetoresistanceof Cd0.98Ga0.02Cr2S4, accordingto Larsen and Voermans (1973)((b) and (c) see
caption fig. 38).

the conduction mechanism shifts from impurity conduction to band conduction.


Although no definite explanation of the phenomena has yet been given, it is clear
that the electrical transport properties show signs of an exchange coupling
between the spin of the charge carriers and the Cr spins. Such effects have also
been observed in optical properties, such as absorption and reflection spectra,
photoconduction, magneto-optic spectra and Raman scattering, to which we will
now turn our attention.
A b o v e 2 eV the optical absorption of CdCr2S4 rises steeply. Towards longer
wavelengths the absorption edge has a long tail with a complicated temperature-
dependent structure (Busch et al. 1966), which also depends on the doping and the
stoichiometry (Miyatani et al. 1971b). The absorption coefficient in the tail is so
large that the optical density of the 4 x 10_3 cm thick single crystal studied by
Harbeke and Pinch (1966) increased at room temperature to log(Io/I)= 3 at
1.57 eV. The position of the " e d g e " shifts to higher energies at lower tem-
peratures, while a structure between 1.6 and 1.7eV develops most strongly
between say 120 and 50 K. In the literature this behaviour has been known as the
"blue shift of the band edge", but it was soon recognized that part of the
absorption at energies below 2 eV is due to crystal field transitions of Cr 3+ and
that the transitions between the valence band and the conduction band occur at
higher energies (Berger and Ekstrom 1969, Wittekoek and Bongers 1969, 1970).
All the lower energy crystal field transitions

4A2g~ 2Tlg + 2Eg' 4A2g i.) 4T2g' 4A2g__+2T2g' 4A2g~ 4Tlg

have been observed. Their energies are listed in table 12 and we have indicated
them at the left hand side of fig. 40. The transition 4A2g---~4Tlg has only been
observed in the polar magneto-optic Kerr effect spectrum (Wittekoek and Rin-
zema 1971), where it occurs at the same energy (2.29eV) as in the absorption
spectrum of Cr-doped CdIn2S4 (table 12). The other crystal field transition
observed in the magneto-optic Kerr effect is 4A2g---~2T2g , but this transition,
660 R.P. V A N S T A P E L E

T A B L E 12
Crystal field transitions of Cr 3+, observed in CdCr2S4 and CdIn2S4: Cr 3+.

Energy (eV)

Transition CdCr2S4 CdIn2S4: Cr 3+ Reference

4A2g --> 2Tlg , 2Eg 1.64 Harbeke and Pinch (1966)


1.62 Berger and Ekstrom (1969)
1.61 Larsen and Wittekoek (1972)
4A2g -~ 4T2g 1.82 Berger and Ekstrom (1969)
Koshizuka et al. (1978a)
1.76 Larsen and Wittekoek (1972)
1.85 Wittekoek and Bongers (1969, 1970)
4A2g -~ 2T2g 2.14 Berger and Ekstrom (1969)
Koshizuka et al. (1978a)
2.12 Wittekoek and Rinzema (1971)
2.1 Larsen and Wittekoek (1972)
4A2g ~ 4Tig 2.29 Wittekoek and Rinzema (1971)
2.29 Wittekoek and Bongers (1969, 1970)

- 2.60

••,otoconduction edge
2.50

2.LO

Z'A2 ~,,STlg ~-.-.-. 2,30

220
Z'A2-~2T2g ~
2.10
[~ C,C" g
2.o0 c~,
LU
1.90
. . . .

1.80

1.70
4A2-~Tlg ,2Eg~--~ Tc 1.60
'~i I I i
100 200 300 LO0 5OO
T IK)
Fig. 40. Energy of the transitions observed in CdCr2S4 as a function of temperature ( A, B and C,
Berger and Ekstrom 1969, - - - - Wittekoek and Bongers 1969 and 1970, - . . . . . Wittekoek and Rinzema
1971, ,,~ Larsen and Wittekoek 1972, ~ A', A~ and C', Koshizuka et al. (1978a).
SULPHOSPINELS 661

together with the transitions from 4A2g t o *Tzg, 2Tlg and 2Eg, has also been observed
in the absorption spectrum of thin films (less than llxm) (fig. 41, Berger and
Ekstrom 1969, Koshizuka et al. 1978a) and in the photoconduction of poly:
crystalline undoped n-type samples (fig. 42, Larsen and Wittekoek 1972). At 6 K
the latter authors also observed a weak luminescence at 1.6eV due to the
transition 2Tlg , 2Eg-->4A2g (fig. 43). A progression of 330+30cm 1 phonons,
possibly associated with this transition, had been observed in absorption by Moser
et al. (1971).
The lower part of the absorption edge, as observed by Harbeke and Pinch
(1966), consists of the crystal field transition 4A2g--">2ylg , 2Eg and the wing of the
crystal field transition 4Azg~ 4T2e (Wittekoek and Bongers 1969, 1970). The blue
shift of the edge is attributed by Wittekoek and Bongers (1969, 1970) to sharpen-
ing of the 4A2g--> 4T2g transition at lower temperatures. Berger and Ekstrom (1969)
are of the opinion that another transition (A) is also present. It shows a small blue
shift (fig. 43) and has a moderate oscillator strength of f ~ 10-3. Because no other
crystal field transitions are expected in this region, Berger and Ekstrom (1969)
conclude that A may be due to an indirect band-to-band transition or to a weak
charge transfer transition.
Below the Curie temperature the spectrum in the region around 2 eV shows
even more structure. A peak C appears that shifts to lower energies with
decreasing temperature (Berger and Ekstrom 1969, Wittekoek and Bongers 1969,
1970). The oscillator strength is small ( f ~ 10-4) and the absorption is strongly
circularly polarized (fig. 41, Berger and Ekstrom 1969). In spite of the small
oscillator strength, the transition has a considerable Kerr rotation (Wittekoek and

1.5
Z,A2
~¢- 1.3

.2 1.1
O
I C A2 /.i/
0.9

0.7
Y... /
_.// "~..~J" Am
.~' ,,,'"- ----- + I
/ -1
0.5 ¢ --- 0
I
,"'I j I I .r I I I 21.1 I
1.7 1.8 19 2.0 2.2
Energy (eV)
Fig. 41. Circularly polarized absorption spectra of CdCr2S4 at 2 K in a magnetic field of several kOe,
according to Berger and Ekstrom (1969). The m = +1 spectra are displaced by 0.12 to higher optical
density.
662 R.P. V A N S T A P E L E

Energy (eV)
2.5 2.0 1.5
I I E I i

z' A2-.-~2 T2

T12E
cD 2
r~
&
LL
D
O
o 1
~5
r-
o_ ~A2-~4T2 L

I L I I

0.5 0.6 0.7 0.8


Wavelength (microns)

Fig. 42. Photoconductivity of a polycrystalline undoped n-type sample of CdCr2S4 as a function of the
wavelength in zero magnetic field and in a transverse magnetic field of 6 kOe, according to Larsen and
Wittekoek (1972).

Energy (eV)
3.0 2.5 2.0 1.5 1.2 1.0 0.8
i i i I i i E i i

4A2--~4T2

2.0
/ / ' 4 A 2 - ~ 2 T 2 ~ 4^ 2T 2 E-

"- V i ~IR line of Cr3+ A


41.0:/o / / i. 200 x enhanced /

g
E
i i I

0.4 0.6 0.8 1.0 1.2 1.4 1.6


Wavelength {microns)
Fig. 43. Luminescence at 6 K (--) and excitation spectrum at 80 K (- . . . . . ) of undoped CdCr2S4
powders, according to Larsen and Wittekoek (1972).
SULPHOSPINELS 663

Rinzema 1971). These authors also found the position of the structure in the Kerr
effect spectrum to depend on heat treatment and doping. This indicates that peak
C is not due to an intrinsic excitation. A possible explanation is a transition from
the valence band to an F-centre-like state of an electron bound to a sulphur
vacancy (Harbeke and Lehmann 1970, Lehmann et al. 1971, Natsume and
Kamimura 1972).
In a repetition of the absorption measurements reported by Berger and
Ekstrom (1969), Koshizuka et al. (1978a) found essentially the same structure in
the region around 2 e V (see fig. 40). Apart from the peak C, however, these
authors found a second red shifting peak A2, which was also observed by Berger
and Ekstrom (1969) (fig. 41) at 2 K and which has the opposite circular polariza-
tion of peak C. Koshizuka et al. (1978a) interpret A2 and C as components of a
transition to an exchange split state, separated at 4.2 K by 0.07 eV.
The onset to strong absorption occurs at B (figs. 40 and 41). Berger and
Ekstrom (i969) attribute this edge to transitions between the valence band and
the conduction band. Its position 2.3 eV at low temperatures agrees well with the
band gap of CdS (2.58eV) and of CdIn2S4 (2.2 eV). The direct edge in the
photoconduction has been observed by Larsen and Wittekoek (1972) to occur at
the slightly higher energy of 2.5 eV (fig. 40). The edge shifts weakly to higher
energies at lower temperatures and shows no trace of a shift to the red in the
ferromagnetic state.
At still higher energies a transition at 3.4 eV has been found by Wittekoek and
Rinzema (1971) as a strong resonance in the polar Kerr effect spectrum. This
transition, which is responsible for the dispersive part of the Faraday rotation
between 0.8 and 10 ~m (Wittekoek and Rinzema 1971, Moser et al. 1971) has
been attributed by Wittekoek and Rinzema to a charge transfer transition of an
electron from sulphur p orbitals to an empty Cr orbital.
Reflection spectra (Wittekoek and Rinzema 1971, Fujita et al. 1971, Ahrenkiel
et al. 1971), thermoreflectance spectra (Iliev and Pink 1979) and measurements of
the reflectance d i c h r o i s m - o n e measures in a magnetic field 2 [ ( R + - R _ ) / ( R + +
R )], where R+ and R_ are the specular reflectivities for right-handed and
left-handed circularly polarized l i g h t - (Ahrenkiel et al. 1971, Pidgeon et al. 1973)
have provided data that generally agree with the results reviewed above. This is
also the case with the Faraday rotation of thin films of CdCr2S4, measured by
Golik et al. (1976) in the visible region.
Values of the refractive index have been derived by Wittekoek and Rinzema
(1971) from the reflectivity and from interference patterns in thin hot-pressed
polycrystalline samples by Moser et al. (1971) and Lee (1971), who measured
n = 2.8 in the wavelength range of 2-10 Ixm. The data, which agree roughly, are
given in fig. 44. Additional data were reported by Pearlman et al. (1973).
T o conclude this review of the spectral properties of CdCr2S4, we return to the
study of the photoconductivity reported by Larsen and Wittekoek (1972). The fact
that they observe the crystal field transitions in the photoconductivity (fig. 42)
reveals that excited C r 3+ ions are not stable with respect to either an indirect
minimum of the conduction band or a direct minimum, to which optical tran-
664 R.P. V A N S T A P E L E

L
c
o
1

E~

"53 ~ X ~ x ~ . x ~ x _ _

'2
I [ , I I I
100 500 1000 5000
W o v e l e n g t h (nm)
Fig. 44. Refraction index of CdCr2S4; ( 0 ) data of Wittekoek and Rinzema (1971) at 80 K; (x) data of
Moser et al. (1971) at 300 K.

sitions from the valence band are forbidden, at 1.6 eV or less above the valence
band. The destabilized Cr ions give rise to holes in the valence band, which
accounts for the observed photoconductivity. Larsen and Wittekoek (1972)
observed that the photoconductivity depended strongly on the strength of an
applied magnetic field (fig. 42), which they attribute to a hole mobility influenced
by spin disorder scattering (Haas 1968). A strong fluorescence line observed at
0.9eV (fig. 43), which has an excitation spectrum showing the crystal field
transitions, indicates that the holes can recombine with electrons from an acceptor
state at 0.9 eV above the valence band (Larsen and Wittekoek 1972). Finally we
mention that the photoconduction experiments have been extended to Ga-doped
CdCr2S4 samples (Larsen 1973) and that the results corroborate the conclusion
that the photoconduction is mainly due to photo-excited holes, which are respon-
sible for the observed dependence on magnetic field and temperature.
We now turn to the phonon structure of C d C r 2 S 4 . White and DeAngelis (1967)
have shown that in normal spinels four vibrations (Tlu) are infrared and five
(Alg + Eg + 3T2g) Raman-active. The four Tlu vibrations have been observed in
infrared absorption and reflection studies and the observed frequencies at room
temperature are given in table 13". Below the ferromagnetic Curie temperature at
79 K, the phonon frequencies are increased by only about 1%, while the oscillator
strength of the highest two vibrations have not changed markedly (Lee 1971). This
is at variance with the Raman-active modes, some of which show a strongly
temperature-dependent intensity. Harbeke and Steigmeier (1968) were the first to
observe this and by scattering light quanta of 1.96 eV they observed the Raman
lines listed in table 14 (Steigmeier and Harbeke 1970). All lines were found to
have an intensity that varies with temperature. This has to do with the tem-
perature-dependent absorption at 1.96 eV, which is the only cause of the variation
of the intensity of the light scattered by the Eg vibration (Raman line C), whereas

* See notes added in proof (f) on p. 737.


SULPHOSPINELS 665

TABLE 13
Frequencies of the four infrared-active phonon modes of CdCrzS4.

Frequencies (cm-1)
at room temperature Reference

381 332 240 97.0 Lutz (1%6), Lutz and Feh6r (1971)
385 337 Riedel and Horvath (1969)
376.9-+0.2 321.6-+0.3 Lee (1971)
385 347 234 Moser et al. (1971)

TABLE 14
Raman lines of CdCr2S4, quoted from Steigmeier and Harbeke (1970).

Assignment Raman shift (cm-1) Rel. intensity


Line 300 K 40 K 300 K 40 K 40 K

A mixed mixed 101 -+2 105 _+2 12


C Eg Eg 256 -+2 257 + 2 44
D T2g T2g 280 -+ 1.5 281 -+ 1.5 49
E mixed mixed 351 _+2 353 -4-2 74
F Alg mixed 394 -4-2 396 -+2 112
G -460 6
H -506 5
I -600 6

it is only partly the origin of the variation of the intensities of the other R a m a n
lines. W h a t has strongly attracted attention was the observation of Steigmeier and
H a r b e k e (1970) that the t e m p e r a t u r e variation of the ratio of the intensity of the
lines A, D, E and F to the intensity of line C resembles that of the correlation
function (Si • S j ) / S 2 of n e a r e s t - n e i g h b o u r Cr spins Si and S/. H o w e v e r , K o s h i z u k a
et al. (1976, 1977a) subsequently s h o w e d that the way in which the R a m a n
intensity varies with t e m p e r a t u r e d e p e n d s on the wavelength of the scattered
light. A b o v e the ferromagnetic Curie t e m p e r a t u r e m a r k e d resonances are not
observed, whereas at 15 K especially the R a m a n lines E and F show a strong
r e s o n a n c e at 650 rim, which m e a n s that light q u a n t a are maximally scattered if
their energy is a b o u t 1.9eV. A s can be seen in fig. 40, this is the wavelength
region w h e r e the red-shifting absorption has been observed. T h e absorption in
this region is strongly circularly polarized and Koshizuka et al. (1978a, b) o b s e r v e d
that the R a m a n scattering of circularly polarized light d e p e n d e d systematically on
the sense of the polarization of the incoming and the scattered light if the sample
was placed in a magnetic field. These experiments were d o n e at 40 K with a Kr
laser at 1.92 e V and with an A r laser at 2.01 e V in an a t t e m p t to observe the
influence of the right-handed circularly polarized absorption C and the left-
h a n d e d circularly polarized absorption A2 (figs. 40 and 41). A t 1.92 e V the effects of
the polarization were m u c h m o r e p r o n o u n c e d than at 2.01 e V but neither spectra
s h o w e d a c o m p l e t e right-left s y m m e t r y (see also Koshizuka et al. (1980)).
666 R.P. VAN STAPELE

An analysis of the phonon frequencies in terms of a simple force model has


been given by Briiesch and D'Ambrogio (1972). These authors as well as
Baltensperger (1970) and Steigmeier and Harbeke (1970) discussed the influence
of magnetic ordering on the Raman and infrared-active phonons on the basis of
the ion position dependence of the superexchange interaction. Suzuki and
Kamimura (1972, 1973) formulated a phenomenological theory of spin-dependent
Raman scattering, starting from the general form for a polarizability tensor that
depends on the Cr spins. They obtained an integrated intensity I ( T )
IR + M ( S o . S~)/$212 with a temperature dependence that is determined by the
signs of R and M and their relative magnitude. The values of R and M can be
calculated for specific microscopic Raman scattering mechanisms. Suzuki and
Kamimura (1973) did this for phonon-modulated transfer integrals and oil-
diagonal exchange and were able to explain the temperature dependence obser-
ved by Steigmeier and Harbeke (1970), mentioned above. However, the experi-
mental situation turned out to be much more complex and a more definite
theoretical picture will have to await more complete experiment data.
Nuclear magnetic resonance studies of CdCrzS4 have been made by Berger et
al. (1968a, 1969a) and Stauss (1969a, b). The spectrum of 53Cr is complex and has
been analyzed by Berger et al. (1968). They described the spectrum by

v = (y]2rr)Hi~o + [(y]2rr)Ha, is + vo(rn - ½)](3 cos 2 0 - 1),

where u is the frequency, 3' the gyromagnetic ratio of 53Cr, m = 3, ½, _ ½for the
three m = +1 transitions between the I = 3 nuclear spin states of 53Cr and 0 is the
angle between the local trigonal axis of the octahedral site and the magnetization
in the randomly oriented Weiss domains. At 4.2 K the isotropic hyperfine field
Hiso = - 1 9 1 . 0 k O e , the anisotropic hyperfine field Hanis= +2.07kOe and the
quadrupole interaction vQ = eZqQ/4h = 0.95 MHz. Berger et al. (1969a) discussed
the strength of /-/~so in connection with the strength of the nearest-neighbour
superexchange interaction, while Stauss (1969b) stressed the importance of trans-
ferred spin density in the Cr 4s states in the process that lowers Hiso from its
purely ionic value.
The hyperfine field on the 111Cd and H3Cd nuclei of the diamagnetic Cd ions is
large, namely +167.0kOe at 4.2K (Stauss 1969b) and +168.10kOe at 1.4K
(Berger et al. 1969a). This field is mainly due to transfer of spins from the filled Cr
3d states to the unfilled Cd 5s state. A density of 2.0% of an electron spin in the
Cd 5s state can explain the observed strength of the hyperfine field (Berger et al.
1969a, Stauss 1969a, b).

5.4. HgCr2S4

HgCr2S4 is a normal spinel (Hahn 1951). Crystallographic data are given in


table 1.
In the paramagnetic region the magnetic susceptibility follows a Curie-Weiss
law with a molar Curie constant close to the spin-only value 3.75 of Cre (table 11).
SULPHOSPINELS 667

As observed by magnetic measurements (Baltzer et al. 1966) and by neutron


diffraction (Hastings and Corliss 1968a, b) the compound behaves as a metamagnet
with an ordering temperature of 36.0 K or 60 K respectively. At 4 K the magnetic
structure is a simple spiral, with a propagation vector parallel to the symmetry
axis of the spiral and directed along a particular cube edge in a given domain. The
spiral wavelength is 42 A at 4 K. It increases with temperature (fig. 45), reaching a
value of about 90 A at 30 K, and shows little further variation up to the N6el point
(Hastings and Corliss 1968a, b). Application of a magnetic field up to 4 kOe along a
cube edge produces a growth of the domains in which the propagation vector is
parallel to the magnetic field. In higher fields the spiral collapses in the field
direction, saturating at about 10 kOe (Hastings and Corliss 1968a, b). These obser-
vations are consistent with the magnetization curves (fig. 46) measured by Baltzer

100

c-

8O
c

o
[D
60
-8
t_
5.
O3

4O
10 20 30 40
T(K)
Fig. 45. Temperature dependence of the spiral wavelength of HgCr2S4 (Hasting and Croliss 1968b).

°1
~5

0
0 2 4 6 8 10
Applied field H (kOe)
Fig. 46. Magnetizationof HgCr2S4 as a function of applied field and temperature (Baltzer et al. 1966).
668 R.P. V A N S T A P E L E

et al. (1966). The saturation magnetization measured by these authors at 4.2 K and
10 kOe corresponds to 5.35 p~B/molecule. This is lower than the theoretical 6/~B,
probably for reasons of stoichiometry.
Srivastava (1969) estimated the shift of the transition temperature with pressure
by measuring the relative shift of the mutual inductance as a function of
temperature and pressure. He did not observe anomalies around 60K, but
reported the ordering temperature to be 36.1 K, in agreement with the value of
36.0 K given by Baltzer et al. (1966). This temperature increases with
o
pressure at a
rate dTJdP = +0.14K/kbar, which means that dTJda = - 1 0 K/A. The pressure
dependence, however, differs in sign from that of the Curie temperatures of
CdCrzX4 and HgCrSe2 (Srivastava 1969).
There have been hardly any measurements of the electrical properties. Baltzer
et al. (1965) have reported that HgCrzS4 is a semiconductor.
The behaviour of the absorption edge of HgCr2S4 has been studied by Harbeke
et al. (1968) and Lehmann and Harbeke (1970). Using relatively thick samples (33
and 20 p~m) they did not observe any structure nor any change in the shape of the
edge between 4 K and 600 K. The position of the edge, defined by the energy at
which the absorption coefficient had risen to 1500 cm -1, showed a large red shift
and a further shift to longer wavelength if a magnetic field was applied (fig. 47).
Lehmann and Harbeke (1970) explain the temperature dependence and the
magnetic field dependence in terms of the properties of the spiral spin structure
and the nearest-neighbour spin correlation function. They also note that the
magnetic field-induced shift at 8 kOe has a maximum at (60.0-+ 1) K, which they
take to be the true N6el temperature.
The infrared absorption spectrum of HgCrzS4 has been measured by Lutz (1966)
and Lutz and Feh6r (1971). At room temperature four infrared-active vibrations
were observed at 376, 336, 227 and 71.4 cm-L

1.5

1.L H =0/.f'" . . . . . ~"-- .... .


@
/"
,m
u 1.3 /
1233

~ ~.2
c

1.1 i ! : 8kOe

1.0
[ I I I ~ I
0 100 200 300 400 500 600
Temperature (K)
Fig. 47. Energ y gap of HgCr2S4 as a function of t e m p e r a t u r e for H = 0 and H = 8 kOe, according to
L e h m a n n and H a r b e k e (1970).
SULPHOSPINELS 669

Hyperfine fields on various nuclei have been measured by Berger etal. (1969a, b).
At 1.4 K the isotropic hyperfine field on the nuclear spin of 53Cr w a s found to be
-189.9 kOe (Berger e t a l . 1969a).
The nmr lines of mgHg and 2°1rig were found to be very anisotropic, which is
due to a reduction of symmetry by the spiral spin structure of HgCr2S4. The
isotropic hyperfine field at the Hg nuclear spins, which corresponds to the high
frequency peak, amounted to +524.3 kOe at 1.4 K (Berger et al. 1969a), while the
centre of the Hg spectra was found at 507 kOe (Berger et al. 1969b). This means a
spin density of 2.0% for an electron spin in the Hg 2+ 6 s state (Berger et al. 1969a).
The same spin density was found in the 5s state of Cd 2+ in CdCr2S4 (see section
5.3). The much larger hyperfine field on the Hg nuclear spins is due to the larger
amplitude of the 6s function on the site of the Hg nucleus.

.5.5. ZnCr2Se4

ZnCr2Se4 is a cubic spinel (Hahn and Schr6der 1952). Values of the lattice
parameters are given in table 1. Using neutron diffraction, Plumier (1965) found
the compound to be less than 1% inverse. The cell edge has been measured as a
function of temperature down to 4 K by Kleinberger and De Kouchkovsky (1966).
The compoufid is cubic down to 20.4 K. Below that temperature it shows a small
tetragonal distortion (fig. 48), which is connected with the magnetic ordering.
Magnetic data are summarized in table 11. Above 300 K, in the paramagnetic
state, the magnetic susceptibility follows a Curie-Weiss law with a molar Curie
constant Cm = 3.54 and an asymptotic Curie temperature 0 = +115 K (Lotgering
1964b).

Ce(l edge (~,1

acubic
io.L8~ /

If I I I I I I
810 L i l i
0 20 AO 60 100 120
Temperature {K}
Fig. 48. Cell edge of ZnCr2Se4 as a function of temperature, according to Kleinberger and De
Kouchkovsky (1966).

In spite of the positive 0, the Cr spins order antiferromagnetically at 20 K as


indicated by the minimum in the reciprocal susceptibility, shown in fig. 49.
Subsequent neutron diffraction experiments by Plumier (1965) showed that the
magnetic structure is a simple spiral with a propagation vector along the symmetry
axis of the spiral and directed along a [001] axis. The turning angle between the
magnetic moments in adjacent (001) Cr planes varies from 42_+ 1 degrees at 4.2 K
(Plumier 1965) to 38 degrees at 2 1 K (Plumier et al. 1975b). Plumier (1966a, b)
studied the influence of magnetic felds up to 15 kOe on the neutron diffraction
670 R.P. VAN STAPELE

300 6

2OO
O0 I L~O~
20

100

l S
/

J
f
If I I I I I I I
O0 200 z,o0 600 800
Temperature (K)
Fig. 49. Molar magnetic susceptibility of ZnCr2Se4, according to Lotgering (1964b).

spectrum. His study shows that in zero magnetic field domains occur in which the
helical axis is oriented along one of the cube edges. In magnetic fields up to 3 kOe
the domains with a preferred orientation of the spiral axis grow. In higher fields a
conical spin structure exists with an increasing net magnetization. The mag-
netization has been measured at 4.2 K in magnetic fields up to 30 kOe and 69 kOe
by Lotgering (1965) and Siratori (1971) respectively and in pulsed magnetic fields
up to 108 kOe by Allain et al. (1965). The magnetization, shown in fig. 50,
saturates at 6.! #B/molecule in fields larger than 64kOe. Hydrostatic pressure
shifts the N6el temperature to higher values at the rate dTN/dp = 0.90 x 10-3 K/bar
(Fujii et al. 1973).
As Lotgering (1964b) has stressed, the combination of a positive asymptotic
Curie temperature and an antiferromagnetic behaviour is a clear indication of the
important role of the more distant exchange interactions in ZnCr2Se4. Attempts
have been made to extract quantitative information about the strength of the
various exchange interactions from the paramagnetic data, the low temperature
spin structure and its initial magnetic susceptibility. Lotgering (1965) and Plumier
(1966a, b) have made such an analysis, assuming J5 = 3"6= 0 (see table 10, and fig.
35) and -/3 = J4.
Lotgering (1965) used the experimental values of the turning angle and the
asymptotic Curie temperature and calculated J1, J3 and the initial susceptibility Xi
as a function of an antiferromagnetic J2 in the range - 7 < Jz/k < 0 K. Under these
circumstances J1/k < 3 0 K , -5<J3/k < O K and Xi has a value close to the
experimental one.
Plumier (1966a, b) used the same experimental data, but calculated the al-
SULPHOSPINELS 671

:D
(3
d~
O
E 3

3
E 2
E
O
~E
1

I I I I I I I
20 40 60 80 100
Magnetic field (k0e)

Fig. 50. Magnetization versus magnetic field of powder samples of ZnCrzSe4. Curve (a) magnetization
at 4.2K, according to Lotgering (1965); curve (b) according to Allain et al. (1%5) and curve (c)
magnetization at 4.2 K, according to Siratori (1971).

gebraic solution with a ferromagnetic -/2: J1/k = 24.9 K, J2/k = 7.8 K and J3/k =
-10.65 K. Dwight and Menyuk (1967) analyzed the stability of the spiral spin
configuration, taking into account all the interactions J 1 . . . J6. They showed that
neither of these interactions can be neglected. Lotgering's set of interactions lies
outside the region where the spiral spin configuration is stable. The values of
Plumier are within this region, but Dwight and Menyuk (1967) judge them to be
physically unreasonable from the point of view of the mechanism of the superex-
change interactions. They do not give a better solution, but mention as an
illustrative "zeroth approximation" the following set of interactions: J~/k =
+25.4 K, J2/k = +2 K, J3/k = - 7 K, J4/k = - 7 K, Js/k = +1 K and J d k = - 1 K.
In a similar analysis A k i n o and Motizuki (1971), who used the restriction -/3 = J4
and J6 = 0, came to the conclusion that the stability of the spiral spin configuration
requires a negative J3 and that this stability is increased by a positive .15.
Paramagnetic resonance with g = 2 had been observed in powder samples
above TN = 21 K by Stickler and Zeiger (1968). Below TN, an antiferromagnetic
resonance appeared abruptly in fields stronger than 4 k G , as the paramagnetic
resonance line disappeared. The antiferromagnetic resonance faded out at lower
fields. T h e antiferromagnetic resonance frequency, extrapolated to zero magnetic
field, was found to follow an S = ~ Brillouin function. At the Ndel t e m p e r a t u r e the
zero field frequency drops to zero abruptly.
The magnetic resonance of single crystals has been studied by Siratori (1971) in
the frequency range 38-83 GHz. His theoretical explanation of the frequency and
angular dependence of the resonance field at 5.5 K gives the values of the Fourier
c o m p o n e n t of the exchange and the magnetic anisotropy energy at three points 0,
672 R.P. VAN STAPELE

k0 and 2ko in the reciprocal space (k0 is the spiral wave vector). Transformation to
local exchange interactions via one and two Se ions gives values for J1 • • • J5 that
do not stabilize the spiral spin configuration, from which Siratori concluded that
superexchange interactions through more than two Se ions (like J6) cannot be
neglected. With the temperature increasing from 5.5 K to 20 K Siratori (1971)
observed a shift of the resonance field and a decrease of the angular dependence
and the line width. Above 20 K, the resonance field is independent of the
direction of the magnetic field, but the shift of the resonance field persists up to
about 70 K.
Specific heat measurements done by Plumier et al. (1975a, b) indicate that, in
addition to a fairly large peak from 17 to 23 K with a mazimum at 20.5 K, two
narrower peaks exist at 45.5 K and 105 K. The nature of these extra peaks is not
yet clear. Plumier et al. (1975b), studying carefully the neutron diffraction
spectrum of a powder sample, observed above 20.5 K broad satellites with angular
positions corresponding to a pitch angle ~-/5, indicating that ZnCr2Se4 at such
temperatures is a metamagnet with a centred tetragonal cell (a/V'-i, a/~/-i, 5a).
The large line width is ascribed to the small size of the reflecting magnetic
domains, 53 A ( - 5 a ) between 21 and 45 K and 32 A ( - 5 a / 2 ) between 45 and
105 K. When a magnetic field was applied below 20.5 K, Plumier et al. (1975b)
observed that the sharp satellites of the helical spin configuration disappeared at a
critical field strength and that broad peaks appeared at higher field strengths. The
width and the position of these peaks corresponded to those observed in zero field
above 20.5 K.
The findings of Plumier et al. (1975a, b) prompted Akimitsu et al. (1978) to
study the neutron diffraction of a single crystal and to reinvestigate the differac-
tion on powder samples. They concluded that the broad peaks observed by
Plumier can be ascribed to critical scattering, observed in their single crystal
experiment in the wide temperature range from 20 to 40 K, and found that there
is no reason to place the N6el temperature higher than 21.2K. However, the
peaks at 45.5 and 105 K in the specific heat then remain unexplained. Another
result of their work that is difficult to explain is the enormous deviation of the
magnetic moment of the C13+ ions (1.71/xB) from the 3#B observed in the
magnetic measurements.
The weak temperature dependence of the turning angle (Akimitsu et al. 1978) and
the susceptibility in the propagation direction (Kawanishi et al. 1978) have been
correlated with the tetragonal deformation (Kleinberger and De Kouchkovsky 1966)
mentioned above and shown in fig. 48. The reciprocal of both quantities was found to
depend linearly on ~5= [1 - (c/a)] (Akimitsu et al. 1978, Kawanishi et al. 1978), which
was interpreted as being due to exchange interactions with a strength depending on ~5
and the wave vector near the propagation vector of the spiral spin configuration*.
The influence of stoichiometry and impurities on the magnetic properties has
scarcely been investigated. The single crystals studied by Kawanishi et al. (1978)
had a small residual ferromagnetic moment, varying from crystal to crystal. Small
amounts of Cu (up to 0.04 per formula unit), substituted for Zn, had been

* S e e n o t e s a d d e d in p r o o f (g) o n p. 737.
SULPHOSPINELS 673

reported to lower the N6el temperature from 20 to 11 K, while they decreased the
high temperature magnetic susceptibility (Menth et al. 1972). Replacement of Zn
by Mn had been reported to increase at 4.2 K the magnetization in magnetic fields
up to 80kOe (Siratori et al. 1973). Neutron diffraction experiments on
Znl-xMnxCr2Se4 with x = 0.1 showed that the spiral spin structure is not seriously
changed by the presence of the Mn ions (Siratori and Sakurai 1975). It was
concluded that the Mn-Cr exchange interaction is weakly ferromagnetic. Single
crystals of Mn-doped ZnCr2Se4 have recently been reported to have a strong cubic
paramagnetic anisotropy (Kawanishi et al. 1979).
ZnCr2Se4 is a semiconductor (Lotgering 1964b, Albers et al. 1965). Replace-
ment of 0.02 Zn per formula unit by Cu lowers the resistivity strongly (Lotgering
1968b). This result has been used as an argument for the monovalent state of
copper because the replacement of Zn 2+ by Cu + gives rise to holes in the valence
band.
The miscibility in the series of mixed crystals Znx-xCuxCr2Se4 between the
spinels ZnCr2Se4 and CuCr2Se4, which have about equal cell edges, has been
reported to be limited to x ~<0.05 (Lotgering 1968b). Complete solubility has been
observed in samples quenched from temperatures above 500°C (Okofiska-
Kozlowska and Krok 1978). These samples have a lattice parameter that varies
linearly with x. They are ferromagnetic for x/> 0.2 with a ferromagnetic Curie
temperature that varies slowly from 380 K at x = 0.2 to 415 K at x = 1. This
behaviour is reminiscent of that of the series CuCrzSe4_yBry (fig. 30). In this
system high Curie temperatures occur for 0 ~<y ~<0.8, e.g., as long as conducting
holes in the valence band induce strong ferromagnetic interactions between the Cr
ions (see section 4.11).
The optical absorption edge of ZnCr2Se4 shows a strong shift to longer
wavelengths with decreasing temperature (Busch et al. 1966). In the optical
transmission of thin (14 to 50 ~m) single crystalline platelets Lehmann et al. (1971)
observed that the shape of the edge changes at lower temperatures (fig. 51): a
slight shift to shorter wavelengths between 298 and 181K is followed by the

'rE 2000

-E

1000
00,;
00I/
[I I I L I I
0 1.10 1.20 1.30
Energy (eV)
Fig. 51. Absorption coefficientversus energy of ZnCr2Se4(Lehmannet al. 1971).
674 R.P. V A N S T A P E L E

development of a precursor band, which shifts strongly towards the red with
increasing absorption coefficient. The strength of this absorption was observed by
Lehmann et al. (1971) to vary from sample to sample, indicating that the band has
a non-intrinsic origin. It shows nevertheless the effects of the influence of the
magnetic structure of ZnCr2Se4. Lehmann et al. (1971) observed that in a strong
external magnetic field the position of the peak of the absorption of circularly
polarized light in the precursor band depends on the sense of the rotation (fig. 52)
and on the strength of the magnetic field (fig. 53). At 4.2 K, the latter curve has a

1.4

- - 1.3 H--0
t_
c-
o9
1.2
(3
O)
0-

1.1

1.0 =
/. I

0.9 I I i i I i I ~ I /
0 20 40 60 80 100
Temperature ( K )
Fig. 52. A b s o r p t i o n peak position of ZnCr2Se4 as a function of t e m p e r a t u r e in zero field and in 67 k O e
for circularly polarized light ( L e h m a n n et al. 1971).

1.3

1.2

_~ 1.1
o~

1.0

0.£ t i Li0 i I I I
20 60 80
Megnetic field (kOe)
Fig. 53. A b s o r p t i o n peak position of circularly polarized light of ZnCr2Se4 as a function of magnetic
field strength at 4.2 K ( L e h m a n n et al. 1971).
SULPHOSPINELS 675

shape that is nearly identical to the magnetization curve, as measured by Allain et


al. (1965) and given in fig. 50. Lehmann et al. (1971) proposed that the absorption
in the precursor band be attributed to transitions from a fourfold valence band
state with degeneracy lifted by the magnetic field to a spin-polarized singlet state,
probably connected with Se vacancies.
Finally, Riedel and Horvath (1969) have measured the infrared spectrum of
powder samples of ZnCr2Se4. They observed at room temperature two infrared-
active vibrations at 287 and 304 cm -1.

5.6. CdCr2Se4

CdCr2Se4 is a normal spinel (Hahn and Schrrder 1952). The lattice parameters are
listed in table 1. The cell edge has been measured as a function of temperature by
Martin et al. (1969), Bindloss (1971) and G6bel (1976). The results are shown in
fig. 54. Whereas CdCr2S4 expands anomalously below its ferromagnetic Curie
temperature, CdCr2Se4 is seen to contract in a normal way in the magnetically
ordered state. As in CdCr2S4, anomalous broadening of the X-ray diffraction lines
has been observed in CdCrzSe4 at lower temperatures, but not correlated with the
magnetic phase transition ( G r b e l 1976), which suggests the existence of in-
homogeneous lattice distortions. G r b e l (1976) also observed a shift of the anion
parameter u from 0.389 at room temperature to 0.392 at 77 K, and he remarks
that such a shift of u changes the atomic distances and bond angles considerably.
This can be an important fact with a view to the understanding of other
phenomena observed at low temperatures, like the red shift of the optical band
edge.
Banus and Lavine (1969) have reported that CdCrzSe4 transforms under high
pressure and temperature to a monoclinic structure related to the defect NiAs
structure. The physical properties of the monoclinic structure were found to differ
completely from those of the spinel structure.
CdCr2Se4 is a ferromagnet (Baltzer et al. 1965, Menyuk et al. 1966). In the
paramagnetic region, the magnetic susceptibility obeys the Curie-Weiss law with
an asymptotic Curie temperature of about 200 K and a Curie constant close to the
spin-only value 3.75 of Cr 3+ ions (table 11). The ferromagnetic Curie temperature
is 130 K and the saturation magnetization measured at 4.2 K in a magnetic field of
10 kOe amounts to 5.6/zB per molecule (Baltzer et al. 1966, Menyuk et al. 1966).
The data given by Baltzer et al. (1966), measured on powder samples, differ from
the single-crystal data reported by LeCraw et al. (1967). The latter authors
measured a saturation moment of 2.99 + 0.2 txB per Cr 3+ ion, very close to the
expected 3/xB, but their Curie constant is too high (table 11).
The details of the magnetic properties apparently depend on the purity and the
stoichiometry of the samples, and Pinch and Berger (1968) observed a small increase
in the magnetization of their polycrystalline samples after annealing in excess Se.
The ferromagnetic Curie temperature of In-doped CdCrzSe4 single crystals was
found not to be affected by heat treatments in various atmospheres (Treitinger et
al. 1978a).
676 R.P. VAN STAPELE

10.77

10.76

o~ 10.75

.,.go
g 10.7z,
o

1.3
::- 10.73

'r

10.72 II
I

10.71

10.70~g
I I I I I I
0 100 200 300 400 500 600
Temperature (K)
Fig. 54. Lattice constant of CdCr2Se4 versus temperature, according to Martin et al. (1969) (I),
Bindloss (1971) ( ) and G6bel (1976) (---).

The values of the ferromagnetic and the asymptotic Curie temperature have
been used by Baltzer et al. (1966) to estimate the strength J of the nearest-
neighbour exchange and the strength K of the more-distant-neighbour inter-
actions (see section 5.1), which resulted in J/k = 14 K and K/k = - 0 . 1 0 K.
The critical behaviour of CdCr2Se4 has been studied by Miyatani (1970), who
measured the magnetization at temperatures close to the Curie temperature in
magnetic fields between 6 0 e and 18 kOe. He was able to describe the tem-
perature dependence of the spontaneous magnetization o-0 by o-0 ~ ( T - T 0 ) ~,
where Tc = 127.67 + 0.005 K and /3 = 0.447 + 0.03, and that of the zero-field sus-
ceptibility X0 by ,to ~ (T - T*)% where T* = 127.7 -+ 0.2 K and 7 = 1.27 -+ 0.02. At
the Curie temperature, the magnetic field dependence of the magnetization o- was
given by o- ~ H 1/8 with 6 = 4.1 _+0.2. Within the experimental error, these values
confirm the scaling law prediction y =/3(6 - 1).
U n d e r hydrostatic pressure the Curie temperature decreases. Srivastava (1969)
o
measured dTddp = - 0 . 8 2 K/kbar, which corresponds to dTdda = +66 K/A, using
the measured value of the compressibility. This agrees well with dTddp =
- 0 . 7 6 K/kbar, measured by Fujii et al. (1970). Whereas these measurements were
done with a mutual induction technique under pressures up to 10 and 6 kbar
respectively, Shanditsev and Yakovlev (1975) used the much higher pressures of
60 and 90 kbar and measured the increase of the resonance line width to locate
the Curie temperature. In this way they found dTddp = - 0 . 4 4 K/kbar. It should
SULPHOSPINELS 677

be noted, however, that Banus and Lavine (1969) observed traces of the high-
pressure modification of CdCr2Se4 at room temperature with pressures higher
than some 80 kbar.
The magnetic anisotropy of CdCr2Se4 is rather weak and dependent on the
purity and the stoichiometry of the crystals. From~- ferromagnetic resonance
measurements at 9.49 GHz, Berger and Pinch (1967) found the cubic anisotropy
constant K1 to have values between 2.2 x 103 and 1.8 x 10 4 erg/cm 3 at 4.4 K. The
temperature dependence differed considerably in the various crystals. The
smallest uniform precession line width found was 37 G. A much smaller line width
of 9.5 G was observed at 13 G H z and 4.2 K by L e G r a w et al. (1967). The crystals
used by these authors has a g-factor of 1.983 + 0.003 and a K1 of approximately
1.5 x 104 erg/cm 3 at 4.2 K. Pinch and Berger (1968) observed an increase of the
anisotropy after the crystal had been annealed in a hydrogen atmosphere, and a
decrease to a very small anisotropy after an anneal in a selenium atmosphere. The
resonance line width was isotropic and the angular dependence of the resonance
field was normal. In contrast with these observations, the crystals studied by
Gurevich et al. (1972) showed at 8.9 G H z a ferromagnetic resonance spectrum
with an angle-dependent line width and resonance field. In one crystal the angular
variation was anomalous, and it could not be described by the cubic anisotropy
constants K1 and K2. The anomalous fe~/ture is the strong increase of the
resonance field in the (111) directions, which was also found in the resonance
spectrum of CdCrzSe4 crystals, in which a small amount of Fe 2+ was substituted for
Cd (Gurevich et al. 1975). Although this suggests that contamination with Fe 2+
ions plays a role, Bairamov et al. (1977) prefer to think in terms of the Cr ions.
This is based on the magnetic anisotropy of Ag-doped CdCr2Se4 crystals. Larson
and Sleight (1968) observed that the positive cubic anisotropy of undoped crystals
was strongly reduced to IKII < 0 . 7 x 103 erg/cm 3 in Ag-doped crystals, while K~
actually changed sign in more heavily doped crystals, reaching the value - 1 . 1 x
10 4 erg/cm 3 for 1.34 mole % Ag (Bairamov et al. 1976). When crystals doped with
1.5mole % Ag are annealed in vacuum a positive anisotropy is restored
(Bairamov et al. 1977). The influence of Ag-doping on the ferromagnetic
resonance line width has been investigated by Ferreira and Coutinho-Filho (1978).
Eastman and Sharer (1967) measured the magnetostriction of a single crystal
with a high ferromagnetic saturation moment of 2.97#B per Cr 3+ ion. Ferro-
magnetic resonance at 4 . 2 K and 23 G H z gave a cubic magnetic anisotropy
constant K1 = 3.5 x 103 erg/cm 3. From the shift of the resonance field, caused by a
compressional uniaxial stress, they found A~00=(2.6_+0.5)x10 -6 and Am =
- ( 2 8 + 4 ) x 10-6. From the measured temperature dependence they concluded
that Am is of single ion origin.
Studies of the paramagnetic resonance have been reported for polycrystalline
samples (Shumilkina 1975) and for single crystals (Samokhvalov et al. 1973,
Krawczyk et al. 1973, K6tzler and Von Philipsborn 1978). The g-factor was found
to have a temperature-independent value of 1.99 (Samokhvalov et al. 1973,
Krawczyk et al. 1973), while the temperature dependence of the line width is
characterized by a broad minimum around 200 K. The increase of the line width
678 R.P. VAN STAPELE

at higher temperatures has been attributed to Raman processes involving local


phonons (Krawczyk et al. 1973). The critical increase of the line width, as the
temperature decreases towards the Curie temperature, was found to depend on
the strength of the applied magnetic field, quantitatively confirming the results of
mode-coupling calculations in the limit of zero magnetic field (K6tzler and Von
Philipsborn 1978). The paramagnetic resonance line width increases when Cd is
replaced by Co (Grochulski and Gutowski 1975).
Some magnetic properties of CdCr2Se4 are sensitive to light. Lems et al. (1968)
discovered this so-called photomagnetic effect in Ga-doped CdCr2Se4. After
cooling the crystals in the dark to 77 K (or a lower temperature), they observed
that the low-frequency permeability (10 kHz) decreased by a factor of two upon
illumination with white light. The return to the original higher-permeability state
after the light was switched off was observed to occur at a rate that strongly
depended on the temperature (fig. 55). Lems et al. proved experimentally that the
lower permeability after illumination is due to a change throughout the material
and not to changes local to domain wails. The spectral dependence of the effect
roughly coincides with the absorption edge of CdCr2Se4 at 1 Ixm. In a simple
model that accounts for the experimental data, stable centres I (probably filled Ga
donors) are ionized, giving rise to less stable centres II. The inverse of the change
in permeability is assumed to be proportional to the density n of centres II, while
the recombination rate is proportional to n 2 as a consequence of the assumption
of random recombination (Lems et al. 1968).
The high-frequency permeabilty of pure and Ga-doped polycrystalline samples
has been reported by Veselago et al. (1972a, b) to show a small decrease upon
illumination. The spectral dependence of the effect shows that it starts at the
optical band edge.
A positive change of the ferromagnetic resonance field of illuminated undoped
single crystals has been observed by Salanskii and Drokin (1975). Anzina et al.
(1976) have reported that undoped and Ga- and Ag-doped samples show an in-
crease in coercive force and a decrease in permeability when illuminated with white
light.

light on 77 K
300

200

~ght off
L.2K
E¢ 100
0

13_

510 I I
0 100 150
time Is]
Fig. 55. Influence of illumination by "white light" on the low-frequencypermeability of Ga-doped
CdCr2Se4, according to Lems et al. (1968).
SULPHOSPINELS 679

Finally, Makhotkin et al. (1978b) studied the influence which illumination by


white light and the application of short magnetic field pulses have on the
high-frequency (4 MHz) permeability of Ga-doped CdCrzSe4 at 77 K. In the dark
the permeability is decreased from 62 to 60 during the 5 ms long magnetic field
pulse with an amplitude of 0.08 Oe. Illumination lowers the permeability to 31,
but upon the rise and the decay of the magnetic field pulse the permeability
increases rapidly to 33, decaying to 31 in about 2 ns. This shows that apart from
the photomagnetic changes throughout the material, which reduce the per-
meability from 62 to 31, a photo-induced pinning of the domain walls exists, which
is not apparent from the low-frequency permeability found by Lems et al. (1968).
Before reviewing the electrical transport properties of CdCrzSe4, we mention
the observation by Salyganov et al. (1973) of a DC electromotive force during
ferromagnetic resonance at 77 K of a Ag-doped CdCr2Se4 single crystal. The emf
was observed between gold electrodes in the centre and on the edge of the
single-crystal platelet. The effect occurred not only during uniform precession, but
also during a magnetostatic mode. It was observed that the sign and magnitude of
the emf did not depend on the direction of the external magnetic field along the
normal to the platelet. The mechanism of the effect is not known, but in-
homogeneous heating of the conduction electrons or dragging of the charge
carriers by propagating spin waves have been suggested as possible causes of the
phenomenon.
CdCr2Se4 is a semiconductor (Baltzer et al. 1965). High-density polycrystalline
undoped samples have a resistivity that decreases with increasing temperature and
the resistivity vs temperature curves do not show any discontinuity at the Curie
temperature. This has been observed in undoped samples with p-type conduction
(Lehmann and Robbins 1966, Haas et al. 1967, Lehmann and Harbeke 1967,
Lehmann 1967, Lotgering 1968b) and in undoped samples with n-type conduction
(Larsen and Voermans 1973). Most of the experimental resistivity curves are
regularly curved (fig. 56) contrary to that of an n-type single crystal, which was
observed to have a weakly temperature-dependent resistivity below the Curie
temperature, a region with an activation energy of 0.18 eV between 130 and 310 K
and of 0.6 eV between 350 and 800 K (Prosser et al. 1974).
Substitution of Ag or Cu for Cd gives rise to strong p-type conduction. The
conductivity of Ag-doped CdCrzSe4 decreases monotonically with decreasing
temperature either without an anomaly (Haas et al. 1967) or with a change of
slope of the log o- vs T ~ curve around the Curie temperature (Lehmann 1967)
(fig. 57). The Cu-doped CdCrzSe4 samples studied by Lotgering (1968b) showed a
similar change of slope. The Ag-doped crystals investigated by Coutinho-Filho
and Balberg (1979) showed a slight "hump" in the resistivity around 130 K.
Lehmann (1967) measured the Seebeck coefficient (fig. 58) and the Hall effect of
hot-pressed Ag-doped samples and found the holes to have a large Hall mobility
(10 to 100 cm2/Vs). The observed rapid decrease of the Seebeck coefficient below
170 K and the maximum of the normal Hall coefficient at 150 K were interpreted
by Lehmann (1967) as being due to the onset of impurity conduction at lower
temperatures. The magnetoresistance of Ag-doped CdCrzSe4 is small, with an
109

///a
107

>" 5
> 10
:,,.%
u3

103

10~ I I 1 I I
2 z, 6 20 8 12 1/. 16
IOS/T (K -1 )
Fig. 56. Resistivity of hot-pressed polycrystalline undoped samples of CdCr2Se4: (a) with p-type
conduction, according to Lehmann (1967); (b) with n-type conduction, according to Larsen and
Voermans (1973); (c) sample b after 60 h at 600°C with Pse = 10-4 mm Hg and quenching to room
temperature, according to Larsen and Voermans (1973).

T__ lOq
{3
10-2
>,
:z. 10-3
*d
D

-~
o
~0-~
O

10-s
@/rc
10-6 I I ~'l I I
2 L 6 10 12 1L 8
103/T (K q )
Fig. 57. Electrical conductivity of p-type Ag-doped and n-type In-doped CdCr2Se4, according to
Lehmann (1967).

680
SULPHOSPINELS 681

600

3
oJ
400

200

0
_••/•
i ] i i /c
Tc ~
Ag

"~
o
300 250 200 150 100"
T e m p e r a t u r e (K)

r /

Fig. 58. Seebeck coefficientof p-type Ag-doped and n-type In-doped CdCr2Se4,accordingto Lehmann
(1967).

2.0

0,8

0.6

?o_
0.4

0,2

0100 ~ , I I l i i ....
120 140 160
T (K)
Fig. 59. Magnetoresistance of n-type Ga-doped CdCr2Se4 (2% Ga) in a magnetic field of 12 kOe,
according to Haas et al. (1967).

anomalous magnetic field dependence at lower temperature (Lehmann 1967).


Changes in the sign of the longitudinal magnetoresistance as a function of applied
magnetic and electric fields were observed in Ag-doped single crystals (Balberg
and Pinch 1972), Strong electric fields influence the properties of Ag-doped
CdCr2Se4. The microwave absorption (Solin et al. 1976), the electrical conductivity
(Samokhalov et al. 1978) and the magnetization (Samokhalov et al. 1979)
decrease, which is attributed to the excitation of spin waves by electron-magnon
interaction.
682 R.P. VAN STAPELE

The substitution of In and Ga for Cd gives rise to a strong n-type conduction.


Unlike that of the p-type samples, the resistivity of the n-type samples shows a
pronounced maximum around the ferromagnetic Curie temperature (fig. 57),
while a large and negative magnetoresistance is maximum in that temperature
region (fig. 59) (Haas et al. 1967, Lehmann and Harbeke 1967, Lehmann 1967,
Amith and Gunsalus 1969, Feldtkeller and Treitinger 1973, Merkulov et al. 1978,
Coutinho-Filho and Balberg 1979). The maximum in the resistivity has also been
observed in the high-frequency conductivity of n-type Ga-doped hot-pressed
samples, where it decreases with increasing frequency (Kamata et al. 1972).
Lehmann (1967) and Amith and Gunsalus (1969) also studied the thermoelectric
effect (fig. 58) and the Hall effect, and found that the Hall mobility of the
electrons was of the order of i cm2/Vs, which is much smaller than that of the
holes in p-type samples. The study reported by Amith and Gunsalus (1969)
revealed a crucial anomaly, namely the coincidence at 150 K of a secondary
minimum of the absolute value of the Seebeck coefficient with the maxima of the
resistivity and the absolute value of the Hall coefficient.
Amith and Friedmann (1970) concluded that this finding cannot be explained in
terms of electrons in a single spin split conduction band which are scattered by the
spin-disorder in the ferromagnetic Cr spin system (Haas 1968, Bongers et al. 1969)
and these authors proposed a two-band model in which one band is an n-type
conduction band and the other a p-type hole band in the band gap. As already
discussed at some length in section 5.1, other models have been proposed. In
particular the model of the magnetic impurity states seems to be applicable in the
case of n-type Ga-doped or In-doped CdCr2Se4 (Larsen and Voermans 1973,
Treitinger et al. 1978a). Treitinger et al. (1978a) varied the concentration of Se
vacancies of In-doped single crystals of CdCrzSe4 and observed that the height of
the maximum of the resistivity and of the magnetoresistance, as well as the
temperature at which this maximum occurs, depended on the concentration of the
Se vacancies (fig. 60), whereas the ferromagnetic Curie temperature was not
affected. They ascribed these properties to conduction in magnetic impurity states
at lower temperatures, the conduction at higher temperatures being dominated by
electrons in the conduction band.
In their study of In-doped CdCr2Se4 Treitinger et al. (1978a) also observed that
the line width of the X-ray diffraction lines increases from the value in the
as-grown state after annealing in Se vapour and also after annealing in hydrogen.
Subsequent annealing in hydrogen after a heat treatment in Se vapour restores
the as-grown value, indicating, as Treitinger et al. conclude, that the crystals in
their as-grown state have a Se deficit that corresponds to a state of minimal
internal stress.
The optical absorption spectrum of stoichiometric pure CdCr2Se4 shows no
structure between the bands near 17 txm, which are due to overtones of lattice
vibrations, and the absorption edge near 1 Ixm (Bongers and Zanmarchi 1968),
whereas doped or non-stoichiometric crystals have some characteristic absorption
lines in that region (Miyatani et al. 1971b) (fig. 61). Hl/dek et al. (1977) have
discovered that a similar absorption spectrum can be induced by illumination with
SULPHOSPINELS 683

T
107 300 200 160 120 100
J i i i

105
3000

~103
(3- 400 C

10
L50C

10-1

10-3 I I
5 7 9 11
103/T {K-1)
Fig. 60. Resistivity of n-type In-doped CdCr2Se4 with an increasing number of Se-vacancies, obtained
by a heating in hydrogen at the temperatures indicated in the figure, after Treitinger et al. (1978a).

Photon energy (eV)


0.15 0.2 0.3 O.Z 0.6 1.0 2.0
i i i i i i JEll

.d

c
o)
1.3

c
0 In-doped
0-

undoped
.13
< 110 i i i I I I ~ i
8 6 4 2 0
Wave length (/urn)
Fig. 61. Absorption spectrum of undoped and In-doped CdCr2Se4between 0.1 and 1 eV at 78 K (
and 300 K (---), according to Miyatani et al. (1971b).

light with a wavelength shorter than the absorption edge. T h e absorption edge
shows a structure that has attracted considerable attention. Busch et al. (1966),
w h o m e a s u r e d the diffuse reflectance of polycrystalline samples, and H a r b e k e and
Pinch (1966), w h o m e a s u r e d the absorption of plane-parallel single-crystalline
samples with thicknesses b e t w e e n 15 and 50 Ixm observed that the edge shifts to
longer wavelengths at t e m p e r a t u r e s below 200 K (fig. 62). H a r b e k e and L e h m a n n
(1970) f o u n d that the strength of the absorption at the lower energy edge d e p e n d s
684 R.P. VAN STAPELE

2.10

2.00 ~.

1.90 ~"

Ill

L ,
f-
1.50 \
ILl "%
>,,
1.40 " ~

1.30
/,, -.
1.20 / / ./..~..
t I
1.10 /Tc
¢"~ t i f
0 100 200 300 400 500
T(K)
Fig. 62. Energy of transitions observed in CdCr2Se4 as a function of temperature: ( ) Harbeke and
Pinch 1966; (---) Sato and Teranishi 1970), (-.-.-.) Stoyanovet al. (1976).

on the sample and that only the lowest energy absorption edge is shifted to the
red with decreasing temperature, whereas the higher energy edge is weakly
shifted to shorter wavelengths (fig. 63).
The intensity at the lower energy edge could also appreciably be changed by a
heat treatment (Prosser et al. 1974). Harbeke and Lehmann (1970) concluded that
the red-shifting absorption is not due to an intrinsic excitation but most probably
to vacancy states. Eagles (1978) explained the absorption profiles observed by
Harbeke and Lehmann (1970) at photon energies below 1.6eV in terms of a
combination of transitions between the valence band and hydrogen-like local

13 103

c 103

.~ 130 2911K

/7
<
5.102
I I
1.15 1.20 1,25 1.30 1.35 1.40
Energy (eV)
Fig. 63. Absorption edge of CdCr2Se4 at various temperatures, according to Harbeke and Lehmann
(1970).
SULPHOSPINELS 685

levels and an indirect band-to-band absorption, as suggested by Sakai et al. (1976).


In his analysis Eagles also made use of the lower energy absorption data of
Shepherd (1970) and the data at higher energies of Sakai et al. (1976).
Application of a magnetic field shifts the absorption edge further to the red
(Busch et al. 1966, Lehmann et al. 1971 and Hl~dek et al. 1976). Measurements
have been done in the Voigt configuration with linearly polarized light and in the
Faraday configuration with circularly polarized light (Lehmann et al. 1971, Hlidek
et al. 1976 and Koshizuka et al. 1978a). They show that the lower energy edge in
the magnetically saturated state is strongly polarized and has a triplet structure
consisting of peaks separated by equal amounts for optical transitions with a
change in magnetic quantum number m = +1, 0 and - 1 . This splitting was
attributed by Lehmann et al. (1971) to a splitting of the valence band by the
magnetic field.
Reflectance spectra have been measured in a much wider range of photon
energies. Ahrenkiel et al. (1971) measured them up to 4 eV, and observed at 4 K a
maximum reflectance at 2.0 eV with a weak feature at 2.9 eV. Fujita et al. (1971)
observed at room temperature a maximum at 1.9 eV and changes of slope at 1.4,
1.6 and 1.9eV. Itoh et al. (1973) measured at room temperature a maximum
reflectivity around 2 e V and sub-bands at 1.6 and 2.9eV. The latter authors
observed that the main peak at 2 eV splits into three overlapping peaks at 1.9, 2.0
and 2.1eV, when the temperature is lowered to 80K. Zv~ra et al. (1979)
measured the specular reflectivity at room temperature up to 12eV. In their
measurements the maximum reflectivity occurred at 1.82eV. They observed
additional fine structure at 1.50, 2.6, 3.14, 4.10, 6.30, 7.25 and a triad around 9 eV.
A Kramers-Kronig analysis resulted in a real part of the dielectric constant with a
maximum at 1.38 eV, in good agreement with the optical absorption edge at room
temperature, described above. The imaginary part of the dielectric constant
showed a broad maximum around 2.05 eV and additional fine structure, in
correspondence with that of the reflectivity. Theses results agree well with those
of Itoh et al. (1973).
Thermoreflectance spectra of CdCrzSe4 have been measured by Stoyanov et al.
(1975) and Taniguchi et al. (1975). Although it is difficult to deduce the positions
of the optical transitions from such spectra, the structure around 1.4 and 2.0 eV
can be recognized as belonging to the optical absorption edge and the transition at
about 2.0 eV, observed in the reflectance spectra.
Reflectance magneto-circular dichroism spectra at 4 K show a doublet with
opposite sense of polarization at 1.8 and 2.6eV (Ahrenkiel et al. 1971). Sato
(1977) also measured the reflectance magneto-circular dichroism at 4 K and found
a similar spectrum. A more involved analysis resulted in clear transitions at 1.9,
2.2 and 2.5 eV, a broad structure at 1.3 eV and some weaker transitions around
3 eV. Bongers et al. (1969) published a magneto-optical Kerr effect spectrum,
measured at various temperatures between 4 and 140 K (fig. 64). Strong magneto-
optical transitions were observed at 1.4, 2.0 and 2.6 eV, which values agree well
with those of Sato. The transition at 1.4 eV shows a shift to longer wavelengths at
lower temperatures. The Faraday rotation, measured in the transparent region
686 R.P. VAN STAPELE

2O
u
c~
"B
c 0
E
t-
o

-20
o

~o

-40

I I 2L2 i 1
-60 10 14 18 26 x 103
Weve n u m b e r (cm -1)
Fig. 64. Kerr rotation of CdCr2Se4as a function of the wave number at 140 K, 120 K and 4 K, accordingto
Bongers et al. (1969).

between 1 and 17 Ixm, is composed of a constant part induced by ferromagnetic


resonance transitions and of a large negative part due to an electronic transition
at 2.6 eV (Bongers and Zanmarchi 1968).
Sato and Teranishi (1970) studied the photoconductivity of undoped p-type
single crystals. The spectral dependence (fig. 65) showed a narrow and weak peak,
shifting from 1.3 eV at 200 K t o 1.1 eV at 70 K, and a broad and strong peak
around 2 eV, which shifts slightly to higher energies with decreasing t e m p e r a t u r e
(fig. 62). The results of later investigations of the photoconductivity are similar,
but with m o r e detail. Using In-doped n-type single crystals Amith and Berger
(1971) and Berger and Amith (1971) observed a red-shifting transition at 1.35 eV
at 200 K and at 1.2eV at 1 0 0 K and transitions at 1.4 and 1.SeV, that have
temperature-independent positions. Stoyanov et al. (1976) observed in undoped
p-type single crystals that at 1.4 eV the peak (A) splits at 200 K into two branches,
one shifting to the red and the other to the blue with decreasing t e m p e r a t u r e (figs.
62 and 66). These authors report two broad structures in the spectral dependence
of the photoconductivity, one around 1.7 eV (B) and the other around 2.0 eV (C)
(fig. 66). At the low energy of about 1.1 eV a shoulder was found, which also
showed a shift to the red (fig. 62).
T h e picture that emerges from the optical m e a s u r e m e n t s is still confused. There
are arguments, however, for assigning the red-shifting precursor absorption to a
transition from the valence band to localized vacancy states ( H a r b e k e and
L e h m a n n 1970, L e h m a n n et al. 1971) and the part of the absorption edge above
about 1.3 eV to an indirect transition from the valence band to the conduction
band (Sakai et al. 1976). The m a x i m u m in the reflectivity around 2.0 eV suggests
that the direct transitions from the valence band to the conduction band fall
SULPHOSPINELS 687

1.0 {a}

"~ 0.5
:3

I I I I
~ o
c
~ 1.o (b}

@
a2

0.5

0 J I I I
1.0 1.5 2.0 2.5
Energy {eV}
Fig. 65. Spectral dependence of the photoconductivity of CdCr2Se4 at 300 K (a) and 77 K (b),
according to Sato and Teranishi (1970).

within this energy region (Sato and Teranishi 1970, Ahrenkiel et al. 1971, Sato
1977 and Zvfira et al. 1979). The wavelength dependence of the Faraday rotation
finally suggests that the charge transfer transition from the valence band to the
empty Cr 2+ 3d states has an energy of about 2.6eV (Bongers and Zanmarchi
1968).
At this point the measurements reported by Batlogg et al. (1978) on the
pressure dependence of the lower absorptive part of the absorption edge should
be mentioned. At room temperature they observed that the edge shifted to higher
energies under hydrostatic pressure, and they concluded from the measured value
of the pressure coefficient that the edge could not be due to transitions involving
s-band states, but that p ~ p interband or p ~ localized state could be reconciled
with the observations.
Using 100 Ixm thick single-crystalline samples, Balberg and Maman (1977)
measured accurately the lower absorptive part of the optical absorption edge at
temperatures close to the ferromagnetic Curie temperature. The position of the
absorption edge, defined by the photon energy at which the absorption coefficient
688 R.P. VAN STAPELE

(/1
t-

b
E0)
t_

"(3
"x.
0J
(A
C
o
0
~D
c
o
o
0

r-
12_

~' 1.5 2.0


Energy (eV)
Fig. 66. Spectral dependence of the photoconductivity of CdCr2Se4 at various temperatures, according
to Stoyanov et al. (1976).

equals 200 cm 1, was used in a determination of the critical parameters that was
based on the theory of Alexander et al. (1976) for the critical behaviour of the
direct optical gap of a ferromagnetic semiconductor. N o arguments were given for
this assignment, which deviates from the view that the direct edge occurs at about
2 eV (see above).
Much discussion has been devoted to the origin of the red shift of the
absorption edge. In addition to the studies mentioned in section 5.1 Cfipek (1977)
calculated the red shift of an absorption edge due to transitions to localized
electronic states on lattice imperfections, influenced by an exchange interaction
with the Cr spins. On the other hand, Zvfira et al. (1979) amplify a suggestion made
by G6bel (1976) that the shift of the edge is not so much directly due to magnetic
exchange interactions as to a magnetostrictive change of the u parameter with
temperature. They argue that small changes of the u parameter give rather large
variations in bond angles and distances, which are expected to give rise to an
appreciable shift of the absorption bands. A small change of the u parameter has
indeed been observed by G6bel (1976).
SULPHOSPINELS 689

As a normal spinel, CdCr2Se4 will have four infrared-active phonons and five
Raman-active phonons (White and DeAngelis 1967). The frequencies of the
infrared-active phonons at room temperature, as measured by far infrared ab-
sorption and reflection, are given in table 15. With decreasing temperature the
frequencies continuously increase, which is largely due to the thermal expansion
(Wakamura et al. 1976b). Below the ferromagnetic Curie temperature there is an
additional small anomalous increase in energy (Arai et al. 1971, Br/iesch et al. 1971
and W a k a m u r a et al. 1976b), which arises from an interaction between the
phonons and the ordering Cr spins (Baltensperger and Helman 1968, Balten-
sperger 1970 and W a k a m u r a et al. 1976b). The low-frequency phonon at 75 cm -1
not only shifts in the ferromagnetic state, but also increases in intensity (Wagner
et al. 1971 and Brfiesch et al. 1971). Such effects have also been observed in the
Raman-active modes. The frequencies of the Raman-active phonons observed by
Steigmeier and Harbeke (1970) are listed in table 16. Three of the expected five
Raman-active modes have been assigned by Steigmeier and Harbeke: F was
observed to have the Alg symmetry, C the Eg and D a T2g symmetry. Brfiesch and
D ' A m b r o g i o (1972), who analyzed the phonons of CdCr2Se4 on the basis of a
simple force model, assigned T2g symmetry to the lines A and E. The temperature
dependence of the frequency of the Raman lines is weak (see table 16) and similar

TABLE 15
Frequencies of the four infrared-active phonon modes of CdCr2Se4.

Frequencies (cm 1) at
room temperature Reference

292 278 Riedel and Horwith (1969)


286.6 264.5 188.0 7 4 . 5 Wagneret al. (1971)
288.1 -+0.6 266.2± 0.2 Lee (1971)
289.3 271.2 189.2 7 5 . 5 Wakamuraet al. (1976b)

TABLE 16
Raman lines of CdCr2Se4, quoted from Steigmeier and Harbeke
(1970).

Raman shift (cm-1) Rel. intensity


Line Assignment 300 K 10 K 50 K

A 84-+2 85-+2 3
B 144±2 147-+2 3
C Eg 154± 1 158-+ 1 33
D T2g 169-+2 172 -+2 23
E 226 -+2 2.5
F Alg 237 ± 2 241 -+2 6.5
G 291 -+2 6
H 300 -+2 7.4
690 R.P. VAN STAPELE

to that of the frequency of the infrared-active phonons. More intriguing is the


temperature dependence of the intensity of the Raman lines. Harbeke and
Steigmeier (1968) were the first to notice that the intensity of the line D decreases
with temperature in the same fashion as the nearest-neighbour spin correlation
function, while the intensity of the lines A, B, C and F were found to be
practically temperature independent (Steigmeier and Harbeke 1970). This has led
to a number of proposed mechanisms for the coupling between the phonons and
the Cr spin system (Baltensperger 1970, Brfiesch and d'Ambrogio 1972), cul-
minating in Suzuki and Kamimura's theory (1973) of spin-dependent Raman
scattering. Experimentally, however, the situation turned out to be rather con-
fused. Steigmeier and Harbeke (1970) scattered H e - N e laser light with a photon
energy of 1.96 eV, well above the absorption edge of CdCr2Se4. This can give the
Raman scattering a resonant character and resonance effects have indeed been
observed by Koshizuka et al. (1977a, b), who found the intensity of line C to
depend only weakly on the temperature and the incident photon energy, whereas
the shape of the temperature dependence of the intensity of the lines D and F
relative to that of line C depends closely on the photon energy between 1.8 eV
and 2.5 eV. In the ferromagnetic state, at 35 K, the intensity of line D was found
to have a pronounced maximum around 2.0 eV, while that of line F had a broad,
less pronounced maximum around 2.2 eV. These maxima fall in the region where
the reflectivity has a maximum, which has been assigned to electronic transitions
near the direct band gap. In the paramagnetic state the main dependence on the
incident photon energy has disappeared. However, the results of the same
measurement done by Iliev et al. (1978a), disagree completely with these findings.
These authors observed that the intensity of line C had a pronounced resonance
peak around 2.0 eV both in the paramagnetic and in the ferromagnetic state, the
peak being slightly broadened in the ferromagnetic state. The intensity of line D
also had a maximum around 2.0 eV, although the dependence on the incident
photon energy was much weaker.
At the much longer incident wavelength of 1.065 ~m the intensity of the Raman
line C has been measured and analyzed by Shepherd (1970). The YAG : Nd 3+ laser
had been chosen by Shepherd because its energy at 200 K is just below the low
absorptive part of the absorption edge of CdCr2Se4, which shifts to the red if the
temperature is decreased below 200 K. He explained the measured temperature
dependence of the intensity of the Stokes l'ine as a combination of three effects:
the change in absorption, the usual temperature dependence of the Stokes
intensity, and the resonant term in the Raman cross section based on transitions
between a parabolic valence and conduction band. The last assumption does not
agree with the conclusion that the lower energy edge is not due to an intrinsic
excitation (Harbeke and Lehmann 1970).
The nuclear magnetic resonance spectra of CdCr2Se4 have been studied by
Berger et al. (1968, 1969a), Stauss et al. (1968) and Stauss (1969a, b). The
spectrum of 53Cr in CdCr2Se4 is very similar to that of 53Cr in CdCr2S4 and has
been analyzed and discussed as described in section 5.3. At 4.2 K the isotropic
hyperfine field Hiso is -182.5 kOe, the axially symmetric hyperfine field Hanis is
SULPHOSPINELS 691

+2.30 kOe, and the strength of the quadrupole interaction Vo is 0.90 MHz (Berger
et al. 1968, Stauss et al. 1968).
The hyperfine field on 77Se is large and negative. At 4 . 2 K the isotropic
component is -98.0 kOe and the axially symmetric component is +9.2 kOe (Stauss
et al. 1968, Berger et al. 1969a).
On lnCd and i13Cd the hyperfine field is large and positive, +136.2 kOe (Berger
et al. 1969a, Stauss 1969a, b). This is less than in CdCr2S4 and corresponds to 1.7%
of an electron spin in the Cd 5s state (Berger et al. 1969a, Stauss 1969b).

5.7. HgCr2se4

HgCr2Se4 is a normal spinel (Baltzer et al. 1966) with lattice parameters as given
in table 1. The lattice parameter has been measured as a function of temperature
by Wakamura et al. (1976b). They found that it decreased steadily with decreasing
temperature between 300 K and 90 K, deviating only Weakly from the Grfineisen-
Debye behaviour below 150 K. HgCr2Se4 is a ferromagnet with a Curie tem-
perature at 106 K, an asymptotic Curie temperature of 200 K and a molar Curie
constant of 3.79, close to the spin-only value of Ca~+ ions (3.75) (Baltzer et al.
1966). The magnetic moment at 4.2 K is 5.64 #B/molecule in an applied magnetic
field of 10 kOe (Baltzer et al. 1966), while Minematsu et al. (1971) report a value of
5.8 _+0.2 ~B/molecule for the saturation moment. The last value agrees well with the
6 p.B/molecule expected for Cr 3+ ions.
The Curie temperature is increased by doping with Cu (Lotgering 1968b,
Okofiska-Kozlowska et al. 1977) and with Ag (Miyatani et al. 1970, Minematsu et
al. 1971). Indium was found to substitute for Cr, and the Curie temperature was
observed to decrease with increasing In concentration (Takahashi et al. 1971,
Miyatani et al. 1970, Minematsu et al. 1971).
From the values of the ferromagnetic and the asymptonic Curie temperature
Baltzer et al. (1966) estimated the strength of the nearest-neighbour exchange
interaction Y and the distant-neighbour interaction K (see section 5.1), which
results in J/k = 15.8 K and K/k = -0.51 K.
Under hydrostatic pressure the ferromagnetic Curie temperature decreases.
Srivastava (1969) measured dTddP = - 0 . 9 5 K/kbar, which, in combination with
the compressibility data, gives an increase of Tc with the lattice parameter at the
rate dTdda = +99 K/A.
HgCrzSe4 is a semiconductor (Baltzer et. al. 1965). Single crystals were observed
to have a p-type conduction with a resistivity that has a maximum at 82K. At
approximately the same temperature, a transition from p-type to n-type conduc-
tion was observed in the Hall coefficient (Lehmann and Emmenegger 1969). Ag
doping (Miyatani et al. 1970, Minematsu et al. 1971) and Cu doping (Lotgering
1968b) results in a much higher p-type conduction. Ag-doped hot-pressed poly-
crystalline samples showed a small positive magnetoresistance at temperatures
near the Curie temperature. The Hall and thermoelectric effects indicate that
p-type carries dominate the electrical conduction of Ag-doped HgCrzSe4 (Mine-
matsu et al. 1971).
692 R.P. VAN STAPELE

In doping increases the resistivity of hot-pressed polycrystalline samples. The


resistivity reached a maximum around the Curie temperature and the samples
have a strong, negative magnetoresistance in that region of temperatures. The
Hall and thermoelectric effects show a complicated behaviour, indicating that
more than one type of carrier takes part in the conduction (Miyatani et al. 1970,
Minematsu et al. 1971). Similar properties have been observed by Takahashi et al.
(1971) in the mixed crystals HgCr2-xInxSe4, in which In can be substituted for Cr
up to x = 0.45. These authors conclude that In also substitutes for Cr in the
In-doped polycrystalline samples studied by Minematsu et al. (1971). This means
that impurities due to deviations from stoichiometry play a role in the conduction
mechanism.
Recently, Goldstein et al. (1978) and Selmi et al. (1980) found that the electrical
transport properties of HgCrzSe4 crystals can be changed appreciably by annealing
in a Hg or a Se atmosphere. Annealing in a Hg atmosphere results in high-
mobility n-type samples, whereas annealing in a Se atmosphere results in materi-
als with p-type conduction at room temperature and n-type conduction at low
temperatures.
Although the analysis of the electrical transport phenomena is far from com-
plete, the observed temperature dependence of the resistivity and the mag-
netoresistance reveals the active presence of an interaction between the charge
carriers and the Cr spin system. A peculiar effect of this interaction has been
observed by Toda (1970), who measured a decrease of the electrical resistivity and
an induced DC voltage in the sample at ferromagnetic resonance.
At wavelengths shorter than 2.5 ~m the optical absorption spectrum of
HgCrzSe4 consists of a broad, weak absorption at 0.6 eV and of an absorption
edge located at 0.84eV at room temperature, shifting strongly to longer
wavelengths with decreasing temperature (Lehmann and Emmenegger 1969). The
position of the absorption edge, which is depicted in fig. 67, starts to move to the
red well above the Curie temperature. It shows a further shift to the red in the
ferromagnetic state of the crystal (Lee et al. 1971, Arai et al. 1973). Application of
a magnetic field shifts the edge further to the red (Lehmann and Emmenegger
1969, Arai et al. 1973), which indicates that the shift is related to the magnetic
ordering. However, simple mechanisms like exchange split bands or mag-
netoelastic coupling fail to explain the observed temperature dependence of the
position of the absorption edge (Arai et al. 1973).
Doping of HgCrzSe4 single crystals results in a shift of the absorption edge. In
In-doped crystals the edge is slightly shifted to shorter wavelengths, whereas in
Ag-doped crystals the edge is shifted to the red (Miyatani et al. 1970). The
Ag-doped crystals also show an additional broad absorption band at 0.62eV,
which becomes sharper at lower temperatures (Miyatani et al. 1970).
The frequencies of the observed optically active phonons are listed in tables 17
and 18. The frequencies of two of the four infrared-active phonons were derived
from the reflectance at room temperature (Lee et al. 1971). In absorption, the
frequencies were measured as a function of temperature (Wakamura et al. 1971,
1976b). Between 300K and 85 K, the frequencies increase continuously with
decreasing temperature, showing a small additional shift to higher frequencies
SULPHOSPINELS 693

(]

0,8

0.6
Z
O..
(D I
/
0.~ /b
/
!
/
,J

0.2

Tc
&

I1~ I I I I I
00 100 200 300 400 500 600
T(K)

Fig. 67. Shift of the optical absorption edge of HgCr2Se4. (a) The position at an absorption coefficient
of 2000 cm -], according to Lehmann and Emmenegger (1969) and (b) the position at an absorption
coefficient of 240 cm 1, according to Arai et al. (1973).

below the ferromagnetic Curie temperature. As in CdCr2Se4, this anomalous shift


can be explained phenomenologically by assuming that the atomic potential
depends on the magnetization (Baltensperger 1970, Wakamura et al. 1976b)*.
The frequency and the intensity of the five Raman-active phonons have been
measured as a function of temperature between 8 K and room temperature and as
a function of the incident photon energy in the range 1.5 to 3 eV (Iliev et al.
1978b). The frequencies are listed in table 18. The assignments made by Iliev et al.
T A B L E 17
Frequencies of the four infrared-active phonons of HgCr2Se4.

Frequencies (cm -]) at


room temperature Reference

286.8 268.6 Lee et al. (1971)


287.7 276.6 170.9 58.7 Wakamura et al. (1976b)

* See notes added in proof (h) on p. 737.


694 R.P. VAN STAPELE

TABLE 18
Raman lines of HgCr2Se4, quoted from Iliev et al.
(1978b).

Raman shift (cm 1)


Line Assignment 300 K 10 K

A T2g 6O.1 66.2


B 140.5
C Eg 152.9 158.8
D T2g 163.4 168.7
E T2g 207 211
F Alg 235.9 238

(1978b) are based on the analysis of Briiesch and D'Ambrosio (1972) of the lattice
vibrations in the spinel structure, since selection rules are not always obeyed in
the (resonant) Raman scattering of compounds like HgCr2Se4. The phonon
frequencies in HgCr2Se4 and in CdCrzSe4 (table 16) have nearly the same value,
except for line A, which means that the diamagnetic cations take part in the
vibration only in mode A. The temperature dependence of the Raman intensities
shows no anomaly at the ferromagnetic Curie temperature, but it depends
significantly on the incident photon energy. This suggests, as Iliev et al. (1978b)
conclude, that the observed changes in Raman intensity are due to temperature-
induced changes of the resonance conditions.
We conclude this review of the properties of HgCr2Se4 with the nuclear
magnetic resonance data published by Berger et al. (1969a). At 1.4K, these
authors measured the hyperfine fields at the nuclear spins of 53Cr, 77Se, 199Hg and
2°~Hg. The isotropic part of the hyperfine field on 53Cr amounts to - 179.4 kOe and
in comparison with the isotropic hyperfine fields in other ferromagnetic semicon-
ductors can be correlated with the strength of the exchange interactions between
the Cr spins (Berger et al. 1969a). The negative isotropic part of the hyperfine
field on the Se nuclear spin (-91.7 kOe) could be understood in terms of a spin
polarization of Se s orbitals by the Cr spins. The large and positive hyperfine field
on the Hg nuclear spins (+446 kOe) was found to be isotropic and was ascribed to
an unpaired spin density in the empty Hg 6s shell of 1.7% of an electron spin.
This compares well with the data on CdCrzSe4 (Berger et al. 1969a).

5.8. Mixed crystals between the compounds ZnCrzX4, CdCr2X4


and HgCrzX4 with X = S, Se

In this section we will briefly review the properties of mixed crystals between
ZnCr2X4, CdCrzX4 and HgCrzX4.
In the sulphides the only series that has been investigated is Cdl-xHgxCr2S4,
between the ferromagnet CdCr2S4 and the metamagnet HgCr2S4, which have
nearly the same cell edge (table 1). Baltzer et al. (1967) observed that the
asymptotic Curie temperature 0 changes gradually from 0 = 152 K for CdCr2S4 to
SULPHOSPINELS 695

0 = 142 K for HgCr2S4, passing through a gentle maximum. The magnetic ordering
temperature varies monotonically from the Curie temperature (84.5 K) of CdCrES4
to the N6el temperature (36 K) of HgCr2S4. The metamagnetic behaviour appears
at x = 0.65. Within the limits of the model given by Baltzer et al. (1966), this
variation of 0 and Tc could be described by a positive effective nearest-neighbour
exchange J, depending only weakly on x, and a negative effective unified more-
distant exchange /(, that drops rapidly in magnitude when the composition
approaches that of CdCr2S4.
In the selenides most attention has gone to the series Znl_xCdxCr2Se4, between
the antiferromagnet ZnCr2Se4 and the ferromagnet CdCr2Se4. The latter com-
pound has a 3 percent larger cell edge than ZnCr2Se4 (table 1). The cell edge of
Znl_xCdxCr2Se4 has been reported to vary linearly with x (Busch et al. 1969,
Wakamura et al. 1976a). The asymptotic Curie temperature increases linearly
(Baltzer et al. 1967, Lotgering 1968b) or nearly linearly (Busch et al. 1969) with x
(fig. 68). The transition from antiferromagnetism to ferromagnetism occurs at
x = 0.4 according to Baltzer et al. (1967) and Lotgering (1968b) or, as measured by
Busch et al. (1969) in very low magnetic fields, between x -- 0.5 and 0.6 (fig. 68).
Although the large number of exchange parameters involved prevents a reliable
analysis of the magnetic ordering in terms of x-dependent exchange interactions

200

150 /°

100 I

Q.
E
t---

50

- - ~TN× ~
× "'"x

I012101 1 i I ' 1
00. .L 06 0.8 1.0
X

Fig. 68. Asymptotic Curie temperature, N6el temperature and ferromagnetic Curie temperature of
Znl-xCdxCr2Se4 as a function of the composition: (Q) data of Baltzer et al. (1967), (O) data of
Lotgering (1968b) and (x) data of Busch et al. (1969).
696 R.P. V A N S T A P E L E

(Lotgering 1968b), some authors conclude that the nearest-neighbour exchange


interaction is essentially constant, but that the more distant interactions vary
drastically with the replacement of Zn by Cd ions (Baltzer et al. 1967, Makhotkin
et al. 1978a). Lotgering (1968b), on the other hand, mentioned the possibility that
the main influence of this substitution is exerted on the nearest-neighbour
exchange interaction. In CdCrzSe4, the ferromagnetic Curie temperature
decreases under hydrostatic pressure. The rate of change, -0.76 x 10-3K/bar,
increases with decreasing Cd concentration to - 1 . 3 x 10-3K/bar for
Zn0.4Cd0.6Cr2Se4 (Fujii et al. 1970). Finally we mention that the small x-dependent
shift of the frequencies of the four infrared-active phonons in Zn>xCdxCr2Se4 has
been measured and analyzed (Wakamura et al. 1976a), and that the magnetic
permeability of crystals with x/>0.76 has been observed to be sensitive to
illumination (Makhotkin et al. 1975).
In the system Cdl_xHgxCr2Se4, between the ferromagnets CdCr2Se4 and
HgCr2Se4, the ferromagnetic Curie temperature decreases linearly from 130 K to
106K (Vinogradova et al. 1978). These authors observed that the dynamic
permeability was sensitive to illumination throughout the series of mixed crystals.
The long-wavelength edge of this photomagnetic effect and of the photoconduc-
tivity was observed to shift gradually to longer wavelengths with increasing Hg
concentrations.
In the system Zn~ xHgxCr2Se4, between the antiferromagnet ZnCr2Se4 and the
ferromagnet HgCr2Se4, the cell edge varies linearly (Wakamura et al. 1973).
Antiferromagnets have been observed for 0 ~<x ~<0.4, whereas at Hg concen-
trations higher than 0.5 a ferromagnetic behaviour was observed (Wakaki et al.
1975). The values of the asymptotic Curie temperature, the N6el temperature and
the ferromagnetic Curie temperature are given in fig. 69. In the ferromagnetic
compositions, the magnetic ordering has been analyzed in terms of a nearest-
neighbour exchange interaction and a unified more distant interaction, which both
change with the composition (Wakaki et al. 1975). The optical absorption edge
gradually shifts to lower energies with increasing Hg concentration, the magnitude
of the anomalous red shift increasing from the value in ZnCr2Se4 to that in
HgCr2Se4 (Wakaki and Arai 1978). Finally, the shift of the frequencies of the
infrared-active phonons has been measured and analyzed by Wakamura et al.
(1973)*.
In the solid solution ZnCr2(S2-xSex)4 the cell edge increases linearly with x,
whereas the chalcogen parameter remains constant (Riedel and Horvfith 1969).
These authors have published some data on the infrared spectra, while the
preparation of single crystals has been reported by Pickardt and Riedel (1971).
Substitution of Te for Se turned out to be only partly possible. Single-phase
samples of ZnCr2(Sei-xTex)4 with x >0.2 could not be obtained (Riedel and
Horvfith 1969).
The lattice parameter of CdCr2(Sl_xSex)4 increases linearly with x, whereas the
u parameter remains constant (Wojtovicz et al. 1967, Riedel and Horvfith 1969). The
asymptotic Curie temperature 0 increases linearly with x, but the ferromagnetic

* See notes added in proof (i) on p. 737.


SULPHOSPINELS 697

200

(3

150

2
0)
fl
E
(9
F-

100

50

TN

1 i i II I OIi~ I I I
O0 012 0 Z,
I'0 0"8
X
Fig. 69. Asymptotic Curie temperature 0, N6el temperature TN and ferromagnetic Curie temperature
T¢ of Znl-xHgxCr2Se4, according to Wakaki et al. (1975).

200

150
D

g
b--

100

I
0 I 0'.2 ' 0'.4 01.6 ' 018 ' 1.0
X
Fig. 70. Asymptotic Curie temperature 0 and ferromagnetic Curie temperature Te of CdCr2(Sl-xSex)4,
according to Wojtowicz et al. (1967).
698 R.P. VAN STAPELE

Curie temperature stays behind 0 for x ~<~ (fig. 70) (Wojtowicz et al. 1967). This
finding has been analyzed by Wojtowicz et al. (1967) in terms of an effective
nearest-neighbour exchange interaction J and an effective unified more distant
interaction/(. Both parameters were found to vary in a non-linear way, the positive 37
passing through a weak maximum and the negative /£ having a pronounced
minimum. The preparation of single crystals has been reported by Pickardt et al.
(1970), data on the infrared spectra have been given by Riedel and Horvgtth (1969),
and data on the position and the structure of the optical absorption edge have been
published by Kun'kova et al. (1976).

5.9. Mixed crystals A1/2A1/2Cr2X4


1+ 3+ with X = S, Se and diamagnetic ions A

An interesting class of compounds is formed by the spinels AlaA1/2Cr2X4,


1+ 3+ in which
the tetrahedral sites are occupied by equal amounts of monovalent diamagnetic
cations like Li +, Cu + or Ag ÷ and trivalent diamagneti c ions like AP +, Ga 3÷ or In 3+.
These compounds are expected to be semiconductors, which has been ascertained
experimentally in the case of CUl/2Inl/zCr2Se4 (Yokoyama and Chiba 1969) and
Cumlnl/eCr2S4 (G6bel .et al. 1974). In a number of these compounds, and
specifically the compounds that contain Li + or Cu ÷, the monovalent and trivalent
cations order on the tetrahedral sites (table 19). This ordering will be discussed
first.
The spinel A sites form two equivalent fcc Bravais lattices (section 2, fig. 2). Since
the sites of one Bravais lattice are surrounded tetrahedrally by four sites of the
other, the A sites are suitable for ionic 1:1 ordering according to the two
sublattices. The driving force is the electrostatic energy arising from a charge
difference, while a difference in ionic radius may also play a role. The ordering
was first observed by Lotgering et al. (1969) in CumFel/2Cr2S4 (section 6.6),
Cul/2Inl/2CraS4 and CUl/2Inl/zCr2Se4. It has also been found in Cul/2Fel/2RhzS4
(section 7.4) and in many of the compounds listed in table 19. The 1 : 1 ordering on
the tetrahedral sites lowers the space group from Fd3m (O 7) of spinel to F43m
(T~). The inversion centres in the spinel lattice (being the B sites) have disap-
peared in T 2. The expected occurrence of piezoelectricity could not be detected
(Pinch et al. 1970)*. Another consequence of the ionic ordering is that the X
tetrahedra around the A + ions and around the A 3+ ions, which have no common X
ions (fig. 2), are not equivalent, so that the A+-X and A3+-X distances are two
parameters that determine the positions of the X lattice completely. Two X
parameters have indeed been observed by means of neutron diffraction in
CUl/2Inl/zCrzX4 for X = S and Se (Plumier, private communication). The
differences between the Cu+-X and In3+-X distances is about 20% in the sulphide
and about 5% in the selenide.
Table 19 gives a survey of the properties of A1/zA1/z[Cr2
+ 3+ 3+]X4 with diamagnetic A
ions. Ionic ordering does not always occur and the data suggest that ordering
occurs for Li + and Cu ÷ and not for Ag + (Pinch et al. 1970). The absence of
ordering in Cul/zGal/z[Cr3+]S4(table 19) and in Cul/zInl/a[B2]84 with B = In or Rh

* See notes added in proof (j) on p. 737.


SULPHOSPINELS 699

TABLE 19
Crystallographic and magnetic data of mixed spinels A1/2A1/2Cr2X4.
1+ 3+ u = chalcogen parameter, Cm =
molar Curie constant, 0 = asymptotic Curie temperature, TN = Nrel temperature and Tc = ferro-
magnetic Curie temperature. References: (a) Yokoyama and Chiba (1969), (b) Lotgering et al. 1969, (c)
Pinch et al. (1970), (d) Locher and Van Stapele (1970), (e) Plumier et al. (1971b), (f) Plumier and Sougi
(1971), (g) Wilkinson et al. (1976), (h) Plumier et al. (1977a), (i) Plumier et al. (1977b).

Ionic Critical
Celledge ordering 0 temperature
Compound (A) u on A sites Cm (K) (K) Ref.

LiI/2Gal/2Cr2S4 9.974 0.385 yes TN -- 14 (c)


Lil/2InmCr2S4 10.127 0.385 yes Ty = 27 (c)
Cul/2A]I/zCr2S4 9.915 0.382 yes TN = 14 (c)
Cul/2Gal/2Cr2S4 9.920 0.381 ? TN = 31 (c)
9.918 0.382 no (g)
Cul/2Inl/zCr2S4 yes (b)
10.065 0.388 yes TN -- 26 (c)
10.060 yes 3.83 -77 TN = 40 (e)
TN = 160.9 (i)
AgmAll/2Cr2Sa 10.067 0,385 no TN = 7 (c)
AgmGal/2Cr2S~ 10.063 0,387 no TN- 10 (c)
Ag~/2Inl/2Cr2Sa 10.215 0,390 ? TN- 14 (c)
10.24 3.43 +142 Ty = 17 (e)
0.3876 no (f)
TN = 138.9 (h)
Cul/2AlmCr2Se4 10.438 0.381 yes Ty = 6 (c)
Cul/2Gal/2Cr2Se4 10.444 0.385 ? TN -- 7 (c)
Cu~/2InmCr2Se~ 10.583 3.8-+ 0.2 + 135 _+5 (a)
yes (b)
10.580 0.386 yes TN- 14 (c)
3.4--+0.1 +105-+5 (d)
10.58 3.58 + 100 no (e)
at 4.2 no
ordering (g)
Agl/2Inl/2Cr2Se4 10.724 0.390 no Tc = 50 (c)
10.72 3.86 + 180 T0 - 60 (e)

( L o t g e r i n g , p r i v a t e c o m m u n i c a t i o n s ) s h o w s t h a t t h e p r o b l e m is less s i m p l e . O n t h e
o t h e r h a n d 1 : 1 o r d e r i n g o n t h e t e t r a h e d r a l sites s e e m s t o b e s u r p r i s i n g l y s t a b l e , as
a p p e a r s f r o m t h e o c c u r r e n c e in F e x C u l - x R h 2 S 4 w i t h 0.46 ~ x < 0.7 ( B o u m f o r d a n d
M o r r i s h 1978) a n d in Inz/3D1/3[In2/3Cr4/3]S4
3+ 3+ 3+ 3+
a n d Inl/zDm[Cr3/2Sn 34- 44-
m]S4 ( L o t g e r i n g
a n d V a n d e r S t e e n 1971b), n o t w i t h s t a n d i n g a s t r o n g d e v i a t i o n f r o m t h e i d e a l
composition.
I n t h e i r p a r a m a g n e t i c state, t h e c o m p o u n d s t h a t h a v e b e e n i n v e s t i g a t e d h a v e a
C u r i e c o n s t a n t t h a t m o r e o r less a g r e e s w i t h t r i v a l e n t C r i o n s (Cm = 3.75) ( t a b l e
19). A l t h o u g h a p o s i t i v e a s y m p t o t i c C u r i e t e m p e r a t u r e has b e e n f o u n d in m o r e
cases, o n l y o n e of t h e c o m p o u n d s l i s t e d in t a b l e 19 o r d e r s f e r r o m a g n e t i c a l l y . T h i s
c o m p o u n d , A g m I n m C r 2 S e 4 , h a s a f e r r o m a g n e t i c C u r i e t e m p e r a t u r e o f a b o u t 60 K.
T h e m a g n e t i c m o m e n t has b e e n r e p o r t e d t o b e 4.7/xB at 4.2 K a n d 10 k O e ( P i n c h
et al. 1970) a n d 5 . 1 / z ~ at 4.5 K a n d 30 k O e ( P l u m i e r et al. 1971b), b o t h v a l u e s
b e i n g l o w e r t h a n t h e 6 tzB e x p e c t e d f o r C r 3+ ions. H o w e v e r , e v e n in 30 k O e t h e
700 R.P. V A N S T A P E L E

material was not saturated (Plumier et al. 1971b). Antiferromagnetic ordering has
been found in all other compounds in table 19, with the exception of
Cul/2Inl/2Cr2Se4, which shows no ordering at 4.2 K (Plumier et al. 1971b, Wilkinson
et al. 1976).
The spin configuration in AgmInmCr2S4 (Plumier and Sougi 1971) and
CUl/2Gal/zCrzS4 (Wilkinson et al. 1976) is not commensurate with the crystallo-
graphic unit cell. In the former case a spiral arrangement like that in ZnCrzSe4
(section 5.5) was found, in the latter case a spiral arrangement with a propagation
vector that does not point in a main crystallographic direction.
The antiferromagnetic ordering found in Cu~/zInmCr2S4 (Plumier et al. 1971b)
consists of four magnetic sublattices with magnetizations pointing along the four
cube diagonals. This structure is commensurate with the unit cell and the four B
sites in one octant (fig. 2) belong to different magnetic sublattices. It can easily be
shown that the Heisenberg exchange energy is degenerate with respect to the
mutual orientation of the sublattice magnetizations if the resulting magnetization
vanishes. Other kinds of interactions must occur in order to remove the
degeneracy. An isotropic biquadratic interaction J(Si • ~)2 with a positive J gives
indeed a minimum energy for the configuration observed. This is a strong
indication for the occurrence of biquadratic exchange in sulphospinels.
The magnetic and electrical properties of Cu0.5+xln0.5-xCrzS4 with -0.1 ~<x ~<0.1
depend strongly on x (G6bel et al. 1974). The resistivity has a sharp maximum at
x = 0, at which composition the conduction changes from p-type (for positive x) to
n-type (for negative x). The paramagnetic Curie temperature shows a pronounced
minimum at x = 0. A transition from ferromagnetism (established by measurements
of the magnetic hysteresis) to antiferromagnetism occurs in the p-type region
near x = 0. These results show Cul/21nl/2CrzS4 to be a semiconducting antiferromag-
net and illustrate again the close connection between the strong ferromagnetic in-
teraction in C u C r 2 S 4 and the hole conduction in the valence band (see section 3).
Recent measurements of the magnetic susceptibility and the specific heat,
together with neutron diffraction experiments, revealed that in Ag~/2Ina/2Cr2S4 a
first-order transition from "helimagnetic macrodomains" to "metamagnetic
microdomains" takes place at 12 K, a transformation to smaller microdomains at
42.5 K, and that the transformation to the paramagnetic state takes place at 138 K
(Plumier et al. 1977a). Similar phenomena have been observed in ZnCr2Se4 (see
section 5.5) and in Cul/2Inl/2Cr2S4. In the last compound the first-order trans-
formation from long range magnetic order to a short range ordering takes place at
31 K, to yet another short range ordered state at 35 K and to the paramagnetic
state at 158 K (Plumier et al. 1977b)*.
Nuclear magnetic resonance of 63Cu, 65Cu and 115In in the compounds
C u l / z l n l / z C r 2 S 4 and CumlnmCr2Se4 revealed the existence of large supertransferred
hyperfine fields on nuclei of the diamagnetic cations. In terms of spin densities in
the first empty s shell of the diamagnetic ion, these fields compare well with those
measured in CdCrzX4 and HgCrzX4 (Locher and Van Stapele 1970).

* See notes added in proof (k) on p. 737.


SULPHOSPINELS 701

6. Ferrimagnetic semiconductors

6.1. Introduction

In this section we will discuss the properties of the normal sulphospinels MCr2S4
with M = Mn, Fe and Co and of some mixed crystal series connected with these
compounds.
The compounds MCrzS4 are interesting because they provide examples of
ferrimagnetism in semiconducting non-oxidic compounds (Lotgering 1956). Their
properties can be compared with the properties of the corresponding oxyspinels in
order to study the influence of the anions. From this point of view it would have
been interesting to compare oxyspinels MFe204 with sulphospinels MFezS4.
However, the fact that many iron sulphides are not stable with respect to FeS2
prohibits the preparation of most of the sulphur analogues of the ferrites MFe204,
which are important materials from a technical point of view. Nevertheless, the
thorough investigation of the physical properties of the sulphochromites, which
was started by Lotgering (1956), has yielded many interesting phenomena, a
better knowledge of superexchange interactions and some insight into the relative
stability of the valencies of cations and anions.
Among the results, worth mentioning in advance is the magnetization versus
temperature curve of MnCrzS4, which provided the first indication of the existence
of a fairly positive exchange interaction between Crs+ ions (Lotgering 1956)
(section 6.2). Another example is the occurrence of a semiconducting ferrimagnet
FemCua/2Cr2S4 in the series of mixed crystals between the semiconducting fer-
rimagnet FeCrzS4 and the metallic ferromagnet CuCrzS4 (section 6.6), which
implies Fe 3+ and Cu 1+ ions. It was found that these ions are ionically ordered on
t h e tetrahedral sites (Lotgering et al. 1969). The same ordering has also been
observed in a number of compounds with trivalent and monovalent diamagnetic
ions on the tetrahedral sites (section 5.9) and later on in the fascinating antifer-
romagnet Fel/zCua/zRhzS4 (section 7.4), in which strongly° negative exchange inter-
actions occur between Fe 3+ ions at a distance of 9.85 A (Plumier and Lotgering
1970).
Finally, the compound FeCrzS4 itself is interesting because of its strong mag-
netic anisotropy at low temperatures and the cooperative Jahn-Teller effect below
9 K (Van Stapele et al. 1971, Spender and Morrish 1972b, Van Diepen and Van
Stapele 1973). Both effects are connected with the 5E ground state of the Fe 2+ ions
(section 6.3).
The magnetic data of the compounds MnCrzS4, FeCr2S4 and CoCrzS4 are
collected in table 20.

6.2. MnCr2S4

MnCr2S4 is a normal (Menyuk et al. 1965) cubic spinel (Passerini and Baccaredda
1931) with lattice parameters as listed in table 1. At high pressures and tern-
702 R.P. VAN STAPELE

0
L)

e~

'6
~SNg

~ N t

.e

p.. ,~

e~
E
¢)

~

+1 o +1 e-- +l

~'.~ I ~
I
+ I cq
I

~8

<
SULPHOSPINELS 703

~t.,~ ~ ~, ~ " ~ , ~ , ...~, .~. ~,.~ 1"'-, t"~

t"q

+1

t~
I i J i

C)
704 R.P. V A N S T A P E L E

2.5

-5 2.0
0

Ig
1.5 S t \

J
1.0
i i 1 r I i
0 '1'0 3'0 5'0 6'o 7'o
temperature (K) .~
Fig. 7]. Magnetization of MnCr2S4 as a [unction of temperature: (a) in 8.4 kOe (Lotgerin g 1956), (b) in
10 k O e (Menyuk et al. 1965) and (c) extrapolated to H = 0 (Lotgering 1968a).

peratures a transformation to a cation-defective NiAs structure has been observed


(Bouchard 1967, Tressler and Stubican 1968, Tressler et al. 1968).
According to the theory of Yafet and Kittel (1952) for ferrimagnetism in
spinels, a canted spin configuration cannot transform at higher temperatures into
the paramagnetic state directly, but has to pass through the collinear state.
Further, the occurrence of a m a x i m u m in the magnetization versus t e m p e r a t u r e
curve, as predicted by N6el (1948), cannot occur in the canted state but only in the
collinear state. These theoretical conditions are nicely demonstrated in MnCr2S4.
The c o m p o u n d is a ferrimagnet with a Curie t e m p e r a t u r e of about 8 0 K
(Lotgering 1956, 1968a) or 66 K (Menyuk et al. 1965, Darcy et al. 1968). In the
paramagnetic state the magnetic susceptibility follows at higher temperatures a
Curie-Weiss law with a Curie constant which is in rough agreement with the
spin-only value of Mn a÷ and C r 3÷, and an asymptotic Curie t e m p e r a t u r e of
- 1 2 _+ 1 0 K (Lotgering 1956) or + 1 0 K (Darcy et al. 1968). In the ferrimagnetic
state the magnetization has a m a x i m u m at about 40 K (fig. 71) (Lotgering 1956),
decreasing to 1.3/xB/molecule at 4.2 K (Menyuk et al. 1965, Darcy et al. 1968).
F r o m the low value of the asymptotic Curie t e m p e r a t u r e and the occurrence of a
m a x i m u m in the magnetization versus t e m p e r a t u r e curve, Lotgering (1956)
concluded that a fairly large positive exchange interaction IBB exists between the
Cr 3÷ ions on the octahedral sites and a weaker negative exchange interaction IAB
between the octahedral Cr 3+ ions and the Mn 2÷ ions on the tetrahedral sites.
Subsequent m e a s u r e m e n t s of the magnetization down to 2 . 2 K in fields up to
30 k O e revealed the existence of a Yafet-Kittel spin configuration with canted "
Mn 2÷ spins on the tetrahedral sites below 5 K and a collinear N6el configuration
SULPHOSPINELS 705

above 5 K (fig. 71) (Lotgering 1968a). The canting of the spins on the tetrahedral
sites indicates a relatively strong exchange interaction IAA between the Mn 2+
spins. From the value of the magnetic moment extrapolated to zero magnetic field
and zero temperature (1.18/zB), the temperature-independent differential suscep-
tibility below 4.5 K and the value of the asymptotic Curie temperature Lotgering
(1968a) found IAB/k = --1.79 K, IAA/k = --1.68 K and IBB/k +10 K. A similar/An,
=

(--0.80 K) was observed in the antiferromagnet MnSc2S4 (Wojtowicz et al. 1969).


The canted spin configuration has been observed by means of neutron diffraction
(Plumier and Sougi 1969). The transition to the collinear N6el configuration takes
place suddenly at 5.5 K. The temperature dependence of the sublattice mag-
netizations measured between 1.5 K and 45 K agrees well with that of the total
magnetization as measured by Lotgering (1968a) and Menyuk et al. (1965). A
molecular field calculation of the sublattice magnetization as a function of
temperature agreed with the experimental results only when exchange striction
effects were taken into account (Nauciel-Bloch et al. 1972).
In high magnetic fields above 100 kOe (fig. 72) yet another magnetic phase
exists at temperatures below 30 K (Denis et al. 1969, 1970). The spin structure in
this phase and the magnetization in high fields have been the subject of molecular
field calculations in which magnetostrictive effects play an important part (Plumier
1970a, b, 1980, Plumier et al. 1971a) (see also Plumier 1980).
Other physical properties of MnCr2S4 have barely been studied. The compound
is reported to be a semiconductor (Bouchard et al. 1965, Albers et al. 1965) and
infrared-active phonons at 381, 323, 260 and 120 cm -1 have been observed in the
infrared absorption spectrum in the course of a general study of the lattice
vibrations of chalcogenide spinels (Lutz and Feh6r 1971).

.,//7ji" 25 K
g3

E
1

I
100 200 300
H {kOe)
Fig. 72. Magnetizationof MnCr2S4as a functionof magneticfield, accordingto Denis et al. (1970).
706 R.P. VAN STAPELE

6.3. FeCr2S4

FeCrzS4 is known as the mineral daubreelite, which Lundqvist (1943) demon-


strated to have the cubic spinel structure. Neutron diffraction studies (Broquetas-
Colominas et al. 1964, Shirane et al. 1964) showed that it is a normal spinel. The
lattice parameters are given in table 1. At high pressures and/or temperatures
FeCr2S4 adopts a cation-defective NiAs structure (Albers and Rooymans 1965,
Bouchard 1967, Tressler and Stubican 1968, Tressler et al. 1968). At low tem-
peratures down to 4 K, FeCr2S4 remains a cubic spinel (Shirane et al. 1964, G6bel
1976), but an anomalous broadening of the X-ray diffraction lines indicates
inhomogeneous lattice distortions (G6bel 1976).
FeCrzS4 is a ferrimagnet (Lotgering 1956) in which the Fe z+ spins on the
tetrahedral sites are coupled antiparallel to the Cr 3+ spins on the octahedral sites
in a collinear N6el configuration (Shirane et al. 1964, Broquetas-Colominas et al.
1964). Values of the asymptotic Curie temperature, reported in the literature, lie
between - 2 3 4 K and - 3 3 0 K, that of the ferrimagnetic Curie temperature be-
tween 170 K and 195 K, while the saturation moment of powder samples lies
between 1.55 and 1.79/xB/molecule (Lotgering 1956, Shirane et al. 1964, Haacke
and Beegle 1967, Lotgering et al. 1969, Hoy and Singh 1968, Gibart et al. 1969,
Shick and Von Neida 1969, Spender and Morrish 1971). Measurements on single
crystals (fig. 73) showed the compound to have a strong magnetic anisotropy at

H f l [1001

30

t [111]
20
',.3

D
121
O
o H = 18kOe
'o
* H =12 '"
10
[] H:6 "
• H=3 '"

0,
0 50 100 150 200
temperature (K)
Fig. 73. Temperature dependence of the magnetization of a single crystal of FeCr2S4with H parallel
to the [100] and [111] direction (Van Stapele et al. 1971).
SULPHOSPINELS 707

low temperatures (Van Stapele et al. 1971), as anticipated by Eibschfitz et al.


(1967b). The magnetization lies preferentially along a cube edge, and the satura-
tion magnetization is 1.86/~s/molecule. This corresponds to a magnetic moment of
4.14/~B for the Fe 2+ ions, if the Cr 3+ ions are assumed to have a moment of 3/~B
(Van Stapele et al. 1971). These values agree well with those derived from
neutron diffraction (Shirane et al. 1964). The magnetic anisotropy increases
strongly with decreasing temperature (fig. 73). At 4 K the cubic anisotropy
constant K1 is 3× 106erg/cm 3 and K2 is positive and of the same order of
magnitude (Van Stapele et al. 1971). The anisotropy finds its origin in the 5E
ground state of the tetrahedral Fe 2+ ions, whose splitting depends strongly on the
direction of the magnetization (Eibschfitz et al. 1967b, Hoekstra et al. 1972). The
straightforward crystal field theory, however, predicts a much larger anisotropy,
which has indeed been observed for small concentrations of tetrahedral Fe 2+ ions
in Cdl-xFexCrzS4 (Hoekstra et al. 1972). The fact that the magnetic anisotropy in
FeCrzS4 is relatively weak is connected with the cooperative Jahn-Teller effect
observed in the M6ssbauer spectrum of this compound (see below).
The 57Fe M6ssbauer spectrum of FeCrzS4 in the paramagnetic state is in
agreement with the perfect cubic symmetry of the tetrahedral sites occupied by
the Fe 2+ ions (Yagnik and Mathur 1967, Eibschfitz et al. 1967b). In the fer-
rimagnetic state, however, the spectra show two anomalies: an electric field
gradient on the Fe 2+ nucleus and a hyperfine field that decreases with decreasing
temperature (Eibschfitz et al. 1967b). Both effects are due to the spin-orbit
interaction that removes the orbital degeneracy of the 5E ground state of the Fe z+
ions in the magnetic state, in which the Fe z+ spins are oriented along the direction
of the magnetization. This leads to a lowest orbital state with a tetragonal
electronic charge distribution, a small induced orbital moment and a strong
preference of the magnetization for the [100] directions (as discussed above). With
such an electronic ground state the Fe z+ nucleus experiences a uniaxial electric
field gradient with its main axis in the [100] directions, increasing in strength with
decreasing temperatures, and a hyperfine field that decreases towards lower
temperatures. Isolated Fe z+ ions on the tetrahedral sites, as in Cdl-xFexCrzS4 and
Col-xFexCrzS4 with a small x, show the full effect, in agreement with the theory
(Van Diepen and Van Stapele 1972, 1973). In FeCr2S4, however, the M6ssbauer
spectrum is more complicated. Above 20 K a uniaxial electric field gradient has
been observed in some cases (Eibschfitz et al. 1967b, Spender and Morrish 1972a,
Van Diepen and Van Stapele 1973), in other cases a gradient with a lower
symmetry (Hoy and Chandra 1967, Hoy and Singh 1968, Spender and Morrish
1972b). A sudden change in the spectrum occurs at about 10 K, which has been
ascribed to a static Jahn-Teller distortion, that is too small to be detected by
powder X-ray or neutron diffraction (Spender and Morrish 1972b, Van Diepen
and Van Stapele 1973, Brossard et al. 1979). Subsequent investigations have
shown that the M6ssbauer spectra and the specific heat are heavily influenced by
the Fe/Cr ratio of the samples (Lotgering et al. 1975, Van Diepen et al. 1976).
Carefully prepared samples in which Fe z+ ions exclusively occupy tetrahedral sites
have M6ssbauer spectra with the narrowest lines. The specific heat of such
708 R.P. VAN STAPELE

O
E 8
-'3
-- 6 • . .."
(D
¢-
& : .if W"
o
E /
2
tn

%
j
1'0 ' 2'0 30
temperature (K) m
Fig. 74. Specificheat of Fe0.97Cr2S4as a function of temperature, according to Lotgering et al. (1975).

samples has a lambda-type peak at 9.25 K (fig. 74), at which t e m p e r a t u r e the


Jahn-Teller transformation occurs (Lotgering et al. 1975, Van Diepen et al. 1976)•
The cooperative J a h n - T e l l e r effect in FeCr2S4 and the single ion behaviour can
reasonably well be described within one model with quite satisfactory values for
the spin-orbit splitting, the J a h n - T e I l e r coupling and the energy of the J a h n -
Teller active vibrations (Feiner 1977, 1982)•
FeCrzS4 is a semiconductor with a resistivity that has an anomalous m a x i m u m
(fig. 75) around the ferrimagnetic Curie t e m p e r a t u r e (Lotgering 1956, Albers et al.
1965). At both sides of the anomaly the Iog p vs T -1 curve is a straight line•
Powder samples with a positive Seebeck coefficient of +338 ixV/deg at room
temperature have an activation energy of 0.018 eV below 135 K and 0.038 eV
above 200 K (Bouchard et al. 1965)• In p-type powder samples the thermoelectric
power also shows an anomalous increase near the Curie t e m p e r a t u r e (Haacke and
Beegle 1966)• Bongers et al. (1969) observed a negative magnetoresistance effect
of Ap/p = - 0 . 0 5 in 12 k O e near the Curie t e m p e r a t u r e in p-type FeCrzS4. These

30
'oL
A

I
20
!

I
i

1o t~
\i~
£

' 200 ' 400I "'7- 600


temperature (K)
Fig. 75. The resistivity of a sintered sample of FeCr2S4 (Lotgering 1956).
SULPHOSPINELS 709

authors reported that powder samples with a n-type conduction do not show a
maximum in the resistivity and the magnetoresistance effect near Tc (Bongers et
al. 1969).
Single crystals in the as-grown state are always reported to have a p-type
conduction (Haacke and Beegle 1968, Goldstein and Gibart 1971, Watanabe
1973). The resistivity has been found to show a similar anomalous increase near
the Curie temperature as in powder samples (fig. 76). The ordinary Hall
coefficient has been observed to be too small to be measured (Haacke and Beegle
1968, Goldstein and Gibart 1969, 1971), the Hall resistivity being determined
mainly by the spontaneous Hall effect (Goldstein and Gibart 1969, 1971). The
resistivity and the magnetoresistance effect were found to be very sensitive to heat
treatments (Watanabe 1973, Gibart et al. 1976). Annealing in a sulphur atmos-
phere slightly lowers the resistivity, whereas annealing in vacuum increases it
(Watanabe 1973, Gibart et al. 1976) (fig. 76). In a qualitative way these obser-
vations can be understood from an energy level diagram (fig. 77) in which an Fe >
band is situated in the energy gap between the valence and the conduction band
(Lotgering et al. 1969). Hole conduction is attributed to holes in the Fe 2+ band,
which are present because of deviations from the stoichiometric composition in
as-grown samples (Watanabe 1973, Lotgering et al. 1975, Gibart et al: 1976).
Sulphur deficiency gives rise to donor states, which are assumed to be situated
above the Fe 2+ band. Annealing in vacuum will then decrease the number of:holes
in the Fe z+ band and increase the resistivity, a s observed, whereas annealing in a
sulphur atmosphere will have the reverse effect (Watanabe 1973).

102

I 1°1

.> 10o

10 -1

I
o 10 15
103/T (K41
Fig. 76. The resistivityof single crystals of FeCr2S4: (a) as-grown, (b) S-annealed at 700°C for 72 h and
(c) vacuum-annealed at 575°C for 68 h 0Natanabe 1973).
710 R.P. VAN STAPELE

T
duction band

Fe2+ EF

Cra÷V / ' / / ' / / / / / / / / ~ band

glE)
Fig. 77. Energy bands in FeCr2S4 (Lotgering et al. 1969).

Apart from the sharp negative magnetoresistance effect near the Curie tem-
perature, p-type single crystals have an additional broad and positive mag-
netoresistance effect, which is maximal at a lower temperature (0.35 To) (Lyons et
al. 1973, Goldstein et al. 1973) (fig. 78). The second effect was ascribed to a
spontaneous anisotropic resistivity, which means a resistivity depending on the
direction of the magnetization relative to the direction of the current and the
crystal axes. A theoretical explanation for this anisotropy has not been given.
The first effect was ascribed to spin disorder scattering of charge carriers in a
broad band (Lyons et al. 1973, Goldstein et al. 1973), although an adequate theory
will have to take acount of strong correlation and a strong exchange interaction,
typical of charge carriers in a narrow Fe 2+ band (Bongers et al. 1969).
The influence of doping has been reported for the dopants Cd and In (Gold-
stein and Gibart 1971) and Cu and Zn (Watanabe 1973). The resistivity of
Zn-doped FeCr2S4 is higher than that of undoped samples, while the mag-
netoresistance peak is slightly shifted to lower temperatures. Doping with Cu
strongly reduces the resistivity in accordance with the valency distribution
2+ 3+ , 1+
Fel-2xFex Cux Cr2S4 (section 6.6) (Lotgering et al. 1969). The magnetoresistance
effect near the Curie temperature is strongly reduced (Watanabe 1973).
Nuclear magnetic resonance of 53Cr nuclei in FeCr2S4 has been measured as a
function of temperature below 150K. Extrapolated to OK the resonance
frequency is 50.8 MHz, which corresponds to a hyperfine field of 211 kOe (Le
Dang Khoi 1966).
SULPHOSPINELS 711

0.05

2-2
o f

r
-0.05

- 0.10 1
100 200
temperature (K)
Fig. 78. Spontaneous anisotropic resistivity (Aps/p(O)) and magnetoresistance (&p/p) in 15kOe of
FeCraS4 (Lyons et al. 1973).

The M6ssbauer spectrum of iagsn in Fel.lCrl.sSn0.1S4 consists of broad lines with


an additional spectrum, indicating variations in the surrounding of the Sn ions.
The large isomer shift relative to SnO2 points to an increased s-electron density.
The hyperfine field was observed to be large and negative (-470 _+ 15 kOe at 80 K)
(Lyubutin and Dmitrieva 1975).
In the near infrared the reflectance circular dichroism of hot-pressed FeCr2S4
has been measured at 80 K. The normal remanence of the dichroism or the polar
Kerr effect of the hot-pressed samples was found to be large (0.80) (Coburn et al.
1973).
We conclude the review of properties of FeCr2S4 by mentioning the frequencies
of the four infrared-active phonons observed in the absorption spectrum at 118,
260, 323 and 382 cm -~ (Lutz and Feher 1971).

6.4. CoCr2S4

In normal circumstances CoCr2S4 has the normal (Raccah et al. 1966) spinel
structure (Hahn 1951). The lattice parameters are given in table 1. A trans-
formation to the ordered cation-defective NiAs structure can be effected by the
application of high pressure at high temperatures (Albers and Rooymans 1965,
Bouchard 1967, Tressler et al. 1968, Tressler and Stubican 1968).
The paramagnetic susceptibility of CoCr2S4 has a shape that is characteristic of
a ferrimagnetic substance (fig. 79). Reported values of the asymptotic and the
ferrimagnetic Curie temperature are - 3 9 0 ± 40 K and 240 _ 5 K (Lotgering 1956),
- 4 8 0 K and 227 K (Pellerin and Gibart 1969), and T~ = 220 ± 1 K (Shick and Von
Neida 1969). The spontaneous magnetization extrapolated to 0 K amounts to
712 R.P. VAN STAPELE

t
:3
U
(~
0
200 l
E
ea
3
100
I I
I/
Ig
I!
I~ I I I
o 2o0 ,,oo 660 860 ooo
temperature(K)
Fig. 79. Magnetization M and inverse molar susceptibilityX~1 of CoCr2S4,according to Pellerin and
Gibart (1969).

2.55 +0.06 p,B/molecule (Lotgering 1956), 2.43 ~B/molecule (fig. 79) (Pellerin and
Gibart 1969), and 2.4/~B/molecule (Shick and Von Neida 1969). With 6/~B of the
Cr 3+ ions opposite to the Co moment in a simple N6el configuration, the Co ions
have a moment of 3.45 to 3.57/~B. This corresponds to a g-factor of 2.30 to 2.38; a
reasonable value for Co 2+ ions on tetrahedral sites (Gilbart et al. 1969). A
molecular field analysis of the magnetic data indicates a dominating exchange
interaction between the Cr s+ and Co 2+ ions and a weak interaction between the Cr
ions (Gilbart et al. 1969).
As-grown single crystals of CoCr2S4 were observed to have a cubic mag-
netocrystalline anisotropy with a cubic anisotropy constant of 3.45 × 105 erg/cm s at
77 K. Annealing in an oxidizing or a reducing atmosphere was found to influence
the cubic anisotropy, while the annealed crystals show a weaker induced uniaxial
anisotropy after cooling down to 77 K in a magnetic field (Gibart et al. 1976).
CoCr284 is a semiconductor with a resistivity that is strongly sample-dependent.
The slope of the log p vs T -1 curve is generally larger in the paramagnetic state
than in the ferrimagnetic state, which observation has been made in poly-
crystalline samples (Bouchard et al. 1965, Albers et al. 1965) as well as in single
crystals (fig. 80) (Pellerin and Gibart 1969, Watanabe 1973). Polycrystalline
samples have either p-type (Bouchard et al. 1965) or n-type conduction (Albers et
al. 1965), while single crystals show n-type conduction (Watanabe 1973, Gibart et
al. 1973). The resistivity of undoped samples does not show an anomalous
maximum near the Curie temperature. However, a magnetoresistance effect near
T~ was observed in p-type doped polycrystalline samples (Bongers et al. 1969) and
n-type single crystal (Watanabe 1973).
Heating of single crystals in vacuum or in a sulphur atmosphere strongly affects
the resistivity (fig. 80). The change was found to be opposite to that in FeCr2S4
(see section 6.3), namely an increase after annealing in a sulphur atmosphere and
a decrease after annealing in vacuum (Watanabe 1973, Gibart et al. 1976).
The mechanism of the n-type conduction is unknown, but the observed proper-
SULPHOSPINELS 713

I I

10 5

10 4

T 10 3
o

102
>,,,
4--

101

10o

I [
0 5 110 15
103/T(K - )
Fig. 80. Resistivity of single crystals of CoCr2S4: (a) as-grown; (b) S-annealed at 700°C for 72 hr and
(c) vacuum-annealed at 600°C for 68 hr (Watanabe 1973).

ties point to donor states due to a sulphur deficiency, already present in as-grown
crystals (Watanabe 1973).
The optical properties of hot-pressed samples were investigated by Carnall et
al. (1972). In the spectral range between 7 and 12 ~m the samples are transparent
with a residual absorption coefficient of about 7 cm -1, a refractive index of 3.56 and a
Faraday rotation that decreases with increasing wavelength from 2100 deg/cm at
5 ~m to 320 deg/cm at 10.6 ~m in the magnetic saturated state at 80 K. At longer
wavelength the absorption spectrum shows four bands at 388, 330, 258 and
120 cm 1 due to the four infrared-active phonons of the spinel structure (Lutz and
Feh6r 1971, Carnall et al. 1972). The onset of the strong absorption at the short
wavelength side is due to crystal field transitions in the Co 2+ ions. Large Kerr
rotations (fig. 81) were observed in dispersion-like peaks at 1.0 and 1.7 ~m,
associated with the 4Az~4T1 transition of tetrahedrally coordinated Co 2+ ions
(Ahrenkiel and Coburn 1973, Ahrenkiel et al. 1974).
These transitions were also observed in the reflectance circular dichroism spectra
(Ahrenkiel et al. 1973, Coburn 1973). The large remanence in the magneto-optical
properties at normal incidence and the large Kerr effect has made hot-pressed
CoCr2S4 interesting from the point of view of optical data-storage materials
(Ahrenkiel et al. 1974).
Nuclear magnetic resonance spectra of 53Cr nuclei in CoCr2S4 were measured at
77 K (Le Dang Khoi 1968) and of 53Cr and 59Co at 4.2, 55 and 78 K (Yokoyama et
al. 1970). The last authors report the 53Cr resonance line to have a triplet
structure, which was not analyzed. The central frequency is 50.0 MHz at 4.2 K,
714 R.P. VAN STAPELE

*8
•" \
+6

+2

o
0)
o
-2
"[3
\....' U
-6

-8

10 1.5 2.0
wavelength (p)
Fig. 81. Double polar Kerr rotation (a) and Kerr ellipticity (b) of C o C r 2 S 4 at 80 K in a saturating
magnetic field (Ahrenkiel and Coburn 1973).

which corresponds to a (negative) hyperfine field of 208 kOe. The nuclear mag-
netic resonance signal of 59Co was found around 31.2 MHz at 4.2 K. The increase
of the resonance frequency in an applied magnetic field shows the hyperfine field
to be positive. The small value of the hyperfine field (31 kOe at 4.2 K) is typical of
tetrahedrally coordinated Co 2+ ions in which a positive orbital hyperfine field and
a negative core polarization hyperfine field have nearly the same magnitude
(Yokoyama et al. 1970).
A large negative hyperfine field (-405 + 20 kOe at 80 K) was observed in the
M6ssbauer spectrum of ngsn in a sample with the composition COl.lCra.sSn0.1S4.
The spectrum is strongly broadened with an additional splitting of the lines,
pointing to non-equivalent positions of the Sn ions in the lattice.The isomer shift
relative to SnO2 is large and indicates an increased density of s electrons at the Sn
ions (Lyubutin and Dmitrieva 1975).
Finally, X-ray absorption spectroscopy has shown that the shift of the cobalt K
absorption discontinuity agrees well with the divalency of the Co ions (Ballal and
Mande 1977).

6.5. The mixed crystals Fel-xCoxCr2S4, Fel_x(Cl~l/2Irtl/2)xer2S4, Fel-xCdxCr2S4,


Coa-xCdxCr2S4 and Col-x( Cul/2Fel/2)xCr2S4

In the mixed crystal series Fel xCoxCr2S4 the lattice parameter decreases roughly
linearly with x (Treitinger et al. 1976b). The Curie temperature increases almost
linearly with x, the value at x = 0.15 being somewhat lower than that of pure
SULPHOSPINELS 715

FeCr2S4. Samples with x ~<0.95 have a p-type conduction with an anomaly in the
resistivity and a negative magnetoresistance effect (3 to 4%) near the Curie
temperature. Pure CoCr2S4 has an n-type conduction with a small Seebeck
coefficient and a very small negative magnetoresistance effect (0.3%) (Treitinger
et al. 1976b). Samples with x = 0.98 (Van Diepen and Van Stapele 1973) and
CoCr2S4 samples with small unspecified amounts of Fe (Tanaka et al. 1973) have
been used in a study of the 5E ground state of tetrahedrally coordinated Fe 2+ ions
by means of Mrssbauer spectroscopy (section 6.3).
Mixed compounds Fel-x(Cul/2Inl/2)xCrzS4 between the semiconducting fer-
rimagnet FeCrzS4 and the semiconducting antiferromagnet Cul/2Inl/zCr2S4 (see
section 5.9) are single-phase spinels with a lattice parameter that varies linearly
with x (Grbel et al. 1975). As measured by the intensity of the (200) X-ray
reflection, ionic ordering of the Cu + and In 3+ ions on the tetrahedral sites exists
down to x = 0.4. For smaller (Cut/2Inl/2) concentrations the Cu +, In 3+ and Fe 2+ ions
are statistically distributed over the tetrahedral sites (Grbel et al. 1975). The
compounds are ferrimagnets in the range 0 ~<x < 0.8 and antiferromagnets for
x > 0.8 with critical temperatures as given in fig. 82 (Grbel et al. 1975). The
compounds are semiconductors with a room temperature resistivity that increases
with increasing x. In the iron-rich samples a negative magnetoresistance effect has
been observed with a high-temperature maximum shifting in position with respect
to the Curie temperature (fig. 83). An explanation of this phenomenon was not
given (Treitinger et al. 1976a).
In the mixed compounds Fel xCdxCr2S4 between the ferrimagnet FeCr2S4 and
the ferromagnet C d C r 2 S 4 the lattice parameter varies linearly with x (Spender and
Morrish 1971, Barraclough et al. 1974). With decreasing Fe content, the magnetic
properties gradually change from ferrimagnetic to ferromagnetic. The asymptotic
Curie temperature increases and the Curie temperature decreases as shown in
figs. 84 and 85 (Bongers et al. 1969, Spender and Morrish 1971, Barraclough et al.
1974).
In view of the rather large magnetoresistance effects the resistivity and the

200

150

t 100
K

50
b e~.~e.--,.-

~ I I 01.8
0 0.2 0.4 0.6 1.0
X ~--
Fig. 82. Curie temperature (a) or Nrel temperature (b) of Fel_x(CumIn~/2)xCr2S 4 as a function of the
composition (Grbel et al. 1975).
716 R.P. VAN STAPELE

T4
Ap
X=0.4

Po

X=0.2

~,~ X=0

050 100 150 200 250


temperoture (K)
Fig. 83. Magnetoresistance effect of Fel ~(Cumlnl/a)xCr2S4as a function of the temperature for x = 0,
0.2 and 0.4 (Treitinger et al. 1976).

%(K) 8 (K)
I lgC ,,-200 l
I
[] + .t+
160 [] o • ,,100

140 [] 0
[]
120 -100
+ o
r-I

100 ÷
6 -200

80 I, I I I
0 Q2 04 '0'6 ' 08 1.0
X
Fig. 84. Asymptotic Curie temperature 0 (O, +) and ferromagnetic Curie temperature Tc (©, [~) of
powder samples of Fei-xCdxCr2S4, according to Bongers et al. (1969) (O, (3) and Spender and Morrish
(1971) (+, E3).

magnetoresistance effect were measured in undoped single crystals (Barraclough


et al. 1974), in hot-pressed Ag-doped powders with a p-type conduction (Bongers
et al. 1969) and in Cu-doped single crystals with a p-type conduction (Treitinger et
al. 1978b). The largest magnetoresistance effect (17%) was observed near the Curie
temperature (147K) of an undoped single crystal with the composition
Fe0.46Cd0.54CrzS4 (Barraclough et al. 1974).
X-ray diffraction of a powder sample with the composition Fe0.83Cd0.17Cr2S4
showed the lattice parameter to decrease regularly with decreasing temperature,
whereas the line width of the diffraction lines broadens unusually. This broaden-
ing, which was also observed in CdCr2S4 and FeCrzS4, indicates inhomogeneous
lattice distortions (G6bel 1976).
Complicated 57Fe M6ssbauer spectra were observed in the compounds
SULPHOSPINELS 717

u
5 • + •
O
E
4

3
E
O
E
o
....,
t'-

I I I I
'0.2 o14 o16'o'.8
X
Fig. 85. Magnetic moment of powder samples of Fel-xCdxCr2S4, (0) extrapolated to 0 K (Bongers et
al. 1969) and (+) extrapolated to infinite magnetic fields at 4.2 K (Spender and Morrish 1971). The solid
line represents a linear variation between 1.86 #B/molecule for FeCr2S4 (section 6.3) and 6 #B/molecule
for CdCr2S4.

Fel_xCdxCrzS4 with 0 < x ~<0.9 (Spender and Morrish 1973). The simple M6ss-
bauer spectrum of Fe0.02Cd0.98Cr2S4 was described in terms of a magnetically
induced orbital hyperfine field and quadrupole splitting (Van Diepen and Van
Stapele 1972).
In the series of mixed compounds C o l - x C d x C r 2 S 4 the lattice parameter varies
linearly with x (Tret'yakov et al. 1975). The magnetic properties change gradually
from the ferrimagnetic properties of CoCr2S4 to those of the ferromagnet CdCr2S4,
the Curie temperature decreasing and the magnetic moment increasing with x
(Coburn et al. 1972, Tret'yakow et al. 1975) (fig. 86). In view of applicable

250
-6

2OO 0 0 0

o
+x
÷+

5 T
x 4
o
'~ 150 x
o
E
O
o 3
e e
S
~- 10(? c-
2 :£
N
@ 50 I C

i I I 0
0.'2 ' o16 o'8 1.o
X ~--
Fig. 86. Curie temperature (O, C)) and magnetic moment (+, x) of Coi-xCdxCr2S4, according to
Coburn et al. (1972) (0, +) and Tret'yakov et al. (1975) (O, x).
718 R.P. VAN STAPELE

magneto-optical properties, the absorption spectrum and the Faraday rotation of


hot-pressed samples were measured. The absorption between 0.8 and 5 Ixm is
roughly proportional to the Co concentration and shows at low Co concentrations
absorption bands at 1.62, 1.72 and 1.95 g m due to the 4A2---~4Tl(F) crystal field
transition of Co 2+ and a weaker band near 3 Ixm, probably due to the 4A2---~4T2(F)
crystal field transition of Co 2+. In the spectral range between 4 and 14 ~xm the
Faraday rotation increases with decreasing wavelength and increasing Co concen-
tration (Coburn et al. 1972).
The reflectance circular dichroism spectrum in the near infrared shows large
dispersions at about 1 txm and 1.7 Ixm due to the crystal field transitions 4A2(F )--~
4TI(P ) and 4A2(F ) ~ 4TI(F ) of the Co 2+ ions (Ahrenkiel et al. 1975).
Single-phase spinels Col-x(CUl/2Fea/2)xCr2S4 between the semiconducting fer-
rimagnets CoCr2S4 and Cu +1/2Fe3+ 1/2Cr2S4 (section 6.6) were prepared in an attempt
to realize sizable magnetoresistance effects near a magnetic ordering at room
temperature. The Curie temperature varies linearly between 211 K at x = 0 and
350 K at x = 1. The magnetoresistance effect is rather small and negative with a
maximum value near the Curie temperature (Treitinger et al. 1976a).

6.6. The mixed crystals ml-xCuxCr2S4 with M = Mn, Fe and Co

Single-phase spinels Mnl-xCuxCr2S4 between the semiconducting ferrimagnet


MnCr2S4 and the metallic ferromagnet CuCraS4 have only been obtained for x i> 0.8
and x < 0.4. The Mn-rich compounds are metastable, heating at 300°C effects a
decomposition into two spinel phases. In the single-phase regions the variation of the
cell edge is smaller than would correspond to a linear variation between the lattice
parameters of MnCr2S4 and CuCr2S4. From the magnetic moments measured for
x = 0, 0.05, 0.2 and 0.8 between 1.6 and 50 K in magnetic fields up to 150 kOe, it
was concluded that the Mn ions remain in the divalent state in the mixed compounds
(Nogues et al. 1979, Mejai and Nogues 1980).
A different situation is encountered in the series Fel-xCuxCr2S4. The electrical
and magnetic properties of these compounds clearly indicate that the Cu ions
ionize the ferrous ions to ferric ions (Lotgering et al. 1969)*. The series does not
show a miscibility gap. Throughout the series, single-phase samples were prepared
and the cell edge varies approximately linearly with x (Haacke and Beegle 1967).
The electrical transport properties vary in a remarkable way (fig. 87) (Haacke and
Beegle 1968, Lotgering et al. 1969). The occurrence of n-type conduction is
observed at composition with x between 0.2 and 0.5, whereas samples with a
smaller or larger Cu concentration have a p-type conduction (Haacke and Beegle
1968, Lotgering et al. 1969), and the observation that Fel/2Cut/2Cr2S4 is a semi-
conductor (Lotgering et al. 1969) led Lotgering et al. (1969) to the conclusion that
the Fe z+ levels fall in the energy gap between the valence and the conduction
band, as sketched in fig. 88. A replacement of 2Fe 2+ in FeCrzS4 by Fe3++ Cu + is
represented by a hole in the Fe z+ level (section 3) and a filled Cu + level below the
top of the valence band. The Fe 2+ levels are empty at x = ½. When x increases

* See notes added in proof (1) on p. 737.


SULPHOSPINELS 719

500,

400t
3OO

T 2OO

100
:3,
ml

o -100

.£3 -200
PA
-300

-4ooi I
0 0.5 1 0 0.5
X----~ X ~-
Fig. 87. Seebeck coefficient c~ and resistivity p at room temperature of Fel-xCuxCr2S4, according to
Haacke and Beegle (1%8) (©), Bouchard et al. (1%5) (0) and Lotgering et al. (1969) (11). Sample A
has been quenched from 700°C, sample B has been slowly cooled with annealing at 500 ° and 100°C.

further, Fe 3+ is replaced by Cu ÷ with charge compensation by two holes in the


valence band. The formal valence distribution is consequently written as
2+ 3+ + 3+ 2 3+ +
Fel-2xFex CuxCr2 $4 in the range 0 ~< x < ½ and as Fel_xCuxCr2S4 with ( 2 x - 1)
holes in the valence band and the Cr 3+ states for ½ < x ~< 1". This explains the
observed p-type conduction for x >½ and for small x as well as the n-type
conduction in the intermediate region, where the conduction is due to the
simultaneous presence of ferro and ferric ions with less than half of the iron ions

nduction
bGnd
Fe3. (X)
2~ --g Fe3+{1/2)~ Fe3+(I_X)r-
Fe *(1-2X EF (2X-1)holes
Cr3*E;~NTZ,{~)~Volence Cr3÷ C r ~ F
~r.'//,/~ hood

g(E)~,,- g(E) ~ g(E) --~


0~X<1/2 X= 112 1/2<X<~1
Fig. 88. Energy level schemes of Fel-xCuxCr2S4 after Lotgering et al. (1969). The number of states per
molecule is indicated between brackets.

* See notes added in proof (m) on p. 737.


720 R.P. VAN STAPELE

in the divalent state. At x = ½, with all iron ions in the trivalent state, the gap
between the empty Fe 2+ states (i.c. Fe 3+) and the filled Cu ÷, Cr 3+ and valence band
states explains the semiconducting behaviour, with a large positive or negative
Seebeck coefficient depending on the details of the preparation (Lotgering et al.
1969).
The magnetic properties show a gradual change from ferrimagnetism in FeCrzS4
to ferromagnetism in CuCrzS4. The Curie temperatures are enhanced with respect
to a linear variation with x (Haacke and Beegle 1967, Lotgering et al. 1969) (fig.
89), which was attributed to a negative Fe3+-Cr 3+ superexchange interaction that is
stronger than the Fe2+-Cr 3+ interaction (Lotgering et al. 1969).

z,0C jo I
,, o :.,.
D
f
30c o f
(9
CL • f
E ...,..:.i
20(

o 10(

01~2 I Oil& I 01,6 I 018 ' 1.0


X J,-
Fig. 89. Curie temperature of Fei-xCuxCr2S4, according to Haacke and Beegle (1967) (0) and
Lotgering et al. (1969) (O).

It is not possible to account for the magnetization data (Haacke and Beegle
1967, Lotgering et al. 1969). It is, however, difficult to measure the saturation
magnetization correctly. At small Cu concentrations, the strong magnetic aniso-
tropy of tetrahedrally coordinated Fe 2+ ions (section 6.3) makes polycrystaltine
samples difficult to saturate, and it is not possible to prepare pure Cu-rich samples
(Lotgering et al. 1969). The saturation m o m e n t of Fel/2CumCr2
3+ 1+ 3+S4 is 3.2/xB/mole-
cule, 10% lower than the 3.5/xB/m01ecule expected for a simple N6el configuration
with Fe in the trivalent state (Lotgering et al. 1969).
M6ssbauer spectra of 57Fe, measured between 4 and 373 K in Fel/zCUl/zCr2S4
(Lotgering et al. 1969) and above the Curie t e m p e r a t u r e as well as at 77 K for the
same and other compositions (Haacke and Nozik 1968) show the presence of Fe 2+
and Fe 3+ ions in Fel-xCuxCr2S4 with x <½ and of exclusively Fe 3+ ions for x ~>½.
The hyperfine field on the iron nucleus in Fel/zCul/zCr2S 4 is positive, i.e. opposite
to the m o m e n t of the Fe sublattice, in agreement with a simple N6el configura-
tion. The hyperfine field versus t e m p e r a t u r e curve is m o r e concave than the
magnetization curve (Lotgering et al. 1969).
The absence of a quadrupole splitting and of dipolar contributions to the
hyperfine field in the M6ssbauer spectrum of Fel/2Cua/zCrzS4 proves the local
symmetry on the iron sites to be perfectly cubic, which is a strong indication of a
1 : 1 ordering of the differently charged Fe 3+ and Cu 1+ ions on the tetrahedral sites
SULPHOSPINELS 721

(Lotgering et al. 1969). Because of the too small difference in scattering power of the
Fe 3+ and Cu 1+ ions, this ordering cannot be detected by means of X-ray diffraction.
However, superstructure X-ray reflections were observed in In 3+ mCu 1+
1/2CrzX4 with
X = S or Se (Lotgering et al. 1969) and later on in a number of similar compounds (see
section 5.9).
The ionic ordering in Fel/2Cul/zCr2S4 permits the observation of Cu nmr lines,
which would otherwise have been strongly broadened. The hyperfine field on the
Cu nucleus was measured as a function of temperature. In terms of spin density in
the empty 4s shell, the magnitude of the hyperfine field is comparable with that on
the nuclei of other diamagnetic ions, like Cd 2+ and Hg 2+ (Locher and Van Stapele
1970).
Finally we mention the recent confirmation of a N6el spin configuration in
Fe0.sCu0.zCr2S4 by means of neutron diffraction (Babaev et al. 1975).
The compounds COl-xCuxCrzS4 have been investigated in much less detail. Lutz
and Becker (1973) have reported the existence of a complete series of solid
solutions. The lattice parameter varies in a non-linear way. As a function of the
composition the Seebeck coefficient does not change sign as in Fel-xCuxCrzS4, but
is positive throughout the series. The conductivity of pressed samples increases
rapidly with x with a positive temperature coefficient up to x = 0.2, and is metallic
at larger Cu concentrations. At x = 0.2 the cell edge starts to decrease to the
smaller value of CuCr2S4, which is reached at x = 0.80 (Lutz and Becker 1973).
The only information on the valencies of the ions comes from the chemical shift of
copper and cobalt K absorption discontinuities in COl/2CUl/2CrzS4, which shows
that Co is in the divalent state, whereas Cu is monovalent (Ballal and Mande
1977). This result and the absence of a change in sign in the Seebeck coefficient
indicates that the Co 2+ levels are well below the top of the valence band.

6. 7. The mixed crystals Ml-xNixCrzS4 with M = Mn, Fe, Co, Cu and Z n

Attempts were made to prepare solid solutions with the spinel structure between
a number of sulphospinels MCrzS4 and NiCr2S4 that itself crystallizes in the
ordered cation-defective NiAs structure. It was found for M = Mn, Fe, Co, Cu
and Zn that the spinel structure is stable in a limited range of Ni concentrations.
This range is given in table 21 for each of the systems investigated. The table also
summarizes some other properties. The lattice parameter generally decreases and
the Curie temperature increases with increasing x.
The only systems whose physical properties have been studied in some detail
are Znl_xNixCrzS4 (Itoh et al. 1977) and Mnl_xNixCr2S4 (Mejai and Nogues 1980).
The latter authors measured the magnetization of Mn0.9Ni0.1Cr2S4 at 7 and 9 K in
magnetic fields up to 150 kOe. They discussed the influence of the Ni ions on the
transitions between the various spin configurations in MnCr2S4 (see section 6.2).
ZnCr2S4 is an antiferromagnetic semiconductor (section 5.2). Substitution of Ni
for Zn does not change the type of conduction, as the resistivity of Zn0.6Ni0.4Cr2S4
(fig. 90) clearly shows. Itoh et al. (1977) consequently conclude that Ni is divalent,
as represented by Zn0.~q'10.4Cr2
2+ .2+ 3+$4. The paramagnetic susceptibility is charac-
722 R.P. V A N S T A P E L E

T A B L E 21
Some crystallographic and magnetic properties of the spinel systems M>xNixCr2S4 with
M = Mn, Fe, Co, Cu or Zn. References: (a) Lisnyak and Lichter (1969), (b) Lutz et al. (1973),
(c) R o b b i n s and Becker (1974) and (d) Itoh et al. (1977).
o
Lattice constant (A) Curie t e m p e r a t u r e (K)

Stability
System range x = 0 at maximal x x = 0 x = 0.3 Ref.

Mnt-xNixCr2S4 x ~0.3 10.11 10.068 74 110 (c)


Fel-xNix Cr2S4 x ~ 0.3 9.995 9.953 185 214 (c)
Col-xNixCr2S4 x ~<0.4 9.918 9.898 (a)
x ~< 0.2 9.923 9.909 (b)
x ~ 0.4 9.936 9.898 235 250 (c)
Cul-xNixCrzS4 x ~ 0.26 9.820 9.801 (b)
Znl-.NixCr2S4 x ~ 0.4 9.986 9.945 TN = 16 -- 170 (d)
at x = 0.4

E
13 °°
°%°
°°.°,
0 %°°
°.

i I i I I I I I I I I
100 200 300
temperoture (K)
Fig. 90. Electrical resistivity of Zn0.6Ni0.4Cr2S4 as a function of temperature (Itoh et al. 1977).

teristic of a ferrimagnet which, for x = 0.4, has an asymptotic Curie temperature


of - 2 9 9 K and a Curie constant that is consistent with the spin-only values of Cr 3+
and Ni >. The magnetization versus temperature curves measured in 16 k O e (fig.
91) show a m a x i m u m at low temperatures. The magnetization cannot be saturated
except for the composition x = 0.4, where it reaches at 4 . 2 K a value of
3.0/xB/molecule. This is much smaller than the 5.2/xB/molecule expected for a
simple Nrel configuration with 2/xB on the Ni 2+ ions (Itoh et al. 1977).

6.8. The mixed crystals MCrz-xlnxS4 with M = Mn, Fe, Co and Ni

In this section we review briefly the properties of mixed crystals between the
ferrimagnets MCr2S4 with M = Mn, Fe and Co and the corresponding indium
sulphospinels MIn2S4. A m o n g the latter c o m p o u n d s MnIn2S4 is a partially inverse
spinel, the other c o m p o u n d s MIn2S4 (M = Fe, Co and Ni) being inverse spinels
SULPHOSPINELS 723

60 I i i , i

•',•.

50
• X=O./.

l 4O •',;,

X= 0.3 '
c~ 30
E
(11

2C
,....
X=0.2

10
x:ol ........ "'.... "...:....
' ' ' ' ' '2oo
temperature (K) ~-
Fig. 91. Magnetization o- per gram of Znl xNixCr2S4 measured as a function of temperature in 16 kOe
(Itoh et al. 1977).

(Hahn and Klingler 1950)• The indium sulphospinels are all paramagnetic down to
4.2 K (Schlein and Wold 1972)• The magnetic susceptibilities follow a Curie-Weiss
law with a negative asymptotic Curie temperature. The molar Curie constant
agrees with the spin-only value in the case of Fe and Ni, but deviates from it in
the other cases (table 22). Electrical resistivity measurements at room temperature
indicate that the c o m p o u n d s are semiconductors (Schlein and Wold 1972). The
large negative asymptotic Curie temperature of NiIn2S4 is anomalous, since the
90 ° Niz+-S-Ni 2+ exchange interaction is expected to be positive• This anomaly and
the lack of antiferromagnetic ordering have been discussed by G o o d e n o u g h (1972)•
In the system MnCrz_xInxS4 single-phase spinels were prepared between x = 0
and x = i o(Darcy et al. 1968)• Theo lattice parameter changes linearly from
a = 1 0 . 1 0 8 A at x = 0 to a = 1 0 . 4 1 8 A at x = 1. It was concluded from X-ray
diffraction data that the In 3+ ions replace Cr 3+ ions on the octahedral sites, so that

T A B L E 22
Cell edge (a), asymptotic Curie temperature (0) and molar Curie
constant (Cm) of the compounds MIn2S4, according to (1) Schlein
and Wold (1972) and (2) Eibschfitz et al. (1967a).

Cm Cm
Compound a (A) 0 (K) (exp.) (spin-only) Ref.

Mnln2S4 10.72 -78 4.00 4.38 (1)


FelnzS4 10.61 -76 3.10 3.00 (1)
10.630 - 122 2.94 (2)
Coln2S4 10.58 - 134 2.84 1.87 (1)
Niln2S4 10.50 -144 1.16 1.00 (1)
724 R.P. VAN STAPELE

the tetrahedral sites are always occupied solely by Mn 2+ ions. MnCr2S4 is a canted
ferrimagnet in which a strongly positive Cr3+-Cr3+ superexchange interaction
combines with weaker negative Mn2+-Cr 3+ and MnZ+-Mn 2+ superexchange inter-
actions (section 6.2). Substitution of In for Cr reduces the magnetic m o m e n t (at 4.2 K
and 10 k O e from 1.27/xB/molecule at x = 0 to 0.85/zB/molecule at x = 0.3) as well as
the Curie t e m p e r a t u r e (fig. 92), effecting a r e m a r k a b l e change from ferrimagnetism
to antiferromagnetism at x = 0.4. The measured paramagnetic m o m e n t s are low
c o m p a r e d to the theoretically expected values (Darcy et al. 1968).

100

'\%

I 50
\
\

o o
~ ~ o

~J
c~. I~L I I I I I I I
E 0 ×~"
~ x ~ 0 ~°
1.'0
X - - ~

x~,~

-50
Fig. 92. Curie temperature Tc, Ndel temperature TN and asymptotic Curie temperature 0 of
MnCrz-xInxS4, according to Darcy et al. (1968).

MnCrInS4 was also prepared by Mimura et al. (1974). These authors confirm
the cation distribution determined by Darcy et al. (1968), but their samples are
paramagnetic down to 4.2 K.
In the system FeCrz_xInxS4 a complete series of mixed crystals can be prepared.
The lattice p a r a m e t e r increases linearly from 9.998 A at x = 0 to 10.610 A at x = 2
(Brossard et al. 1976). From the intensity of X-ray diffraction lines and from
paramagnetic 57Fe M6ssbauer spectra the cation distribution was determined
(Brossard et al. 1976). The fraction y of In ions on tetrahedral sites in
Fel yIny[Cr2 xInx_yFey]S4 increases gradually with x (fig. 93).
Magnetic m e a s u r e m e n t s (Goldstein et al. 1977a) show that the compositions
0 <~ x ~< 0.8 are ferrimagnetic. The Curie t e m p e r a t u r e decreases slowly with x. The
measured saturation magnetizations agree with a collinear N6el spin structure
with Fe 2+ spins on tetrahedral sites antiparallel to Fe 2+ and Cr 3+ spins on
octahedral sites. Compositions in the range 1.3 ~< x ~< 2 are antiferromagnetic with
relatively low N6el temperatures (20 K at x = 1.6). In the range 0.8 ~< x <~ 1.3 the
magnetic behaviour changes from ferrimagnetic to antiferromagnetic. The com-
pound FeCrInS4 with the cation distribution Fe0.41In0.59[CrIn0.41Fe0.59]S4 is ferro-
SULPHOSPINELS 725

1
Y
I 0.8

0.6

0.4

0.2

00-~ ~ ' 1' ' ' ' ' 2


~X
Fig. 93. Cation distribution Fez yIny[Cr2xInx yFey]S4,according to Brossard et al. (1976).

magnetic with Tc = 75 K and a saturation moment of 0.8/xB/molecule. These


results are at variance with the findings of Mimura et al. (1974), who found
FeCrInS4 to be an antiferromagnet with a positive 0 of 82 K and TN = 20 K. In
their sample the cation distribution is inverse, e.g., In[CrFe]S4.
The M6ssbauer spectrum of octahedral Fe z+ ions in FeCrz_xInxS4 measured at
room temperature as a function of x (Brossard et al. 1976) essentially agrees with
the spectrum of FeCrInS4 measured by Mimura et al. (1974) and with the
spectrum of FeIn2S4 measured by Eibschtitz et al. (1967a) and by Yagnik and
Mathur (1967). The large quadrupole splitting of this spectrum is due to the
trigonal crystal field splitting of the 5T2 ground state of octahedral Fe 2+ ions.
In addition to MnCrInS4 and InCrFeS4, Mimura et al. (1974.) prepared
InCrCoS4 and InCrNiS4. In these compounds the In ions occupy the tetrahedral
sites. Both compounds order antiferromagnetically below 36 and 2 2 K respec-
tively. The paramagnetic susceptibility follows a Curie-Weiss law in InCrNiS4
with a positive asymptotic Curie temperature of 26 K and a rather low molar
Curie constant of 1.75. The paramagnetic susceptibility of InCrCoS4 deviates
severely from a Curie-Weiss behaviour. At 4.2 K, the magnetic moment increases
in a non-linear metamagnetic way in fields up to 8 kOe.

6.9. The mixed crystals MnCr2-xVxS4

The system MnCr2-xVxS4 was investigated by Goldstein et al. (1977b). The spinel
phase exists up to x = 0.6 with lattice parameters that increase slightly with x. A
neutron diffraction study of a sample with composition MnCq.sV0.2S4 shows that
most of the V ions occupy octahedral sites. The material is ferrimagnetic with a
N6el spin configuration. A transition to a canted spin configuration, as in MnCr2S4
(section 6.2), does not take place above 1.5 K. Measurements of the magnetization
at 4.2 K in magnetic fields up to 150 kOe revealed a decrease of the spontaneous
magnetization with increasing x. In high magnetic fields the materials show a
transition from the N6el configuration to an "oblique" spin structure. The critical
726 R.P. VAN STAPELE

field for that transition decreases from 110 k O e at x = 0 to 75 k O e at x = 0.6.


From the magnetic behaviour Goldstein et al. (1977b) conclude that the V 3+ spins
couple antiferromagnetically to the Cr s+ spins.

6.10. The mixed crystals FeCr2_xFexS4

In this section we will briefly review the properties of solid solutions between
FeCr2S4 and Fe3S4. The properties of the latter c o m p o u n d will not be discussed
separately in this chapter. W e will confine ourselves to a short catalogue of
properties, referring to the introduction given by Spender et al. (1972) for a m o r e
detailed account.
Two structures of the c o m p o u n d Fe3S4 have been found in nature. One is the
mineral smythite, with a hexagonal crystal structure; the other is the mineral
greigite, which has the spinel structure (Skinner et al. 1964). The spinel com-
pound can be synthesized, but synthetic samples are often contaminated with
other iron sulphides. The cell edge of the mineral is 9.876 A (Skinner et al. 1964);
literature values
o
of the lattice p a r a m e t e r of synthetic materials vary between 9.81
and 9.90 A. Fe3S4 is ferrimagnet with the spins ordered in a simple N6el
configuration (Spender et al. 1972). U d a (1968), who has studied the c o m p o u n d
extensively, measured a Curie t e m p e r a t u r e of 580 K and a saturation m o m e n t of
1.3/xB/molecule. Spender et al. (1972) reported the values 6 0 6 K and
2.2/zB/molecule in their p a p e r on the magnetic properties and the M6ssbauer
spectra of Fe3S4. From conductivity m e a s u r e m e n t s these authors obtained in-
dications for a semimetallic behaviour.
Single-phase samples of Fel+xCr2-xS4 were prepared between x = 0 and x = 0.5
(Robbins et al. 1970b). The solid solutions crystallize in the spinel structure with a
lattice parameter, that decreases from 9.995 A at x = 0 to 9.984 A at x = 0.5. In
agreement with the high Curie t e m p e r a t u r e of Fe3S4, the Curie t e m p e r a t u r e of
Fel+xCr2_~S4 increases from 180 K at x = 0 to 302 K at x = 0.5. The magnetic
m o m e n t measured at 1.5 K on polycrystalline samples (in which the magnetization
is difficult to saturate because of the strong magnetic anisotropy) changes from
1.52/xB/molecule in FeCr2S4 to 1.71/xB/molecule at x = 0.5 with a m a x i m u m of
1.79/xB/molecule at x = 0.3 (Robbins et al. 1970b).
From the low t e m p e r a t u r e resistivity m e a s u r e m e n t s it appears that the com-
pounds are semiconductors. The Seebeck coefficient is positive, decreasing from
+80 fxV/°C in FeCr2S4 to +3 IxV/°C at x __40.4 (Robbins et al. 1970b).
A more recent investigation of the c o m p o u n d FemCrl.8S4 by means of neutron
diffraction and magnetic m e a s u r e m e n t s (Babaev et al. 1975) shows the c o m p o u n d
to be a ferrimagnet with the spins ordered in the N6el configuration. The
magnetization versus t e m p e r a t u r e curve, measured on a powder sample, has a
m a x i m u m at about 90 K. This, together with a large coercive force of 1.5 k O e at
4.2 K, indicates a strong magnetic anisotropy at low temperatures.

6.11. The mixed crystals MCr2S4-xSex with M = Mn, Fe, Co or C u m F e m

In the system MCr2S4-xSex with M = Mn, Fe or Co, solid solutions with the spinel
structure between the semiconducting ferrimagnetic spinels MCr2S4 and the
SULPHOSPINELS 727

corresponding selenides, which crystallize in the cation-defective NiAs structure,


have been prepared in a limited range of sulphur-rich compositions. This range
decreases from 0 ~< x ~<2 for MnCr2S4_xSex to 0 ~< x ~< 1.25 for FeCr2S4 xSex and
0 ~ x ~< 1 for CoCr2S4-xSex. The Curie t e m p e r a t u r e decreases slowly with increas-
ing Se content, which is attributed to a weakening of the superexchange inter-
action between the octahedral Cr 3+ ions and the tetrahedral/VI~+ = Mn 2+, Fe 2+ or
Co 2+ ions (Gibart et al. 1973).
T h e system MnCr2S4 xSex has been investigated in m o r e detail by measure-
ments of the paramagnetic susceptibility, the magnetization at 15.3kOe as a
function of t e m p e r a t u r e and the magnetization at 1.5K as a function of the
strength of the magnetic field up to 60 k O e (Robbins et al. 1973). The m a x i m u m
in the magnetization versus t e m p e r a t u r e curve, which is typical of MnCreS4
(section 6.2), was observed to have disappeared at x = 0.5 (fig. 94), while the
magnetization at 1.5 K remains a linear function of the strength of the magnetic
field. At higher Se concentrations the magnetization at 1.5 K becomes an increas-
ingly non-linear function of the magnetic field. The Curie t e m p e r a t u r e decreases
from 74 K at x = 0 to 56 K at x = 2, whereas the asymptotic Curie t e m p e r a t u r e
increases linearly from - 2 7 K for MnCr2S4 to +50 K for MnCrzSzSe2. 'The obser-
ved change of the magnetic properties is attributed to a decrease of the strength
of the negative Mn2+-Cr 3+ superexchange interaction with increasing Se concen-
tration (Robbins et al. 1973).
In the system FemCul/zCrzS4_xSex a spinel phase exists between x = 0 and
x = 2.75 with a lattice p a r a m e t e r that varies linearly with the Se concentration.
The Curie t e m p e r a t u r e decreases monotonically from 350 K for x = 0 to 280 K for
x = 2.75. Conductivity m e a s u r e m e n t s show that the compounds are semiconduc-
tors with a negative magnetoresistance effect of about 3% at the Curie tem-
perature (Gyorgy et al. 1973).

3.6
x=0.5
3.2
x=0.25
-5 2.8
2.4
~, 2.0
g 1.6
:,=
0
1.2
.N_ 0.8
c~
0.4.

E 0 40 80 120 160 200 24.0 2~,0

temperature (K)

Fig. 94. Magnetization versus temperature of MnCr2S4-xSex in a magnetic field of 15.3 k O e (Robbins et
al. 1973).
728 R.P. VAN STAPELE

7. Some rhodium and cobalt spinels

7.1. Introduction

In this section we will review some Rh and Co spinels, which are interesting
because of their contrasting properties. Although both Rh 3+ and Co 3+ are in the
zero spin t6g state and occupy octahedral sites, the properties of CoRh2S4 and
C03S4, for example, differ strikingly. As will be described later on, CoRheS4 is a
semiconducting antiferromagnet with a high N6el temperature and a lattice
parameter of 9.8 A whereas C03S4 is a metallic paramagnet with a lattice parameter
of 9.4 A. This correlates with a cell edge of C03S4 smaller than and a cell edge of
CoRh2S4 about equal to 9.8A, which is the edge of the smallest cell that can
accommodate sulphur ions with a normal radius of 1.74 A (section 1). In a tight lattice
like that of C03S4 the d electrons of the tetrahedrally coordinated cations are
apparently delocalized, which gives rise to metallic conduction and anomalous
magnetic moments.
Within the class of Rh compounds interesting phenomena were observed in the
system C01-xCuxRh2S4 and Fel-xCuxRh2S4. The properties of Fel-xCuxRh2S4 are
very similar to those of Fel-xCuxCr2S4 (section 6.6), indicating Fe 2+ levels w i t h i n
the energy gap. However, the properties of C01-xCuxRh2S4 indicate Co 2+ levels
below the top of the valence band as in Col_xCuxCr2S 4 (section 6.6).

7.2. CoRh2S4, C01-xFexRh2S4 and FeRh2S4

CoRh2S4 is a normal spinel with a lattice parameter of about 9 . 8 A (table 1)


(Blasse 1965). Since Rh in oxy- and sulphospinels always occurs as trivalent ions in
the low-spin t6g state, the Co ions are the only magnetic ions. In the reciprocal
magnetic susceptibility (fig. 95) a rather sharp minimum has been observed at
400 K, which indicates a surprisingly strong antiferromagnetic ordering of the Co
spins (Blasse 1965). Other susceptibility data in the literature show a less
pronounced maximum around 4 0 0 K (Kondo 1976), which probably can be
connected with the difficulty of preparing pure samples (Lotgering 1968b). In the
paramagnetic region the susceptibility follows a Curie-Weiss law with an asymp-
totic Curie temperature of - 4 0 0 K and a molar Curie constant of 2.3 (Blasse
1965). Without correction for the Van Vleck susceptibility of the Rh 3+ ions the
value of the Curie constant corresponds to a g value of 2.2, which compares well
with values usually found for Co 2+ on tetrahedral sites (for example, g = 2.25 for
ZnS : Co 2+ (Ham et al. 1960).
The magnetic susceptibility clearly shows that the Co ions are divalent. The
observed negative temperature coefficient of the electrical resistivity (Blasse 1965,
Kondo 1976) shows the compound to be a semiconductor, in accordance with the
valencies Co:+Rh3+S4.
The superexchange interaction between neighbouring A ions in spinel takes
place via two anions and is therefore in most cases weak. According to Blasse
(1963) non-magnetic B ions like Rh 3+ or Co 3+ possibly play a part in superex-
SULPHOSPINELS 729

600

\
500 \
\\ x:o/x/
I 400 / x--o.y
\ / /*
x~
3oc " ./'~'f /
.,'~" f\%/ ..-~qs , /
2oc " ~ "~z*~ ....*/ *"~x=zo

,oo :?i"
Oi I I I I i
200 400 600 800 1000
temperature (K)
Fig. 95. Reciprocal molar magnetic susceptibility of CoRh2-xCrxS4 as a function of temperature
(Lotgering 1968b). Curve for x = 0 after m e a s u r e m e n t s of Blasse (1965).

change interactions between neighbouring A ions. The anomalously strong anti-


ferromagnetic interaction in CoRh2S4 might be enhanced by the Rh 3+ ions.
However, the results of an analysis of the magnetic properties of the system
CoRhz_xCrxS4 (section 7.3) do not confirm this assumption (Lotgering 1968b).
It is not possible to prepare single-phase samples of FeRhzS4 (Koerts 1965,
Riedel and Karl 1979). Tressler et al. (1968) have reported that the compound has
an unidentified complex distorted spinel structure.
Attempts were made to prepare solid solutions of FeRhzS4 in CoRh2S4 (Kondo
1976). Single-phase samples could not be obtained. The main phase was found to
be spinel, the main contamination RhzS3. The magnetic susceptibility, the elec-
trical resistivity and the M6ssbauer spectra were measured and the results are
ascribed to the main phase Col_xFexRh2S4 (Kondo 1976). The compounds are
semiconducting antiferromagnets with a N6el temperature that decreases linearly
from about 400 K for CoRh2S4 to 250 K for Co0.25Fe0.75RhzS 4. The susceptibility of
the Fegh2S4 sample (more contaminated than the samples with other com-
positions) has a maximum at 20 K, decreasing monotonically at higher tem-
peratures. However, extrapolation of the linear variation with x gives a N6el
temperature of 200 K, roughly agreeing with the temperature of 190 K below
which a broad paramagnetic line in the complex M6ssbauer spectrum of the
730 R.P. VAN STAPELE

FeRh2S4 sample was observed to be split into six lines. These data are confirmed
by Spender's findings, that FeRh2S4 is a semiconducting antiferromagnet with a
N6el temperature of 205 K (Spender 1973, referred to in Boumford and Morrish
1978).
The M6ssbauer spectra of Col_xFexRh2S4 with x ~<0.75 were observed to be
superpositions of many Fe z+ spectra with different values of hyperfine fields and
quadrupole splittings (Kondo 1976).

7.3. The mixed crystals feRh2-xCrxS4, CoRh2-xfrxS4 and NiRh;-xCrxS4

In the system FeRh2-xCrxS4 relatively pure spinels were prepared in the range
0.8 ~< x ~<2 (Riede! and Karl 1979). The cell edge increases from 9.935 A at x = 0.8
to 9.998 A at x = 2. All spinels are normal with only Fe 2+ on tetrahedral sites. The
materials are p-type semiconductors. The room temperature Mrssbauer spectra of
57Fe consist of several overlapping doublets with almost identical isomer shift but
different quadrupole splittings, which are attributed to tetrahedral Fe 2+ ions with
different numbers of Rh ions as nearest octahedral site neighbours.
In the system CoRh2_xCrxS4, investigated by Lotgering (1968b), single-phase
samples with the spinel structure were prepared for all x. The magnetic properties
vary in an interesting way between the strong antiferromagnetism of CoRh2S4 and
the ferrimagnetism of CoCr2S4 (section 6.4). The paramagnetic susceptibility (fig.
95) follows a Curie-Weiss law with asymptotic Curie temperatures as given in fig.
96 and Curie constants that vary roughly linearly between the values for CoRh2S4
and CoCr2S4. From the magnetic susceptibility of CoRhlsCr0.sS4, which shows a
kink at 360 + 10 K and a ferrimagnetic Curie temperature at 50 K (fig. 95),
Lotgering (1968b) concluded that the Co spins order antiferromagnetically at

700

6O0

T 500
4001 TN ~

z 300
2
~" 200

100

o; I
X
Fig. 96. Absolute value t01 of the negative asymptotic Curie temperature, Nrel temperature TN and
ferrimagneticCurie temperature Tcof CoRh2-xCrxS4as a functionof the composition(Lotgering1968b).
SULPHOSPINELS 731

3 6 0 K and that at the lower temperature of 5 0 K a transition occurs to a


Yafet-Kittel configuration with canted Co spins opposite to the Cr spins. The
observed variation with x of the saturation magnetization (fig. 97) can be
described with a constant ratio c~ of 1.37 between the strength JAA of the negative
C o - C o superexchange interaction and the strength JAB of the negative C o - C r
superexchange interaction. In this description a configuration with canted Co
spins occurs for Cr contents up to x = 1.65. For higher Cr concentrations the
moments agree with a simple N6el configuration. Although the absence of a N6el
temperature for x > 0 . 5 and the observed variation of the asymptotic Curie
temperature and the ferrimagnetic Curie temperature cannot be explained within
a molecular field approximation, the behaviour of the saturation magnetization as
a function of x is typical of an anomalously strong C o - C o interaction in the whole
range of x. This shows that the strong C o - C o interaction is not due to the
presence of Rh 3+ ions (Lotgering 1968b).
In the system NiRh2_xCrxS4 single-phase spinels were prepared in the range
0.3 ~< x ~<0.8 (Itoh 1979). The cell edge increases slightly from 9.702 A at x = 0 to
9.715 A at x = 0.8. The samples are ferromagnetic and have a metallic electrical
conduction. The Curie temperature increases with x (fig. 98). The paramagnetic
moment is roughly proportional to the Cr concentration and is about 0.9 per
gramatom Cr. The ferromagnetic moments are shown in fig. 99. Although the
magnetic behaviour is reminiscent of that observed in the series CuRh2_xCrxSe4
with x > 1 (section 4.6), the anomalous values of the magnetic moments cannot be
reconciled with a specific valency of the Cr ions. This points to delocalized Ni

I
D
2

O
o
o

• /

I /
1 1.65 2
~x
Fig. 97. Saturation moment Ms at 4.2 K of CoRh2-xCrxS4.The straight lines have been calculated with
a Co moment of 3.6/zB and a Cr moment of 3/x~ for: (a) a triangular spin configuration with
a = J A A / J A B = 1.37 (Ms=3(1-a-1)x) and (b) a simple N6el configuration (Ms = 3x-3.6). After
Lotgering (1968b).
732 R.P. VAN STAPELE

100

8O o

B 60 o
o

(9
40
E o
o

I 20
' o'2 ' o'.~ ' o's ' o18 '110
X
Fig. 98. Asymptotic ((2)) and ferromagnetic (O) Curie temperature of NiRh2_xCrxS4(Itoh 1979).

1.0/,

:z,
081
Q

.~_ 0.E
.N_
"$
c-
0.4

E 0.2

l o
I I
o12 o.~ o;
I I I
0'.8
I
1'.o
-----.,. X
Fig. 99. Magnetization at 4.2 K and 14.5 kOe of NiRhz-xCrxS4 (ltoh 1979).

electrons that are partially spin-polarized in a direction opposite to that of the


ferromagnetically coupled Cr spins (Itoh 1979).

7.4. The mixed crystals Fel_xCuxRh2S4 and C01-xCuxRh2S4

Inspired by the 1:1 ionic ordering on the tetrahedral sites in Fet/2Cul/2Cr2S4,


Plumier and Lotgering (1970) investigated the properties of the c o m p o u n d
FemCUl/2Rh2S4. The c o m p o u n d has the spinel structure with a = 9 . 8 5 A and
u = 0.3815. X-ray diffraction clearly showed a 1 : 1 ordering of Fe 3+ and Cu 1÷ ions
on the tetrahedral sites. The material is semiconducting with a large Seebeck
coefficient, the sign depending on the details of the preparation.
The most remarkable observed property of FemCUl/2Rh2S4 is its strong antifer-
romagnetism. A b o v e 170 K, the magnetic susceptibility follows a Curie-Weiss law
with an asymptotic Curie t e m p e r a t u r e 0 = - 4 2 0 K and a molar Curie constant
Cm = 2.32. T h e magnetic susceptibility contains a rather large temperature-in-
SULPHOSPINELS 733

dependent part, due to the diamagnetic susceptibility and the Van Vleck suscep-
tibility of the Rh 3+ ions. Correction for these terms gives 0 = - 3 6 7 K and
Cm = 2.09 (Boumford and Morrish 1978). The value of Cm agrees fairly well with
the value 2.19 for trivalent iron ions. The susceptibility shows a maximum at
145 K below which the iron spins order antiferromagnetically. Neutron diffraction
experiments revealed that the antiferromagnetic ordering is of the second kind, as
in MnO. In this type of ordering the second-neighbour interaction dominates the
nearest-neighbour interaction, which leads in the case of Feu2CuuzRh2S 4 to the
conclusion that a surprisinglYoStrong superexchange interaction exists between
Fe 3+ ions at a distance of 9.85 A.
A study of the 63'65Cu nuclear magnetic resonance as a function of temperature
in the paramagnetic state indicates a N6el temperature of 135 K, slightly below
the value of 145 K mentioned above (fig. 100). The Fe 3+ ions were observed to
give a negative transferred hyperfine field at the Cu nucleus, which would amount
to -12.5 kOe for a saturated Fe 3+ magnetization (Locher and Van Stapele 1970).
The semiconduction and the valencies Fel/zCul/zRh2S4
3+ ~+ clearly indicate that the
Fe 2+ levels fall in the energy gap between the valence and the conduction band as
in Fel-xCuxCrzS4 (section 6.6 and fig. 91 with Rh 3+ instead of Cr3+). This was
recently confirmed by a study of the properties of Fel_xCuxRh2S4 with X/>0.06
(Boumford and Morrish 1978). Using X-ray diffraction these authors observed an
undistorted spinel structure with a linearly varying lattice parameter throughout
the series. Ordering of copper and iron ions was detected in the range 0.3 ~< x ~<
0.54. Magnetic measurements indicate antiferromagnetic behaviour for all com-
positions. The N6el temperature, as determined from M6ssbauer spectra, and the

200

l 100

I
o!2 ' o',~ ' o:~ o:a 1.0
-200 X
G) ×

X
-300
×

• x
X

-4.00 X

Fig. 100. Fel-xCuxRh2S4. N6el temperature (T~) (from M6ssbauer spectra) and asymptotic Curie
temperature (0) before (O) and after correction for the Rh 3+ Van Vleck susceptibility (x), according to
Boumford and Moorish (1978); ([~) data of Plumier and Lotgering (1970), (O) data of M.R. Spender,
Ph.D. thesis (University of Manitoba, 1973, unpublished).
734 R.P. VAN STAPELE

asymptotic Curie temperatures are given in fig. 100. Iron-rich compounds with
0.06 ~< x ~< 0.5 exhibit remanence and displaced hysteresis loops after cooling to
4 K in an external magnetic field of 18kOe. Both the isomer-shift in the
M6ssbauer spectrum and the paramagnetic m o m e n t s indicate the presence of
solely Fe 3+ for x/> 0.5 and a gradual change from Fe 3+ to Fe 2+ for x decreasing
from 0.4 to 0.06. The observed behaviour agrees with the valencies
2+ 3+ +
Fel-2xFex CuxRh2S4 in the range 0 ~ x ~< 0.5 and Fel3 + xCuxRh2S4
+
for x I> 0.5.
In contrast to the behaviour of Fel_xCuxRh2S4 the properties of Col_xCuxRh2S4
do not indicate Co 2+ levels in the energy gap. As has been mentioned in section
6.6, this is not the case either in Col_xCuxCr2S4, in which system indications were
found for Co 2+ levels below the top of the valence band. However, particularly in
the system COl-xCuxRh2S4, this position gives rise to remarkable properties, i.e.
antiferromagnetism for 0 < x ~< 0.4, spontaneous magnetization for 0.4 ~< x ~< 0.7
and paramagnetism for 0.7 ~< x ~< 1.
Single-phase preparations of Col_~CuxRhaS4 were prepared for 0.1~<x ~< 1
(Lotgering 1969). X-ray diffraction, which shows the presence of a spinel phase
with a lattice p a r a m e t e r of 9.78 A, cannot establish the formation of mixed
crystals, because of the equal values of the lattice parameters of CoRh2S4 and
CuRh2S4. However, the change in physical properties proves the existence of solid
solutions. The samples have a nearly temperature-independent, low resistivity
and a positive Seebeck coefficient, decreasing from 125 ixV/deg for x = 0.1 to
25 txV/deg for the Cu-rich compositions with x t>0.5. Measurements of the
magnetic susceptibility show that the Co-rich compositions with x ~<0.4 are
antiferromagnetic with a N6el t e m p e r a t u r e that decreases with increasing x (fig.
101). In the range 0 . 4 4 x ~<0.7 the samples are ferromagnetic with Curie tem-
peratures that link up with the N6el temperatures (fig. 101). The small spon-

_~2,~
&OOt
2 3OO

E \

200 \

100 Tc

0 I I i
0 0.2 0.4 0.6 0.8
. ~ X
Fig. 101. N6el (TN) or Curie temperature (To) of Col xCuxRh2S4, according to Lotgering (1969)
(measured on quenched (ff3)or on annealed samples (mE)).
SULPHOSPINELS 735

I i I

0.8

o
E
0.6
:3,

G
O

0.4
g)
c-
(31
O
E
0.2 !
! I
i I
!
I
0~ 02 0.4 0.6 0.8
X
Fig. 102. Magnetization at 4.5 K in 20 kOe of Col-xCuxRh2S4 after Lotgering (1969) (measured on
quenched (O) or on annealed samples (0)).

t a n e o u s magnetization reaches a m a x i m u m of 1.6/XB/CO ion at x = 0.5 (fig. 102).


C o m p o s i t i o n s with x ~ 0 . 7 are paramagnetic. A definite explanation of these
properties was not given. It was, however, n o t e d that three ranges of compositions
can be distinguished, if o n e tentatively assumes that the valence b a n d can
a c c o m m o d a t e 30 < 1 holes a b o v e the C o 2+ levels:

(i) 0 ~< x ~< 60, in which there are x holes in the valence hand, while all the C o ions
are divalent; symbolized by the f o r m u l a

2+ + 2- -
Col_xCuxeh2{S4_xSx} ;

(ii) 30 ~< x ~< (1 + 60)/2, in which there are (x - 30) holes in the C o 2+ levels (i.e.
C o 3+ ions) and 80 holes in the valence band, symbolized by the f o r m u l a

{COl+6o_2xCox_6o}Cux
2+ 3+ +
Rh2
3+ 2-
{S 4 3oS6o}; -

(iii) ( 1 + 30)/2 ~ x ~< 1, in which range all C o ions are trivalent, while there are
(2x - 1) holes in the valence band, symbolized by

3+ +
COl-xCux Rh23 + { S 5 -22-x S 2 x - 1- } .

With 30 = 0.4, the first range possibly c o r r e s p o n d s to the antiferromagnetic range,


the s e c o n d to the f e r r o m a g n e t i c range and the third to the p a r a m a g n e t i c range
(Lotgering 1969).
In Col/2Cul/2Rh2S4, X-ray absorption discontinuities were measured. T h e chem-
ical shift of the c o p p e r and cobalt K-discontinuities indicate Cu + and C o 2+ ions
736 R.P. VAN STAPELE

(Ballal and Mande 1977) and do not provide evidence for a higher valency of a
part of the Co ions.

ZS. Co3S4 and NiCo2S4

Co3S4, which occurs in nature as the mineral linneite, has the spinel structure o
(Menzer 1926, de Jong and Willems 1927). The lattice parameter is small (9.4 A),
like the other sulphospinels with low spin Co 3+ ions on the octahedral sites (table
1). This combines with a metallic conduction (Bouchard et al. 1965). The com-
pound is paramagnetic down to 20 K (Lotgering 1956) with a magnetic suscep-
tibility that is very sensitive to the purity of the sample. Literature data vary
(Locher 1968) between temperature-independent susceptibilities of 3.9×
10-4cm3/mol (Serres 1953) and 10.4x 10-4cm3/mO1 (Heidelberg et al. 1966) to
susceptibilities that are weakly temperature-dependent and at room temperature
have a value of about 39 x 10-4cm3/mol (Lotgering 1956) or 11x 10-4cm3/mol
(Lotgering (unpublished), quoted by Locher 1968).
The small susceptibility and the metallic conduction are difficult to reconcile
with paramagnetic Co 2+ ions on the tetrahedral sites and indicate a delocalization
of the 3d electrons of the tetrahedral Co ions in a 3d band (Goodenough 1969).
In the nuclear magnetic resonance spectrum of 59Co a strong, single symmetric
line without quadrupole effects is attributed to tetrahedrally coordinated Co ions
(Locher 1968, Saji and Yamadaya 1972, Locher 1973). In the spectrum due to
octahedrally coordinated Co ions a considerable nuclear quadrupole interaction
was observed, almost equal to the quadrupole interaction in the very similar
spectrum of CuCo2S4 (section 4.4). This led Locher (1968) to suggest a similar
charge distribution in the two lattices, which would mean effectively monovalent
Co ions in Co3S4.
NiCo2S4 is a normal spinel (Lotgering 1956) with a lattice parameter of 9.4
(table 1). The compound is paramagnetic with a low susceptibility, which does not
agree with Ni 2+ ions on the tetrahedral sites, (Lotgering 1956), and has a metallic
conduction (Bouchard et al. 1965). Other properties have not been measured, but
the metallic conduction and the anomalous susceptibility point to delocalized Ni
3d electrons.

Acknowledgements

I am grateful to F.K. Lotgering for helpful discussions and critical reading of the
text, to T.J.A. Popma and M.H. van Maaren for their comments and to S.
Heymans for his help with the bibliography.
SULPHOSPINELS 737

Notes added in proof

(a) The electronic structure of transition metal sulphospinels was recently discussed by Haas (Haas, C.,
1980, Jpn. J. Appl. Phys. 19, suppl. 19-3, 171) (see p. 616).
(b) R.M. Fleming, F.J. DiSalvo, R.J. Cava and J.V. Waszczak (1981), Phys. Rev. B24, 2850) observed
charge-density-wave transitions at 90, 75 and 50 K in the resistance, magnetic susceptibility, and by X-ray
diffraction (see p. 624).
(c) More recently less clear data were obtained in the series CuCrz-xRhxS4 with 1.2 ~<x ~<2 (Itoh, H.,
1980, J. Phys. Soc. Jpn. 48, 1130) (see p. 638).
(d) Calculations of the electronic band structure were reported for CdCr2S4, CdCr2Se4 and HgCrzSe4.
(Kambara, T., T. Oguchi and K.I. Gondaira, 1980, J. Phys. C: Solid State Phys. 13, 1493; Oguchi, T., T.
Kambara and K.I. Gondaira, 1980, Phys. Rev. B22, 872 and B24, 3441 (1981)). The results of a
photoemission study of CdCrzS4 and CdCr2Se4 were reported by W.J. Miniscalco, B.C. McCollum, N.G.
Stoffel and G. Margaritondo, 1982, Phys. Rev. B25, 2947 (see p. 653).
(e) CdCri.6In0.4S4was observed to show a typical spin-glass behaviour (Fiorani, D., M. Nogues and S.
Viticoli, 1982, Solid State Commun. 41, 537) (see p. 656).
(f) See also M.N. Iliev and G. Giintherodt (1980, Phys. Status Solidi, B98, K9) (see p. 664).
(g) Recently a magnetoelectric effect in ZnCr2Se4 was observed and discussed by K. Siratori and E. Kita
(1980, J. Phys. Soc. Jpn. 48, 1443) (see p. 672).
(h) A. Selmi, R. le Toullec and P. Gibart (1980, Solid State Commun. 33, 889) reported on the plasmon
reftectivity of HgCr~Se4 (see p. 693).
(i) A neutron diffraction study and magnetic measurements in low fields showed that Zn0.3Hgi.TCr2Se4
has a spin-glass magnetic structure (Sadykov, R.A., A.V. Filatov, P.L. Gruzin, V.M. Novotortsev, I.S.
Kovaleva and V.A. Levshin, 1980, JETP. Lett. 31, 642) (see p. 696).
(j) Piezoelectricity of Cu0.sIn0.sfr2S4 was reported by N . A . Tsvetkova, K.P. Belov, L.I. Koroleva, V.V.
Titov, Ya.A. Kesler and I.V. Gordeav (1979, JETP. Lett. 30, 533) (see p. 698).
(k) See for more recent measurements in high magnetic fields and a further discussion R. Plumier, M.
Sougi, M. Lec6mte and A. Miedan-Gros (1980, Z. Phys. B - Condensed Matter, 40, 227) (see p. 700).
(1) Also in the system Cul-xFexCr2Se4 Cu was found to be replaced by Fe 3+ ions in the spinel phase,
which occurred for 0 ~<x ~<0.6 (Hang Nam Ok, Yun Chung and Jung Gi Kim, 1979, Phys. Rev. B20, 4550)
(see p. 718).
(m) An X-ray photoelectron spectroscopy study of powder samples of Fel-xCuxCr2S4 with 0 < x ~< 1
confirmed the monovalent state of Cu ions throughout the series (Ando, K., 1980, Solid State Commun.
36, 165) (see p. 719).

References

Aers, G.C., A.D. Boardman and E.D. Isaac, and T. Watanabe, 1978, J. Phys. Soc. Japan,
1975, Phys. Lett. 54A, 373. 44, 172.
Ahrenkiel, R.K. and T.J. Coburn, 1973, Appl. Akino, T. and K. Motizuki, 1971, J. Phys. Soc.
Phys. Lett. 22, 340. Japan, 31, 691.
Ahrenkiel, R.K., F. Moser, S. Lyu and C.R. Albers, W. and C. Haas, 1964, Phys. Lett. 8,
Pidgeon, 1971, J. Appl. Phys. 42, 1452. 300.
Ahrenkiel, R.K., T.H. Lee, S.L. Lyu and F. Albers, W. and C.J.M. Rooymans, 1965, Solid
Moser, 1973, Solid State Commun. 12, State Commun. 3, 417.
1113. Albers, W., G. van Aller and C. Haas, 1965,
Ahrenkiel, R.K., T.J. Coburn and E. Carnall, Coll. Int. du C.N.R.S. sur les drriv6s semi-
1974, IEEE Trans. on Magn. 10, 2. mrtalliques, Orsay 1965 (Editions du Centre
Ahrenkiel, R.K., S.L. Lyu and T.J. Coburn, National de la Recherche Scientifique, Paris,
1975, J. Appl. Phys. 46, 894. 1967), p.. 19.
Akerstrom, S., 1959, Arkiv Kemi, 14, 403. Alexander, S., J.S. Helman and I. Balberg,
Akimitsu, J., K. Siratori, G. Shirane, M. Iizumi 1976, Phys. Rev. B13, 304.
738 R.P. VAN STAPELE

Allain, Y., F. Varret and A. Mi6dan-Gros, Barraclough, K.G., W. Lugscheider, A. Meyer,


1965, C.R. Acad. Sc. Paris 260, 4677. H. Schaefer and L. Treitinger, 1974, Phys.
Amith, A. and S.B. Berger, 1971, J. Appl. Phys. Status Solidi, A22, 401.
42, 1472. Batlogg, B., M. Zvfira and P. Wachter, 1978,
Amith, A. and L. Friedman, 1970, Phys. Rev. Solid State Commun. 28, 567.
B2, 434. Belov, K.P., Y.D. Tret'yakov, I.V. Gordeev,
Amith, A. and G.L. Gunsalus, 1969, J. Appl. L.I. Koroleva, A.V. Ped'Ko, E.I. Smirnov-
Phys. 40, 1020. skaya, V.A. Alferov and Y.G. Saksonov,
Anderson, P.W., 1950, Phys. Rev. 79, 350. 1973, Sov. Phys. Solid State, 14, 1862.
Anzina, L.V., V.G. Veselago and S.G. Rudov, Belov, K.P., L.I. Koroleva, M.A. Shalimova and
1976, JETP Lett. 23, 474. S.D. Batorova, 1975, Sov. Phys. Solid State, 17,
Arai, T., K. Wakamura and K. Kudo, 1971, J. 197.
Phys. Soc. Japan, 30, 1762. Benoit, R., 1955, J. Chem. Phys. 52, 119.
Arai, K.I., O. Kubo, N. Tsuya, F. Okamoto and Berger, S.B. and A. Amith, 1971, J. de Phys.,
P.K. Baltzer, 1972, IEEE Trans. on Magn. 8, Coll. C.1.32, 934.
479. Berger, S.B. and L. Ekstrom, 1969, Phys. Rev.
Arai, T., M. Wakaki, S. Onari, K. Kudo, T. Lett. 23, 1499.
Satoh and T. Tsushima, 1973, J. Phys. Soc. Berger, S.B. and H.L. Pinch, 1967, J. Appl.
Japan, 34, 68. Phys. 38, 949.
Babaev, G.Y., A.G. Kocharov, K. Ptasevich, Berger, S.B., J.I. Budnick and T.J. Burch,
LI. Yamzin, M.A. Vinnik, Y.G. Saksonov, 1968a, J. Appl. Phys. 39, 658.
V.A. Alferov, I.V. Gordeev and Y.D. Berger, S.B., J.I. Budnick and T.J. Burch,
Tret'yakov, 1975, Sov. Phys. Crystallogr. 20, 1968b, Phys. Lett. 26A, 450.
336. Berger, S.B., J.I. Budnick and T.J. Burch,
Bairamov, A.I., A.G. Gurevich, V.I. Kar- 1969a, Phys. Rev. 179, 272.
povich, V.T. Kalinnikov, T.G. Aminov and Berger, S.B., J.I. Budnick and T.J. Burch,
L.M. Emiryan, 1976, Sov. Phys. Solid State, 1969b, J. Appl. Phys. 40, 1022.
18, 396. Berger, S.B., T.J. Burch, J.L Budnick and L.
Bairamov, A.I., A.G. Gurevich, L.M. Emiryan Darcy, 1971, J. Appl. Phys. 42, 1309.
and N.N. Parfenova, 1977, Phys. Lett. 62A, Bindloss, W., 1971, J. Appl. Phys. 42, 1474.
242. Blasse, G , 1963, Philips Res. Rep. 18, 383.
Balberg, I. and J.S. Helman, 1978, Phys. Rev. Blasse, G., 1965, Phys. Lett. 19, 110.
B18, 303. Blasse, G. and D.J. Schipper, 1964, J. Inorg.
Balberg, I. and A. Maman, 1977, Phys. Rev. Nucl. Chem. 26, 1467.
B16, 4535. Bongers, P.F., 1957, Dissertation (University of
Balberg, I. and H.L. Pinch, 1972, Phys. Rev. Leiden), unpublished.
Lett. 28, 909. Bongers, P.F. and E.R. van Meurs, 1967, J.
Ballal, M.M. and C. Mande, 1976, Solid State Appl. Phys. 38, 944.
Commun. 19, 325. Bongers, P.F. and G. Zanmarchi, 1968, Solid
Ballal, M.M. and C. Mande, 1977, J. Phys. State Commun. 6, 291.
Chem. Solids, 38, 843. Bongers, P.F., C. Haas, A.M.J.G. van Run and
Baltensperger, W., 1970, J. Appl. Phys. 41, G. Zanmarchi, 1969, J. Appl. Phys. 40, 958.
1052. Bouchard, R.J., 1967, Mat. Res. Bull. 2, 459.
Baltensperger, W. and J.S. Helman, 1968, Helv. Bouchard, R.J., P.A. Russo and A. Wold, 1965,
Phys. Acta, 41, 668. Inorg. Chem. 4, 685.
Baltzer, P.K., H.W. Lehmann and M. Robbins, Boumford, C. and A.H. Morrish, 1978, Phys.
1965, Phys. Rev. Lett. 15, 493. Rev. B17, 1323.
Baltzer, P.K., P.J. Wojtowicz, M. Robbins and Bouwma, J. and C. Haas, 1973, Phys. Status
E. Lopatin, 1966, Phys. Rev. 151, 367. Solidi, B56, 299.
Baltzer, P.K., M. Robbins and P.J. Wojtowicz, Broquetas-Colominas, C., R. Ballestracci and
1967, J. Appl. Phys. 38, 953. G. Roult, 1964, J. de Phys. 25, 526.
Banus, M.D. and M.C. Lavine, 1969, J. Solid Brossard, L., L. Goldstein and M. Gu~ttard,
State Chem. 1, 109. 1976, J. de Phys., Coll. C6, 37, 493.
SULPHOSPINELS 739

Brossard, L., J.L. Dormann, L. Goldstein, P. Enokiya, H., M. 5¢amaguchi and T. Hihara,
Gibart and P. Renaudin, 1979, Phys. Rev. 1977, J. Phys. Soc. Japan, 42, 805.
B20, 2933. Feiner, L.F., 1977, Electron-Phonon Inter-
Brfiesch, P. and F. D'Ambrogio, 1972, Phys. actions and Phase Transitions, ed., T. Riste
Status Solidi, BS0, 513. (Plenum: New York) p. 345.
Br/iesch, P., H. Kalbfleisch and F. Lehmann, Feiner, L.F., 1982, J. Phys. C: Solid State Phys.
1971, Phys. Status Solidi, B46, K99. 15, 1515.
Busch, G., B. Magyar and P. Wachter, 1966, Feldtkeller, E. and L. Treitinger, 1973, Int. J.
Phys. Lett. 23, 438. Magn. 5, 237.
Busch, G., B. Magyar and O. Vogt, 1969, Solid Ferreira, J.M.C. and M.D. Coutinho-Filho,
State Commun. 7, 509. 1978, Solid State Commun. 28, 775.
Callen, E., 1968, Phys. Rev. Lett. 20, 1045. Flahaut, J., L. Domange, M. Guittard and S.
Campbell, J.S. and J.R. Davis, 1939, Phys. Rev. Fahrat, 1961, C.R. Acad. Sc. Paris 253, 1956.
55, 1125. Frankel, R.B., J.J. Huntzicker, D.A. Shirley
Cfipek, V., 1977, Phys. Status Solidi, B81, 571. and N.J. Stone, 1968, Phys. Lett. 26A, 452.
Carnall, E., D. Pearlman, T.J. Coburn, F. Fujii, H., T. Kamigaichi and T. Okamoto, 1973,
Moser and T.W. Martin, 1972, Mat. Res. J. Phys. Soc. Japan, 34, 1689.
Bull. 7, 1361. Fujii, H., T. Kamigaichi, Y. Hidaka and T.
Coburn, T.J., D. Pearlman, E. Carnall, F. Okamoto, 1970, J. Phys. Soc. Japan, 29, 244.
Moser, T.H. Lee, S.L. Lyu and T.W. Martin, Fujita, H., Y. Okada and F. Okamoto, 1971, J.
1972, A.I.P. Conf. Proc. 10, 740. Phys. Soc. Japan, 31, 610.
Coburn, T.J., R.K. Ahrenkiel, E. Carnall and Geschwind, S., 1967, Hyperline Interactions
D. Pearlman, 1973, A.I.P. Conf. Proc. 18, (eds. A.J. Freeman and R.B. Frankel)
1118. (Academic Press, New York-London, 1967).
Colominas, C., 1967, Phys. Rev. 153, 558. Gibart, P., J.L. Dormann and Y. Pellerin,
Cossee, P., 1958, J. Inorg. Nucl. Chem. 8, 1969, Phys. Status Solidi, 36, 187.
483. Gibart, P., L. Goldstein and L. Brossard, 1976,
Coutinho-Filho, M.D. and I. Balberg, 1979, J. J. Magn. Magn. Mat. 3, 109.
Appl. Phys. 50, 1920. Gibart, P., M. Robbins and V.G. Lambrecht,
Darcy, L., P.K. Baltzer and E. Lopatin, 1968, J. 1973, J. Phys. Chem. Solids, 34, 1363.
Appl. Phys. 39, 898. G6bel, H., 1976, J. Magn. Magn. Mat. 3, 143.
Dawes, P.P. and N.W. Grimes, 1975, Solid G6bel, H., H. Pink, L. Treitinger and W.K.
State Commun. 16, 139. Unger, 1975, Mat. Res. Bull. 10, 783.
De Jong, W.F. and A. Hoog, 1928, Z. Krist. 66, G6bel, H., L. Treitinger, H. Pink, W.K. Unger
168. and E. Bayer, 1974, Proc. XIIth Int. Conf. on
De Jong, W.F. and H.W.V. Willems, 1927, Z. the Physics of Semiconductors, ed., M.H. Pil-
anorg, allg. Chem. 161, 311. kuhn (Teubner, Stuttgart 1974) p. 909.
Denis, J., Y. Allain and R. Plumier, 1969, C.R. Goldstein, L. and P. Gibart, 1969, C.R. Acad.
Acad. Sc. Paris B269, 740. Sc. Paris, B269, 471.
Denis, J., Y. Allain and R. Plumier, 1970, J. Goldstein, L. and P. Gibart, 1971, A.I.P. Conf.
Appl. Phys. 41, 1091. Proc. 5, 883.
Dwight, K. and N. Menyuk, 1967, Phys. Rev. Goldstein, L., D.H. Lyons and P. Gibart, 1973,
163, 435. Solid State Commun. 13, 1503.
Dwight, K. and N. Menyuk, 1968, J. Appl. Goldstein, L., L. Brossard, M. Guittard
Phys. 39, 660. and J.L. Dormann, 1977a, Physica, 86-88B,
Eagles, D.M., 1978, J. Phys. Chem. Solids, 39, 889.
1243. Goldstein, L., P. Gibart, M. Mejai and M.
Eastman, D.E. and M.W. Shafer, 1967, J. Appl. Perrin, 1977b, Physica, 86~88B, 893.
Phys. 38, 4761. Goldstein, L., P. Gibart and A. Selmi, 1978, J.
Eibschtitz, M., E. Hermon and S. Shtrikman, Appl. Phys. 49, 1474.
1967a, Solid State Commun. 5, 529. Golik, L.L., S.M. Grigorovich, Z.E. Kunikova,
Eibsch/itz, M., S. Shtrikman and Y. Tenen- Y.M. Ukrainskii and N.M. Shtykov, 1976,
baum, 1967b, Phys. Lett. 24A, 563. Sov. Phys. Solid State, 17, 1420.
740 R.P. VAN STAPELE

Goodenough, J.B., 1960, Phys. Rev. 117, 1442. Harbeke, G. and E.F. Steigmeier, 1968, Solid
Goodenough, J.B., 1965, Coll. Int. du C.N.R.S. State Commun. 6, 747.
sur les d6riv6s semi-m&alliques, Orsay 1965 Harbeke, G., S.B. Berger and F.P. Emmeneg-
(Editions du Centre National de la Recherche ger, 1968, Solid State Commun. 6, 553.
Scientifique, Paris, 1967) p. 263. Hastings, J.M. and L.M. Corliss, 1968a, J. Appl.
Goodenough, J.B., 1967, Solid State Commun. Phys. 39, 632.
5, 577. Hastings, J.M. and L.M. Corliss, 1968b, J. Phys.
Goodenough, J.B., 1969, J. Phys. Chem. Solids, Chem. Solids, 29, 9.
30, 261. Heidelberg, R.F., A.H. Luxem, S. Talhouk and
Goodenough, J.B., 1972, J. Solid State Chem. J.J. Banewicz, 1966, Inorg. Chem. 5, 194.
4, 292. Helman, J.S., I. Balberg and S. Alexander,
Gorter, E.W., 1954, Philips Res. Rep. 9, 295. 1975, A.I.P. Conf. Proc. 29, 495.
Granot, J., 1973, Phys. Lett. 43A, 269. Henning, J.C.M., 1980, Phys. Rev. B21, 4983.
Grochulski, T. and M. Gutowski, 1975, Phys. Hlidek, P., I. Barvik, V. Prosser, M. Vaneck
Status Solidi, B72, K23. and M. Zvfira, 1976, Phys. Status Solidi, B75,
Gurevich, A.G., J.M. Jakovlev, V.I. Karpovich, K45.
A.N. Ageev and E.V. Rubalskaja, 1972, Hlidek, P., M. Zvfira and V. Prosser, 1977,
Phys. Lett. 40A, 69. Phys. Status Solidi, B84, Kl19.
Gurevich, A.G., V.I. Karpovich, E.V. Rubal- Hoekstra, B., 1973, Proc. Int. Conf. on Mag-
skaja, A.I. Bairamov, B.L. Lapovok and netism, Moscow 1973, p. 117.
L.M. Emiryan, 1975, Phys. Status Solidi, B69, Hoekstra, B., 1974, Phys. Status Solidi, B63,
731. K7.
Gyorgy, E.M., M. Robbins, P. Gibart, W.A. Hoekstra, B. and R.P. van Stapele, 1973, Phys.
Reed and F.J. Schnettler, 1973, A.I.P. Conf. Status Solidi, B55, 607.
Proc. 10, 1148. Hoekstra, B., R.P. van Stapele and A.B.
Haacke, G. and L.H. Beegle, 1966, Phys. Rev. Voermans, 1972, Phys. Rev. B6, 2762.
Lett. 17, 427. Holland, W.E. and H.A. Brown, 1972, Phys.
Haacke, G. and L.C. Beegle, 1967, J. Phys. Status Solidi, A10, 249.
Chem. Solids, 28, 1699. Hollander, J.C.T., G. Sawatzky and C. Haas,
Haacke, G. and L.C. Beegle, 1968, J. Appl. 1974, Solid State Commun. 15, 747.
Phys. 39, 656. Hoy, G.R. and S. Chandra, 1967, J. Chem.
Haacke, G. and A J . Nozik, 1968, Solid State Phys. 47, 961.
Commun. 6, 363. Hoy, G.R. and K.P. Singh, 1968, Phys. Rev.
Haas, C., 1968, Phys. Rev. 168, 531. 172, 514.
Haas, C., 1970, Crit. Rev. Solid State Sci. 1, 47. Iglesias, J.E. and H. Steinfink, 1973, J. Solid
Haas, C., A.MJ.G. van Run, P.F. Bongers and State Chem. 6, 119.
W. Albers, 1967, Solid State Commun. 5, Iliev, M. and H. Pink, 1979, Phys. Status Solidi,
657. B93, 799.
Hahn, H., 1951, Z. anorg, allg. Chem. 264, 184. Iliev, M., G. Guentherodt and H. Pink, 1978a,
Hahn, H. and B. Harder, 1956, Z. anorg, allg. Solid State Commun. 27, 863.
Chem. 288, 257. Iliev, M.N., E. Anastassakis and T. Arai, 1978b,
Hahn, H. and W. Klingler, 1950, Z. anorg, allg. Phys. Status Solidi, B86, 717,
Chem. 263, 177. Itoh, H., 1979, J. Phys. Soc. Japan, 46, 1127.
Hahn, H. and K.F. Schr6der, 1952, Z. anorg. Itoh, T., N. Miyata and S. Narita, 1973, Japan J.
allg. Chem. 269, 135. Appl. Phys. 12, 1265.
Hahn, H., C. de Lorent and B. Harder, 1956, Z. Itoh, H., K. Motida and S. Miyahara, 1977, J.
anorg, allg. Chem. 283, 138. Phys. Soc. Japan, 43, 854.
Ham, F.S., G.W. Ludwig, G.D. Watkins and Jonker, G.H. and J.H. van Santen, 1950, Phy-
H.H. Woodbury, 1960, Phys. Rev. Lett. 5, sica, 16, 337 and 16, 599.
468. Kaczmarska, K. and A. Chelkowski, 1977,
Harbeke, G. and H.W. Lehmann, 1970, Solid Phys. Status Solidi, B81, K95.
State Commun. 8, 1281. Kamata, N., S. Yamazaki, S. Kabashima, T.
Harbeke, G. and H. Pinch, 1966, Phys. Rev. Hattanda and T. Kawakubo, 1972, Solid
Lett. 17, 1090. State Commun. 10, 905.
SULPHOSPINELS 741

Kambara, T. and Y. Tanabe, 1970, J. Phys. Soc. Larsen, P.K., 1973, Proc. Int. Conf. on Mag-
Japan, 28, 628. netism, Moscow 1973, g, 484.
Kamigaki, K., T. Kaneko, H. Yoshida, H. Ido Larsen, P.K. and A.B. Voermans, 1973, J. Phys.
and S. Miura, 1970, Ferrites, Proc. Int. Conf. Chem. Solids, 34, 645.
Japan, p. 614. Larsen, P.K. and S. Wittekoek, 1972, Phys.
Kanamori, J., 1959, Phys. Chem. Solids, 10, Rev. Lett. 29, 1597.
87. Larson, G.H. and A.W. Sleight, 1968, Phys.
Kanomata, T., H. Ido and T. Kaneko, 1970, J. Lett. 28A, 203.
Phys. Soc. Japan, 29, 332. LeCraw, R.C., H. von Philipsborn and M.D.
Kasuya, T. and A. Yanase, 1968, Rev. Mod. Sturge, 1967, J. Appl. Phys. 38, 965.
Phys. 40, 684. Le Dang Khoi, 1966, C.R. Acad. Sc. Paris,
Kawanishi, S., A. Tasaki and K. Siratori, 1978, B262, 1555.
J. Phys. Soe. Japan, 45, 80. Le Dang Khoi, 1968, Solid State Commun. 6,
Kawanishi, S., A. Tasaki and K. Siratori, 1979, 203.
J. Phys. Soc. Japan, 47, 1086. Lee, T.H., 1971, J. Appl. Phys. 42, 1441.
Kelly, F.M. and J.B. Sutherland, 1956, Can. J. Lee, T.H., T. Coburn and R. Gluck, 1971, Solid
Phys. 34, 521. State Commun. 9, 1821.
Kittel, C., 1976, Introd. Solid State Physics Lee, T.H., R.M. Gluck, R.K. Ahrenkiel and
(John Wiley, New York). T.J. Coburn, 1973, A.I.P. Conf. Proc. 10, 274.
Kleinberger, R. and R. de Kouchkovsky, 1966, Lehmann, H.W., 1967, Phys. Rev. 163, 488.
C.R. Acad. Sc. Paris, B262, 628. Lehmann, H.W. and F.P. Emmenegger, 1969,
Knop, O., K.I.G. Reid, Sutarno and Y. Nak- Solid State Commun. 7, 965.
agawa, 1968, Can. J. Chem. 46, 3463. Lehmann, H.W. and G. Harbeke, 1967, J.
Koerts, K., 1963, Rec. Trav. Chim. Pays-Bas, Appl. Phys. 38, 946.
82, 1099. Lehmann, H.W. and G. Harbeke, 1970, Phys.
Koerts, K., 1965, Dissertation (University of Rev. B1, 319.
Leiden, unpublished). Lehmann, H.W. and M. Robbins, 1966, J.
Kondo, H., 1976, J. Phys. Soc. Japan, 41, 1247. Appl. Phys. 37, 1389.
Koshizuka, N., Y. Yokoyama and T. Tsushima, Lehmann, H.W., G. Harbeke and H. Pinch,
1976, Solid State Commun. 18, 1333. 1971, J. de Phys. Coll. C1, 32, 932.
Koshizuka, N., Y. Yokoyama and T. Tsushima, Lems, W., P.J. Rijnierse, P.F. Bongers and U.
1977a, Physica, 89B, 214. Enz, 1968, Phys. Rev. Lett. 21, 1643.
Koshizuka, N., Y. Yokoyama and T. Tsushima, Le Nagard, N., G. Collin and O. Gorochov,
1977b, Solid State Commun. 23, 967. 1975, Mat. Res. Bull. 10, 1279.
Koshizuka, N., Y. Yokoyama, T. Okuda and T. Le Nagard, N., A. Katty, G. Collin and O.
Tsushima, 1978a, J. Appl. Phys. 49, 2183. Gorochov, 1979, J. Solid State Chem. 27, 267.
Koshizuka, N., Y. Yokoyama, T. Okuda and T. Lisnyak, S.S. and B.D. Lichter, 1969, Trans.
Tsushima, 1978b, J. Phys. Soc. Japan, 45, Metall. Soc. AIME, 245, 2594.
1439. Locher, P.R., 1967, Solid, State Commun. 5,
Koshizuka, N., S. Ushioda and T. Tsushima, 185.
1980, Phys. Rev. B21, 1316. Locher, P.R., 1968, Z. Angew. Phys. 24, 277.
K6tzler, J. and H. von Philipsborn, 1978, Phys. Locher, P.R., 1973, Phys. Lett. 42A, 490.
Rev. Lett. 40, 790. Locher, P.R. and R.P. van Stapele, 1970, J.
Kovtun, N.M., V.K. Prokopenko and A.A. Phys. Chem. Solids 31, 2643.
Shamyakov, 1978, Solid State Commun. 26, Lotgering, F.K., 1956, Philips Res. Rep. 11, 218
877. and 11, 337.
Krawczyk, M., H. Szymczak, W. Zbieranowski Lotgering, F.K., 1964a, Solid State Commun. 2,
and J. Zmija, 1973, Acta Phys. Polonica, A44, 55.
455. Lotgering, F.K., 1964b, Proc. Int. Conf. on
Kugimiya, K. and H. Steinfink, 1968, Inorg. Magnetism, Nottingham 1964 (Institute of
Chem. 7, 1762. Physics and Physical Society, London), p.
Kun'kova, Z.E., T.G. Aminov, L.L. Golik, M.I. 533.
Elinson and V.T. Kalinnikov, 1976, Sov. Lotgering, F.K., 1965, Solid State Commun. 3,
Phys. Solid State, 18, 1212. 347.
742 R.P. VAN STAPELE

Lotgering, F.K., 1968a, J. Phys. Chem. Solids, Mejai, M. and M. Nogues, 1980, J. Magn.
29, 2193. Magn. Mat. 15-18, 487.
Lotgering, F.K., 1968b, J. Phys. Chem. Solids, Menth, A., A.R. von Neida, L.K. Shick and
29, 699. D.L. Maim, 1972, J. Phys. Chem. Solids, 33,
Lotgering, F.K., 1969, J. Phys. Chem. Solids, 1338.
30, 1429. Menyuk, N., K. Dwight and A. Wold, 1965, J.
Lotgering, F.K. and E.W. Gorter, 1957, Phys. Appl. Phys. 36, 1088.
Chem. Solids, 3, 238. Menyuk, N., K. Dwight, R.J. Arnott and A.
Lotgering, F.K. and R.P. van Stapele, 1967, Wold, 1966, J. Appl. Phys. 37, 1387.
Solid State Commun. 5, 143. Menzer, G., 1926, Z. Krist. 64, 506.
Lotgering, F.K. and R.P. van Stapele, 1968a, J. Merkulov, A.I., S.I. Radautsan and V.E.
Appl. Phys. 39, 417. Tezlevan, 1978, Phys. Status Solidi, B87,
Lotgering, F.K. and R.P. van Stapele, 1968b, K141.
Mat. Res. Bull. 3, 507. Methfessel, S. and D.C. Mattis, 1968, Magnetic
Lotgering, F.K. and G.H.A.M. van der Steen, Semiconductors, Encyclopedia of Physics
1969, Solid State Commun. 7, 1827. XVIII/1 (Springer Verlag, Berlin).
Lotgering, F.K. and G.H.A.M. van der Steen, Mimura, Y., M. Shimada and M. Koizumi,
1971a, J. Solid State Chem. 3, 574. 1974, Solid State Commun. 15, 1035.
Lotgering, F.K. and G.H.A.M. van der Steen, Minematsu, K., K. Miyatani and T. Takahashi,
1971b, J. inorg, nucl. Chem. 33, 673. 1971, J. Phys. Soc. Japan, 31, 123.
Lotgering, F.K. and G.H.A.M. van der Steen, Miyatani, K., 1970, J. Phys. Soc. Japan, 28, 259.
1971c, Solid State Commun. 9, 1741. Miyatani, K., Y. Wada and F. Okamoto, 1968,
Lotgering, F.K., R.P. van Stapele, G.H.A.M. J. Phys. Soc. Japan, 25, 369.
van der Steen and J.S. van Wieringen, 1969, Miyatani, K., T. Takahashi, K. Minematsu, S.
J. Phys. Chem. Solids, 30, 799. Osaka and K. Yoshida, 1970, Ferrites, Proc.
Lotgering, F.K., A.M. van Diepen and J.F. Int. Conf. Japan, p. 607.
Olijhoek, 1975, Solid State Commun. 17, 1149. Miyatani, K., K. Minematsu, Y. Wada, F.
Lundqvist, D., 1943, Ark. Kemi. Min. Geol. Okamoto, K. Kato and P.K. Baltzer, 1971a,
17B, Nr. 12. J. Phys. Chem. Solids, 32, 1429.
Lutz, H.D., 1966, Z. anorg, allg. Chem. 348, 36. Miyatani, K., F. Okamoto, P.K. Baltzer, S.
Lutz, H.D. and R.A. Becker, 1973, Monat- Osaka and T. Oka, 1971b, A.I.P. Conf. Proc.
shefte F. Chem. 104, 572. 5, 285.
Lutz, H.D. and M. F6her, 1971, Spectrochem. Moser, F., R.K. Ahrenkiel, E. Carnall, T.
Acta, 27A, 357. Coburn, S.L. Lyu, T.H. Lee, T. Martin and
Lutz, H.D. and K. Grendel, 1965, Z. anorg. D. Pearlman, 1971, J. Appl. Phys. 42, 1449.
allg. Chem. 337, 30. Motida, K. and S. Miyahara, 1970, J. Phys. Soc.
Lyons, D., L. Goldstein and P. Gibart, 1973, Japan, 29, 516.
Proc. Int. Conf. on Magnetism, Moscow Mott, N.F., 1949, Proc. Phys. Soc. (London)
1973, 6, 208. A62, 416.
Lyubutin, I.S. and T.V. Dmitrieva, 1975, JETP Nagaev, E.L., 1975, Soy. Phys. Usp. 18, 863.
Lett. 21, 59. Nagaev, E.L., 1977, JEPT Lett. 25, 76.
Makhotkin, V.E., G.G. Shabunina, T.G. Nakatani, I., H. Nose and K. Masumoto, 1977,
Aminov, G.I. Vinogradova, V.G. Veselago J. Japan Inst. Metals, 41, 939.
and V.T. Kalinnikov, 1975, Soy. Phys. Solid Natsume, Y. and H. Kamimura, 1972, Solid
State 16, 2034. State Commun. 11, 875.
Makhotkin, V.E., V.G. Veselago and V.T. Natta, G. and L. Passerine, 1931, R.C. Accad.
Kalinnikov, 1978a, Sov. Phys. Solid State, 20, Lincei 14, 33.
777. Nauciel-Bloch, M., A. Castets and R. Plumier,
Makhotkin, V.E., G.I. Vinogradova and V.G. 1972, Phys. Lett. 39A, 311.
Veselago, 1978b, JETP Lett. 28, 78. Ndel, L., 1948, Ann. de Phys. 3, 137.
Martin, G.W., A.T. Kellogg, R.L. White, R.M. Nogues, M., M. Mejai and L. Goldstein, 1979,
White and H. Pinch, 1969, J. Appl. Phys. 40, J. Phys. Chem. Solids, 40, 375.
1015. Ohbayashi, K., Y. Tominaga and S. Iida, 1968,
Matsumoto, G., K. Ohbayashi, K. Kohn and S. J. Phys. Soc. Japan, 24, 1173.
Iida, 1966, J. Phys. Soc. Japan, 21, 2429. Ohsawa, A., Y. Yamaguchi, H. Watanabe and
SULPHOSPINELS 743

H. Itoh, 1976, J. Phys. Soc. Japan, 40, 986 Plumier, R., M. Lecomte, A. Miedan-Gros and
and 40, 992. M. Sougi, 1977a, Physica 86--88B, 1360.
Okofiska-Kozlowska, I. and J. Krok, 1978, Z. Plumier, R., M. Sougi and M. Lecomte, 1977b,
anorg, allg. Chem. 447, 235. Phys. Lett. 60A, 341.
Okofiska-Kozlowska, I., M. Jelonek and Z. Prosser, V., P. Hlidek, P. Hoeschl, P. Polivka
Drzazga, 1977, Z. anorg, allg. Chem. 436, and M. Zvfira, 1974, Czech. J. Phys. B24,
265. 1168.
Passerini, L. and M. Baccaredda, 1931, R.C. Raccah, P.M., R.J. Bouchard and A. Wold,
Accad. Lincei, 14, 38. 1966, J. Appl. Phys. 37, 1436.
Patil, C.G. and B.S. Krishnamurthy, 1978, Phys. Rehwald, W., 1967, Phys. Rev. 155, 861.
Status Solidi, B86, 725. Riedel, E. and E. Horvath, 1969, Z. anorg, allg.
Pauling, L., 1960, The Nature of the Chemical Chem. 37!, 248.
Bond (Cornell University Press, lthac~i). Riedel, E. and E. Horvath, 1973a, Mat. Res.
Pearlman, D., E. Carnall and T.W. Martin, Bull. 8, 973.
1973, J. Solid State Chem. 7, 138. Riedel, E. and E. Horvath, 1973b, Z. anorg.
Pellerin, Y. and P. Gibart, 1969, C.R. Acad. Sc. allg. Chem. 399, 219.
Paris, B269, 615. Riedel, E. and R. Karl, 1979, private com-
Pickardt, J. and E. Riedel, 1971, J. Solid State munication.
Chem. 3, 67. Riedel, E., J. Pickardt and J. Soechtig, 1976, Z.
Pickardt, J., E. Riedel and B. Reuter, 1970, Z. anorg, allg. Chem. 419, 63.
anorg, allg. Chem. 373, 15. Robbins, M., R.H. Willens and R.C. Miller,
Pidgeon, C.R., R.B. Dennis and J.S. Webb, 1967a, Solid State Commun. 5, 933.
1973, Surface Science, 37, 340. Robbins, M., H.W. Lehmann and J.G. White,
Pinch, H.L. and S.B. Berger, 1968, J. Phys. 1967b, J. Phys. Chem. Solids, 28, 897.
Chem. Solids, 29, 2091. Robbins, M., P.K. Baltzer and E. Lopatin,
Pinch, H.L., M.J. Woods and E. Lopatin, 1970, 1968, J. Appl. Phys. 39, 662.
Mat. Res. Bull. 5, 425. Robbins, M., M.A. Miksovsky and R.C. Sher-
Pink, H., W.K. Unger, H. Schaefer and H. wood, 1969, J. Appl. Phys. 40, 2466.
Goebel, 1974, Appl. Phys. 4, 147. Robbins, M., A. Menth, M.A. Miksovsky and
Plumier, R., 1965, C.R. Acad. Sc. Paris, 260, R.C. Sherwood, 1970a, J. Phys. Chem. Solids,
3348. 31,423.
Plumier, R., 1966a, J. Appl. Phys. 37, 964. Robbins, M., R. Wolff, A.J. Kurtzig, R.C.
Plumier, R., 1966b, J. de Phys. 27, 213. Sherwood and M.A. Miksovsky, 1970b, J.
Plumier, R., 1970a, C.R. Acad. Sc. Paris, B271, Appl. Phys. 41, 1086.
184. Robbins, M., P, Gibart, L.M. Holmes, R.C.
Plumier, R., 1970b, C.R. Acad. Sc. Paris, B271, Sherwood and G.W. Hull, 1973, A.I.P. Conf.
277. Proc. 10, 1153.
Plumier, R., 1980, J. Phys. Chem. Solids, 41,871. Robbins, M., P. Gibart, D.W. Johnson, R.C.
Plumier, R. and F.K. Lotgering, 1970, Solid Sherwood and V.G. Lambrecht, 1974, J.
State Comrnun. 8, 477. Solid State Chem. 9, 170.
Plumier, R. and M. Sougi, 1969, C.R. Acad. Sc. Rys, F., J.S. Helman and W. Baltensperger,
Paris, B268, 1549. 1967, Phys. Kondens. Materie, 6, 105.
Plumier, R. and M. Sougi, 1971, Solid State Riidorff, W. and K. Stegemann, 1943, Z. anorg.
Commun. 9, 413. allg. Chem. 251, 376.
Plumier, R., R. Conte, J. Denis and M. Saji, H. and T. Yamadaya, 1972, Phys. Lett.
Nauciel-Bloch, 1971a, J. de Phys. Coll. C1 32, 41A, 365.
55. Sakai, S., T. Sugano and Y. Okabe, 1976, Japan
Plumier, R., F.K. Lotgering and R.P. van J. Appl. Phys. 15, 2023.
Stapele, 1971b, J. de Phys., Coll. C.1. 32, Salanskii, N.M. and N.A. Drokin, 1975, Sov.
324. Phys. Solid State, 17, 205.
Plumier, R., M. Lecomte, A. Miedan-Gros and Salyganov, V.I., Y.M. Yakovlev and Y.R.
M. Sougi, 1975a, Phys. Lett. 55A, 239. Shil'nikov, 1973, JETP Lett. 18, 215.
Plumier, R., M. Sougi, A. Miedan-Gros and M. Samokhvalov, A.A., V.S. Babushkin, M.I.
Lecomte, 1975b, A.I.P. Conf. Proc. 29, Simonova and T.I. Arbuzova, 1973, Sov.
410. Phys. Solid State, 14, 1883.
744 R.P. VAN STAPELE

Samokhvalov, A.A., V.V. Osipov, V.G. Kal- Spender, M.R. and A.H. Morrish, 1973, Proc.
linikov and T.A. Aminov, 1978, Sov. Phys. Fifth Int. Conf. Moessbauer Spectrometry,
Solid State, 20, 344. Bratislava 1973 (Czechoslovak Atomic Energy
Samokhvalov, A.A., V.V. Osipov, V.T. Kal- Commission, Nuclear Information Centre,
linikov and T.G. Aminov, 1979, JETP Lett. Zbraslav, 1975) part I, p. 125.
28, 382. Spender, M.R.M., J.M.D. Coey a n d A.H.
Sato, K., 1977, J. Phys. Soc. Japan, 43, Morrish, 1972, Can. J. Phys. 50, 2313.
719. Srivastava, V.C., 1969, J. Appl. Phys. 40, 1017.
'Sato, K. and T. Teranishi, 1970, J. Phys. Soc. Stasz, J., 1973, Acta Phys. Polonica, A43, 177.
Japan, 29, 523. Stauss, G.H., 1969a, J. Appl. Phys. 40, 1023.
Schaetter, G.M. and M.H. van Maaren, 1968, Stauss, G.H., 1969b, Phys. Rev. 181, 636.
Proc. 11th Int. Conf. Low Temp. Phys., St. Stauss, G.H., M. Rubinstein, J. Feinleib, K.
Andrews, p. 1033. Dwight, N. Menyuk and A. Wold, 1968, J.
Schlein, W. and A. Wold, 1972, J. Solid State Appl. Phys. 39, 667.
Chem. 4, 286. Steigmeier, E.F. and G. Harbeke, 1970, Phys.
Sekizawa, H., T. Okada and F. Ambe, 1973, Kondens. Materie, 12, 1.
Proe. Int. Conf. on Magnetism, Moscow, Stickler, J.J. and H.J. Zeiger, 1968, J. Appl.
1973, 2, 152. Phys. 39, 1021.
Selmi, A., P. Gibart and L. Goldstein, 1980, J. Stoyanov, S.G., M.N. Iliev and S.P. Stoyanova,
Magn. Magn. Mat. 15--18, 1285. 1975, Phys. Status Solidi, A30, 133.
Serres, A., 1953, J. de Phys. 14, 689. Stoyanov, S.G., M.N. Iliev and S.P. Stoyanova,
Shanditsev, V.A. and E.N. Yakovlev, 1975, 1976, Solid State Commun. 18, 1389.
Sov. Phys. Solid State, 17, 1161. Strick, G., G. Eulenberger and H. Hahn, 1968,
Shelton, R.N., D.C. Johnston and H. Adrian, Z. anorg, allg. Chem. 357, 338.
1976, Solid State Commun. 20, 1077. Suzuki, N. and H. Kamimura, 1972, Solid State
Shepherd, I.W., 1970, Solid State Commun. 8, Commun. 11, 1603.
1835. Suzuki, N. and H. Kamimura, 1973, J. Phys.
Shick, L.K. and A.R. von Neida, 1969, J. Cryst. Soc. Japan, 35, 985.
Growth, 5, 313. Takahashi, T., K. Minematsu and K. Miyatani,
Shirane, G., D.E. Cox and S.J. Pickart, 1964, J. 1971, J. Phys. Chem. Solids, 32, 1007.
Appl. Phys. 35, 954. Tanaka, M., T. Tokoro and T. Mori, 1973,
Shumilkina, E.V., 1975, Sov. Phys. Solid State, Proc. Fifth Int. Conf. Moessbauer Spec-
17, 800. trometry, Bratislava, 1973 (Czechoslovak
Shumilkina, E.V. and A.I. Obraztsov, 1975, Atomic Energy Commission, Nuclear Infor-
Sov. Phys. Solid State, 17, 802. mation Centre, Zbraslav, 1975) part I, p. 118.
Siratori, K., 1971, J. Phys. Soc. Japan, 30, 709. Taniguchu, M., Y. Kato and S. Narita, 1975,
Siratori, K. and J. Sakurai, 1975, J. Phys. Soc, Solid State Commun. 16, 261.
Japan, 38, 701. Toda, M., 1970, Appl. Phys. Lett. 17, 1.
Siratori, K., A. Tasaki and H. Asada, 1973, Int. Treitinger, L., H. Pink, H. Mews and R. Koepl,
J. Magnetism, 4, 273. 1976a, J. Magn. Magn. Mat. 3, 184.
Skinner, B.J., R.C. Erd and F.S. Grimaldi, Treitinger, L., H. Goebel and H. Pink, 1976b,
1965, Am. Min. 49, 543. Mat. Res. Bull. 11, 1375.
Sleight, A.W., 1967, Mat. Res. Bull. 2, 1107. Treitinger, L., H. Pinl~ and H. Goebel, 1978a, J.
Sleight, A.W. and H.S. Jarrett, 1968, J. Phys. Phys. Chem. Solids, 39, 149.
Chem. Solids, 29, 868. Treitinger, L., K.G. Barraclough and E. Feldt-
Solin, N.I., A.A. Samokhvalov and V.T. Kal- keller, 1978b, Mat. Res. Bull. 13, 667.
linikov, 1967, Sov. Phys. Solid State, 18, 1226. Tressler, R.E. and V.S. Stubican, 1968, J. Am.
Spender, M.R., 1973, Ph.D. Thesis, University Cer. Soc. 51, 391.
of Manitoba. Tressler, R.E., F.A. Hummel and V.S. Stubi-
Spender, M.R. and A.H. Morrish, 1971, Can. J. can, 1968, J. Am. Cer. Soc. 51, 648.
Phys. 49, 2659. Tret'yakov, Y.D., M.A. Vinnik, Y.G. Sak-
Spender, M.R. and A.H. Morrish, 1972a, Can. sonov, V.K. Kamyshova and I.V. Gordeev,
J. Phys. 50, 1125. 1975, Sov. Phys. Phys. Solid State, 17, 1184.
Spender, M.R. and A.H. Morrish, 1972b, Solid Uda, M., 1968, Sci. Pap. Inst. Phys. Chem. Res.
State Commun. 11, 1417. 62, 14.
SULPHOSPINELS 745

Ullrich, J.F. and D.H. Vincent, 1967, Phys. Watanabe, T., 1973, Solid State Commun. 12,
Lett. 25A, 731. 355.
Unger, W.K., 1975, J. Magnetism Magn. Mat. 1, White, R.M., 1969, Phys. Rev. Lett. 23, 858.
88. White, W.B. and B.A. DeAngelis, 1967, Spec-
Unger, W.K., O. Scherber and H. Stremme, trochim. Acta 23A, 985.
1974, Int. J. Magnetism, 6, 313. White, J.G. and M. Robbins, 1968, J. Appl.
Van Diepen, A.M. and R.P. van Stapele, 1972, Phys. 39, 664.
Phys. Rev. B5, 2462. Wilkinson, C., B.M. Knapp and J.B. Forsyth,
Van Diepen, A.M. and R.P. van Stapele, 1973, 1976, J. Phys. C: Solid State Phys. 9,
Solid State Commun. 13, 165L 4021.
Van Diepen, A.M., F.K. Lotgering and J.F. Williamson, D.P. and N.W. Grimes, 1974, J.
Olijhoek, 1976, J. Magn. Magn. Mat. 3, 117. Phys. D: Appl. Phys. 7, 1.
Van Gorkom, G.G.P., J.C.M. Henning and Wittekoek, S. and P.F. Bongers, 1969, Solid
R.P. van Stapele, 1973, Phys. Rev. BS, 955. State Commun. 7, 1719.
Van Maaren, M.H. and H.B. Harland, 1969, Wittekoek, S. and P.F. Bongers, 1970, IBM J.
Phys. Lett. 30A, 204. Res. Develop. 14, 312.
Van Maaren, M.H., G.M. Schaeller and F.K. Wittekoek, S. and G. Rinzema, 1971, Phys.
Lotgering, 1967, Phys. Lett. 25A, 238. Status Solidi, B44, 849.
Van Maaren, M.H., H.B. Harland and E.E. Wojtowicz, P.J., 1969, IEEE Trans. on Magn.
Havinga, 1970a, Solid State Commun. 8, 5,840.
1933. Wojtowicz, P.J., P.K. Baltzer and M. Robbins,
Van Maaren, M.H., H.B. Harland and E.E. 1967, J. Phys. Chem. Solids~ 28, 2423.
Havinga, 1970b, Proc. 12th Int. Conf. on Low Wojtowicz, P.J., L. Darcy and M. Rayl, 1969, J.
Temp. Phys. Kyoto 1970, p. 357. Appl. Phys. 40, 1023.
Von Neida, A.R. and L.K. Shick, 1969, J. Appl. Wollan, E.O., 1960, Phys. Rew 117, 387.
Phys. 40, 1013. Yafet, Y. and C. Kittel, 1952, Phys. Rev. 87,
Van Stapele, R.P. and F.K. Lotgering, 1970, J. 290.
Phys. Chem. Solids, 31, 1547. Yagnik, C.M. and H.B. Mathur, 1967, Solid
Van Stapele, R.P., J.S. van Wieringen and P.F. State Commun. 5, 841.
Bongers, 1971, J. de Phys., Coll. C1 32, 53. Yamashita, O., Y. Yamaguchi, I. Nakatani, H.
Veselago, V.G., E.S. Vigeleva, G.I. Vino- Watanabe and K. Masumoto, 1979a, J. Phys.
gradova, V.T. Kalinnikov and V.E. Makhot- Soc. Japan, 46, 1145.
kin, 1972a, JETP. Lett. 15, 223. Yamashita, O., H. Yamauchi, Y. Yamaguchi and
Veselago, V.G., E.S. Vigeleva., G.I. Vino- H. Watanabe, 1979b, J. Phys. Soc. Japan, 47,
gradova, V.T. Kalinnikov and V.E. Mok- 450.
hotkin, 1972b, Proc. Int. Conf. Phys. Semi- Yanase, A., 1971, Solid State Commun. 9, 2111.
cond., Warsaw, p. 1300. Yanase, A., 1972, Intern. J. Magnetism, 2, 99.
Vinogradova, G.I., V.G. Veselago, V.E. Mak- Yanase, A. and T. Kasuya, 1968a, J. Appl.
hotkin, I.S. Kavaleva, V.A. Levshin, G.G. Phys. 39, 430.
Shabunina and V.T. Kalinnikov, 1978, Sov. Yanase, A. and T. Kasuya, 1968b, J. Phys. Soc.
Phys. Solid State, 20, 827. Japan, 25, 1025.
Wagner, C., 1935, Z. Tech. Phys. 16, 327. Yanase, A., T. Kasuya and T. Takeda, 1970,
Wagner, V., H. Mitlehner and R. Geick, 1971, Ferrites, Proc. Int. Conf. Japan, p. 604.
Optics Commun. 2, 429. Yokoyama, H. and S. Chiba, 1969, J. Phys. Soc.
Wakaki, M. and T. Arai, 1978, Solid State Japan, 27, 505.
Commun. 26, 757. Yokoyama, H., S. Chiba and N. Ichinose, 1970,
Wakaki,x M., T. Arai and K. Kudo, 1975, Solid Ferrites, Proc. Int. Conf. Japan, p. 611.
State Commun. 16, 679. Yokoyama, H., R. Watanabe and S. Chiba,
Wakamura, K., S. Onari, T. Arai and K. Kudo, 1967a, J. Phys. Soc. Japan, 22, 659.
1971, J. Phys. Soc. Japan, 31, 1845. Yokoyama, H., R. Watanabe and S. Chiba,
Wakamura, K., T. Arai, S. Onari, K. Kudo and 1967b, J. Phys. Soc. Japan, 23, 450.
T. Takahashi, 1973, J. Phys. Soc. Japan, 35, Zener, C., 1951, Phys. Rev. 81, 440 and 82, 403.
1430. Zheru, I.I., I.G. Lupya, K.G. Nikiforov, S.I.
Wakamura, K., T. Arai and K. Kudo, 1976a, J. Radautsan and V.E. T6zl6van, 1978, Sov.
Phys. Soc. Japan, 40, 1118. Phys. Solid State, 20, 884.
Wakamura, K., T. Arai and K. Kudo, 1976b, J. Zvfira, M., V. Prosser, A. Schlegel and P.
Phys. Soc. Japan, 41, 130. Wachter, 1979, J. Magn. Magn. Mat. 12, 219.
chapter 9

TRANSPORT PROPERTIES OF
FERROMAGNETS

I.A. C A M P B E L L A N D A. FERT
Laboratoire de Physique des Sofides
Universit6 Paris-Sud, 91405 Orsay
France

Ferromagnetic Materials, Vol. 3


Edited by E.P. Wohlfarth
O North-HollandPublishing Company, 1982

747
CONTENTS

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 751
1. G e n e r a l p r o p e r t i e s of t r a n s p o r t in f e r r o m a g n e t s . . . . . . . . . . . . . 751
1.1. R e s i s t i v i t y a n d H a l l effect of a m o n o d o m a i n p o l y c r y s t a l . . . . . . . . . 751
1.1.1. S p o n t a n e o u s resistivity a n i s o t r o p y . . . . . . . . . . . . . . 752
1.1.2. E x t r a o r d i n a r y H a l l effect . . . . . . . . . . . . . . . . . 754
1.1,3. P l a n a r H a l l effect . . . . . . . . . . . . . . . . . . . . 754
1.2. R e s i s t i v i t y a n d H a l l effect in single c r y s t a l f e r r o m a g n e t s . . . . . . . . . 755
1.3. T h e r m a l a n d t h e r m o e l e c t r i c effects in p o l y c r y s t a l s . . . . . . . . . . . 756
2. E l e c t r i c a l resistivity of f e r r o m a g n e t s . . . . . . . . . . . . . . . . . 757
2.1. T h e o r e t i c a l m o d e l s . . . . . . . . . . . . . . . . . . . . . 757
2.1.1. S p i n d i s o r d e r s c a t t e r i n g . . . . . . . . . . . . . . . . . 757
2.1.2. T w o c u r r e n t m o d e l . . . . . . . . . . . . . . . . . . . 758
2.2. R e s i s t i v i t y of p u r e m e t a l s . . . . . . . . . . . . . . . . . . . 762
2.2.1. T a b u l a r r e s u l t s . . . . . . . . . . . . . . . . . . . . 762
2.2,2. R e s i s t i v i t y at l o w t e m p e r a t u r e s . . . . . . . . . . . . . . . 762
2.2.3. R e s i d u a l resistivity . . . . . . . . . . . . . . . . . . . 764
2.2.4. H i g h field b e h a v i o u r . . . . . . . . . . . . . . . . . . . 765
2.3. A l l o y s : r e s i d u a l resistivity a n d t e m p e r a t u r e d e p e n d e n c e o f resistivity . . . . 766
2.3.1. N i c k e l h o s t . . . . . . . . . . . . . . . . . . . . . . 768
2,3.2. C o b a l t h o s t . . . . . . . . . . . . . . . . . . . . . . 771
2.3.3. I r o n h o s t . . . . . . . . . . . . . . . . . . . . . . . 771
2.3,4. A l l o y s c o n t a i n i n g i n t e r s t i t i a l i m p u r i t i e s . . . . . . . . . . . . . . 772
2.4. H i g h t e m p e r a t u r e a n d critical p o i n t b e h a v i o u r . . . . . . . . . . . . 773
3, O t h e r t r a n s p o r t p r o p e r t i e s of Ni, C o , F e a n d t h e i r all~oys . . . . . . . . . . 776
3.1. O r d i n a r y m a g n e t o r e s i s t a n c e . . . . . . . . . . . . . . . . . . 776
3.2. O r d i n a r y H a l l coefficient . . . . . . . . . . . . . . . . . . . 778
3.3. S p o n t a n e o u s resistivity a n i s o t r o p y . . . . . . . . . . . . . . . . 779
3.4. E x t r a o r d i n a r y H a l l effect . . . . . . . . . . . . . . . . . . . 783
3.5. T h e r m o e l e c t r i c p o w e r . . . . . . . . . . . . . . . . . . . . 790
3.6. N e r n s t - E t t i n g s h a u s e n effect . . . . . . . . . . . . . . . . . . 792
3.7. T h e r m a l c o n d u c t i v i t y . . . . . . . . . . . . . . . . . . . . . 792
4. D i l u t e f e r r o m a g n e t i c a l l o y s . . . . . . . . . . . . . . . . . . . . 793
4.1. P a l l a d i u m b a s e d a l l o y s . . . . . . . . . . . . . . . . . . . . 793
4.1.1. R e s i s t i v i t y a n d i s o t r o p i c m a g n e t o r e s i s t a n c e . . . . . . . . . . . 793
4.1.2. M a g n e t o r e s i s t a n c e a n i s o t r o p y . . . . . . . . . . . . . . . 794
4.1.3. E x t r a o r d i n a r y H a l l effect . . . . . . . . . . . . . . . . . 794
4.1.4. T h e r m o e l e c t r i c p o w e r . . . . . . . . . . . . . . . . . . . 794
4.2. P l a t i n u m b a s e d a l l o y s . . . . . . . . . . . . . . . . . . . . 795

748
5. A m o r p h o u s alloys . . . . . . . . . . . . . . . . . . . . . . . 795
5.1. Resistivity of a m o r p h o u s alloys . . . . . . . . . . . . . . . . . 795
5.2. Hall effect and resistivity anisotropy of a m o r p h o u s alloys . . . . . . . . 800
References . . . . . . . . . . . . . . . . . . . . . . . . . . . 800

749
Introduction

The scope of this chapter will in fact be rather more restricted than its title might
suggest. W e will outline some very general properties of transport in ferro-
magnets and will summarize the models that have been used. We will then
review in detail results on particular systems. W e will not treat magnetic semi-
conductors, and we only occasionally mention rare earth metals and their alloys,
as they will be treated in another chapter*. Because of the lack of systematic data,
we will not review the properties of magnetic intermetallic compounds as such,
but will mention results on these compounds as and when they exist. A section is
devoted to a m o r p h o u s ferromagnets. However, we will be mostly concerned with
the transport properties of the transition ferromagnets Fe, Co and Ni and their
alloys, while drawing examples from other classes of magnetic metals to illustrate
particular types of behaviour.
The subject has advanced so much over the last 25 years that this chapter has
little to do with the equivalent one in Bozorth's book. However, a n u m b e r of
books and review articles have been very useful; a list of them is given at the end
of the chapter (Jan 1957, Mott 1964, H u r d 1974 and 1972, Dorleijn 1976).

1. General properties of transport in ferromagnets

W e will discuss effects which arise quite generally as a consequence of the


symmetry properties of the ferromagnetic state.

1.1. Resistivity and Hall effect of a monodomain polycrystal

The components Ei of the electric field inside a conductor are related to the
current density ~ through

J~i ~" E PiJJ] , ( 1)


J

* Vol. 1, oh. 3, by Legvold.

751
752 I.A. CAMPBELL AND A. FERT

where the pij coefficients form the resistivity tensor. Suppose we have a random
polycrystal with its magnetization saturated in the direction z. From symmetry
arguments (Birss 1964, H u r d 1974) one finds that such a magnetized isotropic
medium has a resistivity tensor of the form:

[P~i] = |[_
p H (orP±(B)
B ) -pH(B)
pa(B) 00 1 . (2)
0 RjI(B)

This form of the resistivity tensor corresponds to the following expression of the
electric field E:

E = p I ( B ) J + [PlI(B)- p±(B)][o~ • J ] ~ + pH(B)oz x J, (3)


where J is the current vector and a is a unit vector in the magnetization direction.
The pii(B) are functions of the induction B, which depends on the external field H
and on the demagnetizing factor D of the particular sample geometry,

B = H + 4rrM(1 - D ) . (4)

This is because of Lorentz force effects which exist in any conductor. That the
effective field acting on the electron trajectories in a ferromagnet is indeed B
seems physically reasonable (Kittel 1963) and has been verified experimentally
(Anderson and Gold 1963, Tsui 1967, Hodges et al. 1967). Conventionally, the
coefficients Pij are split up into two parts pij(B) = pij + p°(B) where the p~j are the
" s p o n t a n e o u s " or "extraordinary" coefficients, and the p°(B) are the "ordinary"
coefficients. It should be noted that the Pit cannot be measured directly without
some form of extrapolation because of the presence of the internal Lorentz field.
In practice, this extrapolation causes little difficulty except for fairly pure samples
at low temperatures where B/p can be high and the ordinary effects become large.
Assuming that this extrapolation to zero B has been made we have the three
spontaneous parameters:
Pll = the resistivity for J parallel to M at B = 0,
p± = the resistivity for J perpendicular to M at B = 0,
PH = the extraordinary Hall resistivity.

1.1.1. Spontaneous resistivity anisotropy


The fact that the diagonal elements Pll and p± in (2) are unequal means that the
resistivity depends on the relative orientation of M and d. Taking the geometry of
fig. 1 and calling 0 the angle between M and J, as by definition the resistivity is

p = E. J/fJI 2 ,

we have from eq. (3),


TRANSPORT PROPERTIES OF FERROMAGNETS 753

[
7

Fig. 1. Experimental geometry for transport properties. T h e current J is constrained to flow along the
direction x, and whatever the applied field or magnetization direction, the resistivity is proportional to
the voltage between probes A and B, p = ( V B - VA)bt/J1. For the conventional Hall geometry, the
applied field or magnetization direction is along z and the Hall voltage is m e a s u r e d between C and D.
T h e n PH = ( V O - Vc)t/J and R0 = ( V D - Vc)t/JH or R~ = ( V D - Vc)t/J 4~rM.

ps=0 = Pll + 32p~ + (cos 2 0 - ½)(Pll- P~) (5)

The relative spontaneous anisotropy of the resistivity is defined as:

Ap _ PlI- P±
1
(6)
P 3Pll + 2p±

It can have either sign, and while generally values of a few percent are typical, certain
systems show more than 20% anisotropy.
Figure 2 shows schematically typical resistivity changes as a function of applied
field, and the method of extrapolation to obtain Pll and p± is indicated. N o t e that
the resistivity of the zero field state depends on the exact domain configuration, so
it is history dependent and is not well defined even for a given sample at a given
temperature. In the same way, the change of resistivity with field below technical

{8=0}. . . . .

~a=01-- ,
H
I1

B/p 8=0,1
Fig. 2. Schematic extrapolation for a ferromagnetic of p(H) resistivity curves to B = 0. T h e heavy
lines indicate observed resistivity as a function of field when the field is applied parallel (11) or
perpendicular (±) to the current direction. Arrows show the regions of incomplete technical saturation.
Dotted lines indicate extrapolations from the saturated regions to the respective B = 0 points
(B = H + 4~-M(1 - D ) where D is the demagnetizing factor for transverse or longitudinal fields).
754 I.A. C A M P B E L L A N D A. F E R T

saturation depends on the magnetization process. We have assumed above that in


the saturated region only Lorentz force effects are important for the variation of
Pll or p± with B. In fact, the external field H may also increase the magnetization
of the sample which will affect p, because of a reduction in spin disorder
resistivity. This is particularly important near T¢.

1.1.2. Extraordinary Hall effect


The off-diagonal terms +pH in (2) lead to the extraordinary Hall voltage,

E H ( B = 0) = p H a × J , (7)

perpendicular to M and J. This is usually measured in the conventional Hall


geometry with M perpendicular to J (fig. 1). We can also define the extraordinary
Hall angle

¢hH= P~/Pi (8)

and, by analogy with the definition of the ordinary Hall coefficient R0 = p ° / B we


have the extraordinary Hall coefficient

Rs = pn/4~M. (9)

Once again, extrapolation of the Hall voltage curve as a function of B from the
saturated region back to B = 0 is necessary to obtain PH (fig. 3). This is easy in the
low field regime oJc~ 1 (we is the cyclotron frequency and ~- is the electronic
relaxation time) but the separation between ordinary and extraordinary effects is
difficult when ~o~-~ 1 where the ordinary Hall voltage is no longer linear in B.

~(~0] -
/
H
I,

B=O 0
Fig. 3. Schematic extrapolation for a ferromagnet of a Hall resistivity curve to B = 0. On (B = 0) is the
extraordinary Hall resistivity.

1.1.3. Planar Hall effect


This is a misnomer, as was already pointed out by Jan (1957). When J is in the
direction x, the voltage is measured perpendicular to J in the direction y and M is
rotated in the plane xy, then from eq. (4):
TRANSPORT PROPERTIES OF FERROMAGNETS 755

Ey = (P/I- P±) cos 0 sin OJ ,

where j r . M = cos 0. This is a manifestation of the resistivity anisotropy and is an


effect even in field which has nothing to do with the Hall effect.

1.2. Resistivity and Hall effect in single crystal ferromagnets

The symmetry arguments can be extended to single crystals (see, e.g., Hurd 1974).
It is found that even for cubic crystals, the resistivity becomes dependent on the
orientations of the current and the magnetization with respect to the crystal axes ,
and the extraordinary Hall effect depends on the magnetization direction.
For a cubic ferromagnet with M in the direction (11, 12, 13) and J in the
direction (/31,/32,/33) we have the D6ring expression (D6ring 1938):

O ~. Oo[l+k1(a2/321+ 2 2 ..{- 0/3/33


Of 2/32 2 2 -- ~)
1

+ k2(20L1o~2/31/32 + 2ol2o,3/32/33 + 2ce3oq/33/31 ) + k 3 ( S _ 1)


..~ k , ( ~ 14 / 321 + ~ / 3 2 + 4 2 + z3S - I)
13/33

+ k5(2~i~2,~2/31/3~ + 2~2,~,-,~/32/3, + 2 - , ~ 1 - ~ / 3 , / 3 3 1 ,

where S = O /2l a 22.~_ OL20~3


2 2 -}- 0~30~
2 21 (10)
Other equivalent definitions have been given. Kittel and Van Vleck (1960)
suggested that a physically more significant definition for the magnetoelastic
coefficients (which have the same symmetry as the resistivity coefficients) would
regroup the terms to give new coefficients: K1 = k l + 6k4; K2 = k2 + ~k5;I K3 =
k3 + 2k4; K4 = k4; /(5 = }ks and p; = p0(1- 2k4). This would separate out terms
homogeneously second order (K1,/£2) and fourth order (/(3, K4, Ks) in the mag-
netization.
The resistivity

p = t~ + ( p , - a i ) ( c o s 2 0 - 1) (11)

for a magnetized cubic polycrystal is related to the two types of coefficients by:

t~ = t~o(1 - ~ k 3 ) -= o ;

and

2 3 12 3 2
Pfl- P. = po(~kl + 3k2 + ggk4 + 5~k5) ~ po(~K1 + 3/(2). (12)

The resistivity of a demagnetized cubic polycrystal is

memag = P ~--p0(1 -- ~k3) --=p ; (13)


756 I.A. CAMPBELLAND A. FERT

if the magnetization is randomly oriented with respect to the crystal axes. In


contrast, if there are easy axes, Pdemagis different from/5, for example,

Pdemag = /90

for (111) easy axes.


The D6ring expression only gives the leading terms in an infinite expansion.
Higher order terms can be necessary in interpreting experimental results.
For the extraordinary Hall effect in a cubic monocrystal, instead of defining the
z-axis as the M direction, we use the crystal axes to define x, y and z. Then the
asymmetric part of the resistivity tensor becomes (Hurd 1974).

[ A3
o -a3
0 -A1
A2 1 with Ai = o~i(eo+ ela~ + e2c~4+ e3S). (14)
-A2 A1 0
This gives a Hall voltage which must be perpendicular to J, but is no longer
necessarily perpendicular to M, and which is anisotropic.
We might note that in cubic crystals the ordinary Hall effect is isotropic in the
low field limit.
For ferromagnetic monocrystals with other symmetries such as hexagonal,
analogous expressions can be derived (Birss 1964, Hurd 1974).

1.3. Thermal and thermoelectric effects in polycrystaIs


The thermal conductivity and the Righi-Leduc effects are the pure thermal
analogues of the resistivity and the Hall effect, with heat currents replacing
electrical currents and thermal gradients replacing electrical field gradients. The
phenomenological separation into spontaneous and ordinary effects carries over
entirely. The only difference lies in the necessity to define whether the experi-
mental conditions are isothermal or adiabatic (Jan 1957).
Similarly, for a magnetized ferromagnetic polycrystal subject to a temperature
gradient VT, the resultant electric field is

E, = E Si,V,T,

with

S/j = NE S± 0 .
0 S~l

Here, SII and Sl are the isothermal thermoelectric powers in parallel and per-
pendicular geometries. SNE represents the spontaneous Nernst-Ettingshausen
effect.
TRANSPORT PROPERTIES OF FERROMAGNETS 757

For both the thermal effects and the thermoelectric effects the generalization to
monocrystals is just the same as for the electrical effects.

2. Electrical resistivity of ferromagnets

2.1. Theoretical models

For non-magnetic metals, current is carried by electrons which are scattered by


phonons and by impurities or defects. To a first approximation the two scatterings
give additive contributions to the resistivity

p ( r ) = Po + pp(T),

where p0 is the residual resistivity and pp(T) is the pure metal resistivity at
temperature T. This is known as Matthiessen's rule. Deviations from this rule are
small and are generally explained in terms of differences in the anisotropies of the
relaxation time of impurity and of phonon scattering over the Fermi surface
(Dugdale et al. 1967). In magnetic metals a number of new effects appear.

2.1.1. Spin disorder scattering


On the simple model of well defined local moments in a simple conduction band,
an exchange interaction between the local and conduction electron spins, J and s
respectively, of the type F s . J will give rise to spin disorder scattering (Kasuya
1956, 1959, De Gennes and Friedel 1958, Van Peski Tinbergen and Dekker 1963).
Well above the ordering temperature there will be a temperature independent
paramagnetic resistivity

- kv(mF)2 J ( J + 1)
pM - 4~e2zh3
(15)

where kF is the Fermi wave vector and z the number of conduction electrons per
atom. This paramagnetic resistivity is actually the sum of a non-spin-flip term (due
to szJz interactions and proportional to j2) and a spin-flip term (due to s - J + + s+J
and proportional to J). For the region around Tc the resistivity depends on the
spin-spin correlation function (see section 2.4). As T drops both the spin-flip and
non-spin-flip scattering begin to freeze out. Kasuya (1956) finds

~m -- j ( j
PM
_}_ 1)
(J-I<S)I)(J + 1 + KS)t) (16)

Finally, at low temperatures magnetic scattering only remains as magnon-electron


scattering. A number of authors have calculated the low temperature form of this
scattering. If the spin '~ and spin + conduction electron Fermi surfaces are
assumed to be identical spheres with the same relaxation rates, the contribution to
758 I.A. CAMPBELL AND A. FERT

the resistivity is proportional to T 2 (Vonsovskii 1948, Turov 1955, Kasuya 1959,


Mannari 1959). Mannari (1959) finds*

_ ¢r3 N J m F 2 (_~) 2
Pm 32 eZzhE~ (kT)2 ' (17)

where/xe is the effective mass of the magnons (/Xe= h2/2D). The resistivity varies
as T 2 because the loss of momentum of the total electron system due to collisions
with the magnons is proportional to

q
f exp(Dq~/kT)- 1
q2 dq
'

which is equal to a constant times T 2. Here q is the magnon wave vector. The
final q2 factor inside the integral is a small angle scattering factor, and the rest of
the integrand represents the number of magnons with vector q at temperature T
which undergo a collision. The reasoning is essentially the same as that leading to
the well known T 5 limiting dependence for electron-phonon scattering in non-
magnetic metals except that:
(i) the dispersion relation for magnons is E = D q 2 instead of E = aq for
phonons.
(ii) the coupling strength for electron-magnon collisions is independent of q
instead of proportional to q as it is for electron-phonon collisions.
If the conduction band is polarized or if there is scattering from s to d Fermi
surfaces or if there is a magnon energy gap, the electron-magnon scattering will
tend to drop off exponentially at low temperature instead of the T 2 behavior
(Abelskii and Turov 1960, Goodings 1963). Other mechanisms giving various
temperature dependences have been also suggested (Vonsovskii 1955, Turov and
Voloshinskii 1967) but appear to give extremely small contributions (see section
2.2).
The previous models generally assume that the spin T and spin { electrons
have the same relaxation times and carry the same current. If there are different
spin 1' and spin { currents, there will be a contribution of the magnons to the
resistivity through spin mixing. This appears to be the major magnon contribution
in many ferromagnetic alloys. This question Will be treated in section 2.1.2.
Finally, in dilute random ferromagnets (Pd_Fe at low concentrations is the best
example of this type of system) the incoherent part of the magnon scattering-
without momentum c o n s e r v a t i o n - b e c o m e s predominant. The scattering rate is
then simply proportional to the number of magnons giving p ( T ) - p o - T 3/2
(Turner and Long 1970, Mills et al. 1971).

2.1.2. T w o current model


All the discussion given up to now has neglected any spin independent potential.
If there is a scattering potential V at magnetic sites in addition to Fs • J, then at

* Note that the result of Kasuya (1959) differs from eq. (17) by a factor 3z/'n'.
TRANSPORT PROPERTIES OF FERROMAGNETS 759

low temperatures the spin I' and spin + electrons will be subject to potentials
V + F J and V - F J respectively, together with weak spin-flip scattering by
magnons (from F(s+J - + s-J+)). This will give rise to different spin t and spin $
currents. Alternatively, in terms of the s-d band model which is widely used for
transition ferromagnets, the d t and d + densities of states at the Fermi level are
different so the s to d scattering rates will be different for spin 1' and spin $
conduction electrons. This approach was used early on to explain the p ( T ) of
transition ferromagnets up to and above the Curie temperature (Mott 1936, 1964).
It now appears that the simple s-d band model approach is questionable at high
temperatures where the local spin aspect seems to be dominant even for a typical
" b a n d " ferromagnet such as Ni. At low temperatures however where Sz is almost
a good quantum number the s-d model is more appropriate.
The electronic structure of the transition ferromagnets has been studied in-
tensively, and low temperature measurements such as the de Haas van Alphen
effect in pure metal samples show that they have complex Fermi surfaces of much
the same type as those of non-magnetic transition elements except that k 1' and
k ~ states are not equivalent. Detailed band structure models have been set up
which are in good agreement with the measurements and which show that the
electrons wave functions are s and d like, generally hybridized (Visscher and
Falicov 1972). An extreme s-d model where s-d hybridization is assumed to be
weak provides a good basis for the discussion of a wide range of alloy properties
(Friedel 1967).
We will now consider the resistivity again. Quite generally, at low temperature
the electron spin direction is well defined if we ignore magnon scattering and
spin-orbit effects. Then in any model we will have conduction in parallel by two
independent currents. If the corresponding resistivities are Pt, P+ the total
observed resistivity is*:

= PtP+ (18)
P Pt + P+

Inside each p~ we can have complications such as s and d bands but eq. (18) still
remains strictly valid. If now there is transfer of momentum between the two
currents by spin mixing scatterings (e.g. electron-magnon or spin-orbit scattering)
then again quite generally (Fert and Campbell 1976),

PtP+ + Pt +(Pt + P+)


P= pt+p~+4pt~ , (19)

where

P* = P * t / X ~ + P t $ / X t X ~ , P$ = P ~ $ / X 2 ~ + P t $ / X t X ~ , (2o)
and
* Electrons with magnetic moment parallel to the total magnetization, i.e., electrons of the majority
spin band, are indicated by t ; electrons of the minority spin band by $ :
760 I.A. C A M P B E L L A N D A. F E R T

P~ ~ = - P t + / X , X+ .

H e r e p~, are integrals over the transition rates p(kcr, k'cr') for scattering from one
electronic state to another and X~ are integrals over driving terms:
Vk " (e Ofk/OEk)E where E is the electric field, vk the electron velocity andfk the Fermi
function (Ziman 1960).
If the occupation n u m b e r of each k state under zero electric field is ~ , then we
can write the occupation n u m b e r under applied field as

fk = ~ - 4)k(df°kldEk). (21)

We can use the variational principle, using k . u as a trial function for the
function k, where u is a unit vector parallel to the applied field. We then get:

P'~ - X ~ k B T 1 ~, f (k. u)[(k - k')" u]P(ko-, k'o") d k d k ' (22)

P* ~ 1
X t X,t k B T f (k' • u)(k • u ) P ( k t k'+ ) d k d k ' , (23)

where P(ko-, k'o-') is the equilibrium scattering rate between (kcr) and (k'o-').
Matthiessens' rule is assumed for each of the three terms

x x

where x refers to different types of scattering centres (phonons, magnons, each


sort of impurity). T h e r e are t e m p e r a t u r e independent " p u r e metal" terms Pit ( T ) ,
pi+(T) and p~ ~(T) which go to zero at zero t e m p e r a t u r e and are assumed to be
independent of impurity concentration in dilute alloys, and there are impurity
terms which are assumed to be t e m p e r a t u r e independent. The spin-flip scattering
by impurities (via spin-orbit coupling) is generally neglected, which leaves two
impurity terms P0 t and p0 ~.
The model then predicts a deviation from Matthiessens' rule in the residual
resistivity of ternary alloys (Fert and Campbell 1976):

(O/A -- aB)2pAPB
A p = PAB -- (PA + PB) -- a , , ) 2 a B p A + (1 + aB)Za oB ' (24)

where

a A = P A $ / P A "r and O/B = PB $/PB t "

The analysis of the deviations in alloys with different relative concentrations of A


and B can be used to determine aA, aB (Campbell et al. 1967, Cadeville et al.
TRANSPORT PROPERTIES OF FERROMAGNETS 761

1968, Dorleijn and Miedema 1975a, Fert and Campbell 1976, Dorleijn 1976).
For a binary alloy at finite temperature the general equation (19) can be used,
with

P~ = Po~ + pi~(T).

For the low temperature range where

Pot, po+ >>p~t (T), pi+ (T), Pt +( T ) ,

this reduces to:

(1 p~(T)+ k ~ + - l ] Pt $ ( T ) , (25)

where

tx = Pil (T)
pi~: ( T) '

Po = PotPo+ (residual resistivity)


Pot +Po~

and

pi(T) = PiT(T)Pi~(T) (which is not the ideal pure metal resistivity).


pi,r(T)+ pi$ (T)

The term of eq. (25) proportional to p$ ~(T) will give a strong variation of the
resistivity as a function of the temperature when a is very different from unity; in
nickel, for example, Co, Fe or Mn impurities enhance the low temperature
resistivity variation by almost an order of magnitude. The form of the electron-
magnon contribution to p~ $(T) has been calculated by Fert (1969) and Mills et al.
(!971) using a spin-split spherical conduction band model. In alloys with a -~ 1 the
temperature dependence of the resistivity will nearly be that of pi(T).
Using experimental data on binary and ternary alloys, it turns out to be possible
to obtain consistent values for the parameters for a number of impurities, together
with estimates of the temperature behaviour of the pure metal terms (see section
2.3). It is important to note that it is not possible to obtain a full description of the
pure metal behaviour, i.e., the three pure metal terms, without analyzing alloy
data.
Another way of treating the two current conduction has been presented by
Yamashita and Hayakawa (1976). They start from a realistic band structure model
for Ni and calculate the resistivity by numerically solving coupled spin 1' and spin
$ Boltzmann equations for series of k vectors. They find that the electron-
magnon contribution to the resistivity is very small when there is no impurity or
762 I.A. CAMPBELL AND A. FERT

phonon scattering but becomes important when impurity or phonon scattering


makes the spin 1" and spin $ mean free paths different; in the latter case, there is
no more "cancellation between outgoing and incoming scatterings". This is
another way of describing the spin-mixing effect of the magnons.

2.2. Resistivity of pure metals

2.2.1. Tabular results


The resistivity (and also the thermoelectric power and thermal conductivity) of
pure Fe, Co and Ni are given in tabular form over a wide range of temperatures
by Laubitz et al. (1973, 1976) and and Fulkerson et al. (1966).
Data on polycrystalline hexagonal Co should be treated with caution, as the
transport properties of monocrystals are highly anisotropic. At room temperature
(Matsumoto et al. 1966)

pc = 10.3 ixf~cm, Po = 5.5 ixlIcm

where Pc is the resistivity measured along the c axis while pp is measured in the
plane perpendicular to the c direction. Texture effects in polycrystals will certainly
be important.

2.2.2. Resistivity at low temperatures


With high purity samples at low temperatures the presence of the induction in
each ferromagnetic domain means that the Lorentz wO- is not negligible even in
zero applied field. There is an associated "internal" magnetoresistance and to get
meaningful results for the intrinsic low field resistivity this effect must be eli-
minated as well as possible. Careful extrapolations to B = 0 using Kohlers' law
2Xp/po = f(B/po) have been done for Ni (Schwerer and Silcox 1968), Co (Volken-
shtein et al. 1973) and Fe (Volkenshtein and Dyakina 1971). For Fe in particular
the "internal" magnetoresistance is very important, partly because of the high
value of 4~-M and partly because Fe behaves as a compensated metal (see section
2.2.4) so that the transverse magnetoresistance can become very strong, whereas
the longitudinal magnetoresistance is relatively much weaker. This means that the
observed resistivity of a high purity Fe sample at low temperature is strongly
dependent on the domain configuration, which regulates how much of the sample
is submitted to transverse magnetization and how much to longitudinal; the
domain configuration is a function of applied field, stresses and measuring current.
The resistivity behaviour arising from such effects of internal magnetoresistance
has been studied in detail in Fe whisker monocrystals (Taylor et al. 1968, Shumate
et al. 1970, Berger 1978); an example of experimental results is given in fig. 4. A
contribution to the resistivity from the internal Hall effect has also been found in
Co monocrystals; this contribution is associated with the zig-zag path of the
conduction electrons which is induced by the Hall effect in a polydomain sample
(Ramanan and Berger 1978, Berger 1978).
The low temperature resistivity of Ni, Co and Fe has been found to vary as
TRANSPORT PROPERTIES OF FERROMAGNETS 763

1800-

U
1600-

-~ 1400-

1200-
g

N 1000-
%
800
~3
~- 0 I ~ I { I I L I

-0.2 -

-0,4
0_.

~-- -0.6

-0.8 oo>
I I I I I I I
100 200 300 400 500 600 700 800 Oe
Magnetic FieLd

Fig. 4. Magnetization of iron single crystals (above) and magnetoresistance (below) of (100) and (111)
iron whiskers at 4.2 K as a function of applied field (after Taylor et al. 1968).

p o + A T 2 by most workers; above 1 0 K an additional term in T 4 is generally


needed (White and Woods 1959, Greig and Harrison 1965, White and Tainsh
1967, Schwerer and Silcox 1968, Beitcham et al. 1970). In early work no mag-
netoresistance corrections were made, with the result that rather varied values of
A were obtained. In a number of samples a term linear in T was also needed, but
careful experiments on the effect of magnetic fields show that this linear term is
not intrinsic but is a result of the internal magnetoresistance (Volkenstein et al.
1971, 1973).
The best values of A from data to which the magnetoresistance correction was
applied to obtain values at B = 0 are:
A = 9.5 x 10-12 ~ c m K -2 for Ni (Schwerer and Silcox 1968)
A = 16 × 10-12 l"~cmK-2 for Co (Volkenstein et al. 1973)
A = 15 × 10-12 ~ c m K -2 for Fe (Volkenstein et al. 1971, 1973).
These values however are not quite consistent with measurements on other
samples, even if internal magnetoresistance effects are taken into account. It is
likely that p(T) depends in some way on the nature of the residual impurities. As
the scattering by residual impurities is still predominant up to 10 K in the purest
samples, p(T) is actually expected to show deviations from Matthiessens' rule
similar to those observed in alloys and explained in the two-current model
764 I.A. CAMPBELL AND A. FERT

(section 2.3). More precisely, the two-current model relates p(T) to the parameter
a characteristic of the impurity scattering (eq. (25)). According to whether a is
large or close to unity p(T)-po is large and varying as p , +(T) or small and
varying as pi(T). The values of p(T)-po for pure metals, although scattered, are
relatively small, which suggest that a is generally close to unity. The variation in
T 2 can be then ascribed to pi(T).
The variation in T 2 has been attributed either to electron-magnon scattering or
to s-d electron-electron scattering (Baber 1937) of the same type as leads to a T 2
term in the resistivity of non-magnetic transition metals at low temperatures, and
which is much the same magnitude as the T 2 term in the ferromagnets. If the
thermal conductivity of the ferromagnetic metals is also measured at low tem-
peratures the Lorentz ratio corresponding to the non-impurity scattering is about
1 × 10-8 W~)K -2 both in Ni (White and Tainsh 1967) and in Fe (Beitcham et al.
1970). This is close to the value estimated theoretically (Herring 1967) for s-d
electron-electron scattering. However the way in which the experimental data are
analyzed has been criticized (Farrel and Greig 1969). Secondly, we can consider
data on pit(T), pi+(T) and p~ +(T) obtained from an analysis of dilute Ni alloys
(see section 2.3). The spin-flip magnon-electron resistivity p~ +(T) is roughly
5 × 10 9 l~cm at 10 K in Ni (see fig. 9); because of the small angle scattering factor
the electron-magnon contributions to p~, (T), pi+ (T) should be lower by a factor
of the order of T/Tc, giving pi~ ( m a g n o n s ) - 1 0 - 1 ° ~ c m ; as the "observed" pi~
values in Ni at 10 K are much higher (these are higher than 10 -9 ~~cm), we can
infer that the electron-magnon contributions to the p~r are not dominant. This
actually is in agreement with predictions of calculations based on a realistic band
structure model of Ni (Yamashita et al. 1975). It thus turns out that electron-
electron collisions play the major role in the low temperature T 2 term of the pure
metals.

2.2.3. Residual resistivity


In pure ferromagnetic metals the "internal magnetoresistance" enhances the
resistivity which is no longer proportional to the concentration of impurities. This
effect is particularly important for Fe which has a high value of 47rM and a high
transverse magnetoresistance (fig. 5). For demagnetized Fe polycrystals it was
pointed out that the apparent residual resistivity ratio p(300 K)/p(4.2 K) would
never increase beyond about 300 however pure the sample (Berger and de
Vroomen 1965). It is now standard practice to measure the low temperature
resistivity of Fe samples in a saturating longitudinal magnetic field so as to
eliminate transverse magnetoresistance. This can reduce the apparent resistivity
by a factor of 5 or more. In principle a correction should still be made for the
longitudinal magnetoresistance. In Ni samples the enhancement of the residual
resistivity by the internal magnetoresistance is less important than in Fe but still
significant (Fujii 1970, Schwerer and Silcox 1970).
The contribution of the domain walls to the residual resistivity of ferromagnetic
metals has been subject to many discussions. It now appears that domain walls are
too thick to scatter electrons appreciably. However, as it has been pointed out in
TRANSPORT PROPERTIES OF FERROMAGNETS 765

1000 -

x
500
B nO
0
o xrl
x x O 0 O
" O V ^Jm O0 A
c~ 200 n
.4.. On

~" 0
100 - - o
c,l
0... x[3
X
50 x

°o
0 *o Increasing p u r i t y
20
x I , i i , ~,,I I , i I L,,,I ,
20 50 100 200 500 1000 2000
((3295 / (34.2)mox.

Fig. 5. Residual resistance ratio in zero field and in longitudinal applied field for iron samples of
increasing purity (after Berger 1978).

the preceding section, the resistivity depends indirectly on the domain walls as it
depends on the domain configuration in the sample (Berger 1978).

2.2.4. High field behaviour


It is well known that the magnetoresistance and Hall effect of pure metals under
high fields such that wc~">> 1 give information on the Fermi surface.
In Ni under high fields applied along a non-symmetry direction of a monocrys-
tal, the transverse magnetoresistance saturates, and R0 corresponds to an effective
carrier density of 1 electron per atom, even though Ni has an even number of
electrons and so would normally be expected to behave as a compensated metal.
Reed and Fawcett (1964a) showed that a ferromagnetic metal did not have to obey
the same rules as non-magnetic metals because of the inequivalence of spin 1' and
spin $. They deduced from their results that the minority d band in Ni was
electron-like in character. The behaviour of the magnetoresistance for certain
field directions indicated the presence of open orbits for certain field orientations.
The results could be compared with de Haas-Van Alphen data (Hodges et al.
1967, Tsui 1967, Ruvalds and Falicov 1968). No obvious transition corresponding
to a major difference in mobility for d-like and s-like parts of the Fermi surface
was observed.
In Co the transverse magnetoresistance is again saturated (Coleman et al. 1973)
with open orbit behaviour for certain special directions.
In Fe up to fields of about 100kG, the magnetoresistance tends to a B 2
dependence indicating that the metal is compensated (Reed and Fawcett 1964b).
There appears to be a considerable spread of o)c~-values. In the same field range
at low temperatures the ordinary Hall coefficient R0 is strong, negative and
766 I.A. C A M P B E L L A N D A. F E R T

weakly field dependent (Klaffky and Coleman 1974). This also is consistent with a
metal having compensated character, for which R0 bears no relationship to any
effective number of electrons per atom.
At still higher values of ~o~- the magnetoresistance increases much more slowly
than B 2 (Coleman 1976); magnetic breakdown and intersheet scattering have been
invoked.

2._3. Alloys: residual resistivity and temperature dependence of resistivity

The residual resistivity per atomic percent impurity has been measured for a wide
range of impurities in Ni, Co and Fe (p0 in tables 1, 2, 3) and the deviations from
Matthiessens' rule have been studied both for ternary alloys (fig. 6) and for binary
alloys as a function of temperature. Using the two current model equations of
section 2.1.2 the experimental data have been used to determine the spin t and
spin ~, residual resistivities Pot and p0~ for each impurity in each host (tables 1, 2,
3).

TABLE 1
Values of a = P o l / P o t , po, Pot, Po,, for dilute impurities in nickel*.

Po po t Po
Impurity in nickel c~ = Po ~/Po t (ixf~cm) (ix~cm) (~f~cm)

Co 13 ("), 300,), 20 (c),


13(d) 20(f), 20(g) 0.16 ± 0.03 0.18 --+0.03 3.5 --+1

Fe 11 ("), 200,), 7.3 (d) 0.36-+0.04 0.4-+0.04 6-+ 1.5

Mn 6.3 ("/, 150,), 5.4 (d) 0.7 -+ 0.1 0.75 -+ 0.2 7.5 + 2.5
Cr 0.21 ("), 0.45 (b), 0.4 (c),
0.21(a), 0.2#), 0.4(g) 5 -+ 0.1 22 -+ 6 6.5 -+ 0.5

V 0.45 ("), 0.55 °,), 2.3 (d) 4.4-+0.2 13-+ 1 6.7-+0.5

Ti 0.9 (a), 40,), 2.7 (d) 3.3 -+ 0.6 5.6 -+ 2 10.5 -+ 4

Pd 1(d) 0.2 -+ 0.05 0.3 Id) 0.3 (a)

Rh 0.3 ("), 0.17 ("), 0.290) 1.8 _+ 0.1 10 _+ 2 2.1 _+0.2

Ru 0.075 ("/, 0.15 (e) 4.8 _+ 0.2 56 -+ 15 5.8 -+ 0.5


Mo 0.28 e), 0.37 °) 6.4 -+ 0.6 25 + 4 8+ 1
Nb 0.44 ("/, 0.470) 5 _+0.2 16 -+ 1 7 _+ 0.2
Zr 7.5 (e) 2.8 ± 0.5 4 (e) 30 (e)

Pt 0.24 (a), 0.17 (~) 0.85 ± 0.2 5.3 ± 1.6 1 _+0.2


Ir 0.24 ("), 0.13 (~) 3.8 ± 0.2 28 ± 7 4.8 ± 0.2
Os 0.13 ("~, 0.13 o) 5.5 _+0.5 50 ± 2 6.4 ± 0.5
Re 0.3 ("), 0.26 e) 5.8 ± 0.5 26 ± 3 7.5 ± 0.5
T A B L E 1 (continued)

Po pot Po ,~
Impurity in nickel a = po ~/Po t (txf~cm) (ixf~cm) (pf~cm)

W 0.4 (e), 0.5 o) 6 ±0.5 16.5 -+ 1 7 -+ 0.5

Ta 0.53 (e), 0.46 o) 5.2 ± 0.5 16 ± 1 7.5 ± 0.5


Hf 8.6 (~), 8.1 o) 3.6 ± 0.5 3.5 ± 0.5 30 - 1

Cu 2.9 (a), 3.7 (d) 0.9 ± 0.1 1.1 ± 0.2 3.7 ± 0.2
Au 5.9 (") 0.36 (") 0.44 (a) 2.6 (a)
AI 1.7 (a) 2.13 (a) 3.4 (a) 5.8 (a)
Si 1.3 (") 2.83 °) 5 (a) 6.4 (a)
Zn 2.2 (~) 1 ± 0.1 1.3 (a) 2.9 (a)

Ga 1.7(g) 1.91 (g) 3cg) 5.2 (g)


Ge 1(g) 2.84 (g/ 5.7 (g/ 5.7 (g)
In 1.50') 3.60') 6 °`) 90`)

Sn 1.6(a), 1.350') 3.2 ± 0.4 5.2 -+ 0.8 7.7 ± 0.5


Sb 0.8 ~) 1.6c~) 3.60') 2.90')

* For a we give the values found by:


(a) Dorleijn and M i e d e m a (1975a), Dorleijn (1976); (0 Cadeville et al. (1968);
0`) Fert and Campbell (1976); ~) Hugel (1973);
(c) Leonard et al. (1969); 0') Ross et al. (1978);
ca) Farrell and Greig (1969); (i) D u r a n d (1973).
(e) D u r a n d and Gautier (1970);
T h e values of a given in (a), (b), (f), (g) have been mostly derived from m e a s u r e m e n t s of the
residual resistivity of ternary alloys, which is the most direct method. T h e values of a given in (d), (e),
(h), (i) have been obtained on binary alloys from the deviations from Matthiessen's rule at low
temperature (h), 77 K (i), 300 K with the assumption of complete spin mixing (d) or 300 K with the

lo]
assumption of no spin mixing (e). T h e values that we give for p0, Pot, p0~ have been estimated from
the spread of the values found in the literature.

N.~ (AUl_xCO x)
o_£ 1 0.5[
t // Ni (Co1_xRhx)
i 01
0 05 10 0 015 1.0
× X
Fig. 6. Residual resistivities of Ni(COl-xRhx) and Ni(Aul xCox) alloys. The large deviations from
Matthiessen's rule (broken line) for the Ni(COl-xRhx) alloys are accounted for by very different values
of C~coand aRh; the solid curve is calculated from eq. (24) with aco = 13 and aRh = 0.3. T h e very small
deviations from M R in the Ni(Aul-xCox) are associated with values of aAu and aCo both m u c h larger
than 1 (after Dorleijn 1976).

767
768 I.A. C A M P B E L L A N D A. F E R T

TABLE 2
Values of a = Po ~/po ~, po, Po t , po ~ for dilute impurities in cobalt*.

Po Po ~ po;
Impurity in cobalt c~ = Po +/Po ~ 0xf~cm) (Ixf~cm) (~ftcm)

Fe (") 12 0.5 0.54 6,7


Mn c°/ 0.8 5.5 12 10
Cr e°} 0,3 1.8 7.3 2.4
V (b) 1 3.8 7.7 7.7
Ti (u~ 1.4 4.5 7.6 11
Rh c°~ 1 1.4 2.8 2.8
Ru ("/ 0.22 4.0 22.4 4,86
Mo c°~ 0.7 6.0 14.4 10
Nb ~ 1 6.5 13 13
Zr c°~ 3.3 4.0 5.2 17
Ir (a~ 0.33 2.9 11.7 3.82
Os (a~ 0.29 5,3 23.5 6.84
Re c°) 0.43 5.3 18 7.7
W (b~ 0.84 5.7 10.5 12.5
Ta c°~ 1.23 5.5 10 12.3
Hf (b) 2.5 4.0 5.5 14
Sn (c~ 1.2 2.9 5.3 6.4
Sb (c) 0.9 2 4.2 3.8

• ca)Loegel and Gautier 1971; ~b~Durand 1973; to) Ross et al. 1978. The data
have been obtained from deviations from Matthiessen's rule in the residual
resistivity of ternary alloy (a) or in the resistivity of binary alloys at low
temperature (c) or at 77 K (b). We have preferred the results given by Durand
(1973) to slightly different ones given previously by Durand and Gautier
(i970).

2.3.1. Nickel host


The general picture of c~ (=p0~/p0t) values for impurities in Ni estimated by
different groups is consistent, although numerical values are not in perfect
agreement (table 1, fig. 7). It is found that Co, Fe, Mn, Au and Cu have a >> 1
while Cr, V and a number of other transition impurities have a < 1. As has been
pointed out (Durand and Gautier 1970, Fert and Campbell 1971, 1976, Hagakawa
and Yamashita 1975) there is a very clear connection between the electrical and
the magnetic properties of the impurities. Those impurities with high values of o~
are those which, on the Friedel analysis of the magnetic properties (Friedel 1967),
do not have d I' virtual bound states at or near the spin 1' Fermi energy. These
impurities have low P0t values because the d 1" phase shift at the Fermi energy is
small; in contrast, when the impurity is such that a d ]' virtual bound state is close
to the Fermi energy, Pot is large so a is small. The difference in Pot values
between these two types of impurities can be quite striking: Pot -~ 0.16 Ixf~cm/%
for Co impurities while Pot = 5 6 1 x ~ c m / % for Ru impurities! Detailed com-
parisons between calculated and experimental values of spin 1' and spin
resistivities have been made.
The temperature dependence of the resistivity of binary alloys of Ni can be
TRANSPORT PROPERTIES OF FERROMAGNETS 769

TABLE 3
Values of a = PoUPot, Po, Pot, P01 for dilute impurities in Fe*.

Po Po i' Po $
I m p u r i t y in Fe oL = po ~ Ipo t (Ixl~cm) (ixf~cm) " (Ixf~cm)

Ni 3 (a), 70") 2 _+0.2 2.4 _+0.2 12 _+5


Co 1("), 3.70') 0.9_+0.1 " 1.6_+ 0.4 3.3_+ 1.3
Mn 0.09 (a), 0.170') 1.5 _+0.2 13 _+5 1.7 _+ 0.2
Cr 0.17 ("), 0.37 (8) 2.2_+0.3 12.5_+6 2.8_+0.2
V 0.12 C"), 0.13 °') 1.1_+0.3 10.5_+3 1.3_+0.3
Ti 0.25 (a), 0.66 (8) 2.75 _+0.25 10.5 _+4 4 _+0.4
Rh 5.8 (8) 0.95 ~ 1.1 °') 6.4 0')
Ru 0.380') 2 _+0.1 7.3 °') 2.80')
Mo 0.210') 1.75 ~ 0.2 110') 2.30')
Pt 80') 1.3 (8) 1.5 °') 120')
Ir 90') 2 (8) 2.2 °') 200')
Os 0.33 °') 3.5 _+0.5 130') 4.3 °')
Re 0.31 c°) 2.7 _+0.5 8.7 °') 2.7 (8)
W 0.240') 1.6 _+0.1 7.50') 1.8 °')
Be 6.2 °') 40') 4.7 °') 29 °')
A1 8.6 °') 5.3 _+0.2 5.6 °') 48 °')
Si 5.60') 6 _+0.6 6.40') 36 °')
Ga 8.1 (8) 4.8 °') 5.40') 440')
Ge 6.2 °') 6.8 _+0.2 7.9 (8) 49 °')
Sn ~ 1(c) 8.7 _+ 1
Sb ~ 1(~) 9.8 _+0.4

* F o r a we give the values derived by: (~) Fert and Campbell (1976); 0") Dorleijn
and M i e d e m a (1977), Dorleijn (1976); (C)Ross et al. 1979, f r o m the residual
resistivity of ternary alloys (b) or f r o m p ( T ) of binary alloys (a) and (c). W h e n
there are data f r o m several a u t h o r s we have estimated m e a n values of p0, p0t,
Po~.

ID

, J,
E 20.

15

0_.-
~_~ 10
c~_

Z
I I ! I
Ti V Cr iqn Fe Co Ni
Fig. 7. S u b - b a n d residual resistivities p0 t and p0 ~ of 3d impurities in nickel. (References in f o o t n o t e to
t a b l e 1.)
770 I.A. C A M P B E L L A N D A. F E R T

analyzed by using the two-current model equations to estimate the pure metal
parameters Pt ~(T), Pit (T) and pi,(T). Figure 8 shows the agreement between
experimental results below 50 K for series of Ni alloys and curves obtained from
eq. (25) by using a values derived from independent measurements on ternary
alloys, /x = 3.6, pi(T) = 9.5 x 10-12T2+ 1.7x 1 0 - 1 4 T 4 (in f~cm if T is expressed in
K) and Pt +(T) of fig. 9 (dashed line). At temperatures up to about 50 K the
analysis can be done unambiguously but at higher temperatures different sets of
solutions fitting the experimental data are possible. At 300 K a reasonable
estimate is Pt ~(300) = 11 tx~cm, pit (300) = 6.7 tM2cm, pi~(300) = 27 FxlIcm (Fert
and Campbell 1976).
The contribution to Pt ~(T3 from electron-magnon collisions has been cal-
culated by Fert ~(1969) and Mills et al. (1971) in a model of spin-split spherical
Fermi surfaces. The calculation gives the correct order of magnitude. The
variation obtained for p~ $(T)/T 2 as a function of T is shown in fig. 9 (solid line)
together with the variation needed to fit the experimental results (dashed line).
The calculated curve drops at low temperature, which results from a freezing out
of electron-magnon, scattering in the presence of a gap between spin I' and spin
Fermi surfaces; the experimental curve shows a similar drop below about 30 K
and then an upturn below 5 K; this upturn seems to be associated to a variation in
T 3/2 at very low temperature and has been ascribed to electron-magnon scattering

~ T (10-115~cm °K-2)

x × x

{Cx'C °
10 x o

/~"/~:atc. _. / o
/* /NiMnOA*/, d
,¢// o/O
? ,//.x./ /
x x r o cole. a

,,x x ~ / x7 7 - o ° "~
/_~a_NiCrt6°lo~

4
"'-u tx
~,1° ~ • cole. _ %M.n0
t~

10 20 30 40 50
Fig. 8. pr/T 2= Co(T)-p(O))/T 2 against T for several nickel based alloys. The solid curves are
calculated from eq. (25) in the way described in the text (after Fert and Campbell 1976).
TRANSPORT PROPERTIES OF FERROMAGNETS 771

in regions where the spin 1' and spin $ Fermi surfaces touch or are very near
(Fert and Campbell 1976).
The resistivities pi~(T) are expected to include contributions from electron-
electron, electron-phonon and electron-magnon collisions. Because of the small
angle scattering factor the electron-magnon contributions to pi~, pi+ should each
be equal to roughly (T/Tc)p~ + and therefore relatively small at low temperatures.
If then the electron-electron or electron-phonon contributions are dominant, the
possibility of scattering of the spin $ electrons to the d $ band makes pi$(T)
larger than pi~(T), in agreement with /~ > 1. At very low temperatures the
variation of pi(T) in T 2 can be attributed to electron-electron scattering, as it has
been concluded in section 2.2.2. Above 10K the electron-phonon collisions
become progressively more important. When approaching room temperature the
electron-magnon collisions should begin to make a substantial contribution to
pi~(T). Theoretical estimates of the electron-phonon contributions to Pit and pi+
at 300 K are 4.25 ~ c m and 19.2 ~l"~cm respectively (Yamashita and Hayakawa
1976); we can reasonably infer that additional contributions of a few ~ c m from
electron-magnon scattering account for the experimental pi~(300). Without mag-
non contributions to pi~(300) and without p~ ~ term the resistivity of pure nickel at
3 0 0 K would be predicted to amount to roughly 4.25x 19.2/(4.25+ 1 9 . 2 ) -
3.5 ~ c m , instead of 7 p ~ c m experimentally. We conclude that:
(i) at low temperature the main contributions to p~(T) arise from electron-
electron and electron-phonon scatterings; electron-magnon collisions come into
play through p~ ~(T) and are important in alloys with ~ very different from unity;
(ii) at near room temperature the electron-magnon collisions contribute to
both pi~ and p~ +; they will become increasingly important as temperature
increases.
The analysis of the experimental data on Ni alloys by Yamashita and Hayakawa
(1976), although based on a different treatment of the two current conduction,
arrives at similar conclusions.

2.3.2. Cobalt host


The o~ values of a large number of impurities have been obtained in Co metal
(Durand and Gautier 1970, Loegel and Gautier 1971, Durand 1973, Ross et al.
1978), table 2. They are again consistent with the magnetic structures of the
impurities. The parameters Pit (T), pi+ (T) and p, +(T) of Co have been evaluated
by Loegel and Gautier (1971); the behaviour of p~ ~(T) is similar to that of Ni.

2.3.3. Iron host


Extensive work has been done on Fe based alloys (Campbell et al. 1967, Fert and
Campbell 1976, Dorleijn 1976, Dorleijn and Miedema 1977, Ross et al. 1979),
table 3. The resulting ~ values from different authors, both from ternary alloy
data or from temperature dependence, are in reasonable agreement with each
other. The range of c~ values is very great, po ~/po ~ varies from 0.13 for F__eeVto 9 for
FeIr (table 3). There is a good correlation between the resistivity p0 in each band
and the charge screening in that band for each impurity (Dorleijn 1976).
772 I.A. C A M P B E L L A N D A. F E R T

,_NC
12

T
C)
v

k-

-4P

J I I I I I
10 20 30 40 50

Temp6rature (K)

Fig. 9. Experimental (dashed line) and calculated (solid line) curves for p$ ,~/T 2 in nickel. The
experimental curve is after Fert and Campbell (1976); the calculated curve is obtained from the model
calculation of Fert (1969) with 01 = 38 K.

The behaviour of p, +(T) in Fe is similar to that in Ni and the value of pi+/pi,


seems to be near 1 (Fert and Campbell 1976). More complete low temperature
measurements would be necessary to decide this. As in Ni, the low temperature
p(T) data cannot be understood without including the p, ~ term.

2.3.4. Alloys containing interstitial impurities


Ni, Co or Fe based alloys containing small concentrations of interstitial impurities
of B or C can be prepared by rapid quenching. Swartz (1971), Schwerer (1972) and
Cadeville and Lerner (1976) have investigated the resistivity of NiC, C__o_oC,
Nil_xF_eexC alloys. The residual resistivity of these alloys is equal to about
3.4 ~flcm/at.% for C in Ni and 6.6 txOcm/at.% for C in Co. From the deviations
from Matthiessens' rule in N__iiCrCand CoCuC alloys, Cadeville and Lerner (1976)
have estimated that the resistivity P0+ was about twice as large as Pot. This result,
together with magnetization and thermo-electric data by the same authors, are
consistent with a predominant screening by the electrons of the d + band. In
Ni~-xFexC alloys the resistivities Pot and p0+ of the C impurities are found to
become nearly equal for x >0.4, which has been ascribed to the change from
strong to weak ferromagnetism (Cadeville and Lerner 1976).
The resistivity of B impurities in Ni and Co have been found to be fairly small
( - 1 ixl)cm/at.%). This has been ascribed to a predominant screening by the d ~,
electrons resulting in a small resistivity for the spin 1' electrons (Cadeville and
Lerner 1976).
TRANSPORT PROPERTIES OF FERROMAGNETS 773

2.4. High temperature and critical point behaviour

It was observed a long time ago that the resistivities of ferromagnetic metals
changed slope as a function of temperature at the Curie temperature. For Ni this
was originally interpreted by Mott (1936) as indicating a reduction of the spin T
resistivity on ordering. Later work (Kasuya 1956, Yoshida 1957, Coles 1958, Weiss
and Marotta 1959) showed that spin disorder scattering provided a more general
explanation. When the resistivities of the 3d ferromagnetic metals are compared
with those of their non-magnetic 4d and 5d counterparts it can be seen clearly that
there is an extra magnetic scattering contribution which is approximately constant
above T~ and which decreases gradually below T~ (fig. 10). The simplest disorder
model shows that the paramagnetic term above T~ is equal to

kv(mF)2 t t r
Pm = 4~e2zfi3 ~ + 1), (26)

where J is the effective local spin and f ' the local spin conduction electron spin
coupling parameter. De Gennes and Friedel (1958) suggested that the critical
magnetic scattering near Tc was similar in type to the critical scattering of
neutrons and that it should lead to a peak in p(T) at To. Later work by Fisher and

I0[
[3 Tc
Qcm ~/
80

6C

Tc

Pd
20

0
~ I r I L I
T//80 I
0 1 2 3 /~
Fig. 10. Resistivity of several transition metals as a function of T/OD. OD is the D e b y e t e m p e r a t u r e .
774 I.A. CAMPBELL AND A. FERT

Langer (1968), using a better approximation for the spin-spin correlation function
near To, modified this prediction to that of a peak in dp/dT at To. They also made
the important remark that just above To the same leading term in the spin-spin
correlation should dominate dp/dT and the magnetic specific heat, so that these
two parameters should have the same critical behaviour as T tends to Tc from
above. Both magnetic entropy S and magnetic scattering rate should be propor-
tional to

fo2kvF(k, T)k 3dk, (27)

where F(k, T) is the spin-spin correlation function. Later theoretical work


showed that the same correspondence should hold equally in the region just
below Tc (Richard and Geldart 1973).
Renormalization theory can predict the critical coefficients for dp/dT (Fisher
and Aharony, 1973) but it is difficult to decide over what range of temperature
each side of To the strictly "critical behaviour" should be observed; Geldart and
Richard (1975) discussed the cross-over from a regime near To where the short-
range correlations dominate to a long-range correlation regime. The theory of
resistivity behaviour at To in weak ferromagnets has been developed by Ueda and
Moriya (1975), Der Ruenn Su and Wu (1975).
Experimentally, the critical behaviour of dp/dT has been studied very carefully
for Ni, Fe, Gd and the compound GdNi2 (Craig et al. 1967, Zumsteg and Parks
1970, Shaklette 1974, Kawatra et al. 1970, Zumsteg and Parks 1971, Parks 1972,
Zumsteg et al. 1970). For Ni (Zumsteg and Parks 1970) and Fe (Shaklette 1974) it
is found that dp/dT and the specific heat do indeed show the same A point type of
behaviour around To (fig. 11). The data are parameterized using

I J i i i l l ,

ooo o°
oo 1.04
.0.05 o
o
o

oo
o
oo
1.03
oO o
o
.OOl o
o 1.02
o
o o

1.01 -~
.005
1,00 n,-

099
.002 o°°2 n*"

o
o 0.98
,001 o
o
0.97
o

096
.000 ~ , , , I i I I I I I I I

348 352 356 360 T(oC)36& 368 372 376

Fig. 11. Resistivity R(T) of nickel and dR/dT versus temperature in the region of the Curie point
(after Zumsteg and Parks 1970).
TRANSPORT PROPERTIES OF FERROMAGNETS 775

1 dp A
pcdT-h (e-*-l)+B, T > To, (28)

and

1 dp A'
p o d T - -h (lel-~'- 1 ) + B " T < To, (29)

where

e = ( T - Tc)lTc. (30)

Renormalization theory predicts h = h ' ~ 0 . 1 0 and A/A'~-1.3 (Zumsteg et al.


1970) for a 3 dimensional exchange ferromagnet. The accurate determination of A
and h' is extremely delicate especially as Tc must be fitted self-consistently from
the data and it appears essential to have the theoretical predictions as a guide.
In pure Fe, Kraftmakher and Pinegina (1974) find h, h ' = 0-+ 0.1 while Shaklette
(1974) observes A, A'=-0.12--_0.01 by imposing h----h'. Agreement with the
magnetic specific heat data in Fe is very good (Shaklette 1974, Connel!y et al.
1971). For Ni, the values obtained were h = 0.1 _+0.1, h' = 0.3_+0.1 (Zumsteg and
Parks 1970) but within the fitting accuracy this is presumably also consistent with
theoretical values.
In G d which is hexagonal the critical behaviour looks very different when
measured along the c- and the a-axes. Zumsteg et al. (1970) suggest that the
resistivity changes are complicated by the critical behaviour of the lattice
parameters, but this has been questioned (Geldart and Richard 1975).
GdNi2 was investigated in the hope that it would correspond to a simple local
moment system, dp/dT shows similar critical behaviour to Fe and Ni but has more
complicated temperature dependence a few degrees above Tc (Kawatra et al.
1970, Zumsteg and Parks 1971). The significance of this has been discussed
(Geldart and Richard 1975). The resistivity variation has also been measured at
the structural and ferromagnetic transition in T b Z n (Sousa et al. 1979).
The critical behavi0ur of dp/dH has been studied for Ni (Schwerer 1974) and
for Gd with the current in the basal plane (Simons and Salomon 1974).
The behaviour of transport properties near Tc can also be studied in alloys, but
local inhomogeneity leads to a spread in the local values of Tc at different parts of
the sample and so the critical behaviour is smeared out. This has been observed in
NiCu alloys (Sousa et al. 1975) and in PdFe (Kawatra et al. 1970, Kawatra et al.
1969).
Finally, behaviour at the critical concentration for ferromagnetism (the concen-
tration at which To-> 0) can be studied. Very varied behaviour has been found in
different alloy systems. In N iCu alloys there is a peak in dp/dT at T~ as long as Tc
exists and there is a maximum in p(T) some degrees higher, while for c > cent a
minimum in p(T) is observed (Houghton and Sarachik 1970). In NiAu (splat
cooled to avoid segregation) p(T) shows a maximum at T0 for c < cent (Tyler et al.
776 I.A. CAMPBELL AND A. FERT

1973). For NiCr, Yao et al. (1975) find weak minima in p ( T ) for c > Ccrit while
Smith et al. find giant minima in the region c - cent (Smith et al. 1975). In NiPd
alloys, Tari and Coles (1971) express the low temperature resistivity behaviour as
p = po = A T n and find A is sharply peaked at cc~t while n has a minimum with
n - 1. The Curie point "is not easy to detect on the p ( T ) curves". A m a m o u et al.
(1975) using the same way of expressing the resistivity behaviour found n --> 1 and
strong peaks in A at the critical concentrations of a large number of alloys
systems.
The transition from low temperature two current behaviour to high tem-
perature spin disorder behaviour has been studied in Fe based alloys (Schwerer
and Cuddy 1970). The high temperature resistivity behaviour of the alloy seems to
depend essentially on the local impurity moment.

3. Other transport properties of Ni, Co, Fe and their alloys

Here we will summarize results on different transport properties in these metals


and alloys and outline the interpretations which have been given. We will
generally find that Ni has been studied in most detail while rather less is known
about Fe and Co. In the interpretation of the results, we will refer to what has
been learnt about the different systems from the resistivity measurements which
we have already discussed.

3.1. Ordinary magnetoresistance

We have outlined the situation for pure metals in section 2.2. For non-magnetic
alloys the low temperature magnetoresistance behaviour generally follows Koh-
ler's rule ( p ( B ) - p ( O ) ) / p ( O ) = f(B/p(O)), where f is a function which varies from
metal to metal but which is rather insensitive to the type of impurity present for a
given host. In a ferromagnet above technical saturation the same effect, due to the
Lorentz force on the electrons, can be observed but as B includes the mag-
netization term 47rM, p(0) cannot be attained except by extrapolation. Schwerer
and Silcox (1970) showed by a careful study of dilute Ni alloy samples that for a
given series of alloys (e.g. NiFe samples) the ordinary magnetoresistance follows a
Kohler's rule, but that the Kohler function f varied considerably with the type of
scatterer (fig. 12). Other work (Fert et al. 1970, Dorleijn 1976) is consistent with
these data.
It can be seen in fig. 12 that the strongest magnetoresistances are associated
with impurities having large values of p0+/P0t (i.e. N__iiFe, __NiCo...). The lon-
gitudinal magnetoresistance of these alloys is also high [Apll/P(O ) saturates at about
10 in NiFe (Schwerer and Silcox 1970)] considerably greater than that observed
for Cu based alloys for instance, where ApJp(O) saturates at about 0.7 (Clark and
Powell 1968). Attempts have been made to understand this behaviour in the two
current model. In its simplest form the two types of electron (spin 1' and spin $ )
can be represented by electron-like spheres in k space with different relaxation
TRANSPORT PROPERTIES OF FERROMAGNETS 777

,.,o oo////

1.05

IDO ~ ~ R u ~ , , ,
0 10 20 30 &O 50
B/~O (k G/,u.~.cm)
Fig. 12. Kohler plots for the transverse magnetoresistance at 4.2 K of nickel containing Co, Fe, Mn,
Ti, A1, Cr, Pt, V or Ru impurities (after Dorleijn 1976).

times. In this approximation (Fert et al. 1970) the transverse magnetoresistance is


indeed an increasing function of p~ (O)/pt (0), but the model is not satisfactory as it
predicts a zero longitudinal magnetoresistance in disagreement with experiment.
As a next step, it is possible to invoke relaxation time anisotropy within each spin
band. Dorleijn (1976) suggests that the intrinsic magnetoresistivity of the spin 1'
band of Ni is much greater than that of the spin band so that the longitudinal and
transverse magnetoresistances are much greater when the current is carried
mainly by the spin 1' electrons. Jaoul (1974) proposes that there is a mixing
between spin 1' and spin ~ currents which is an increasing function of B/p(O).
This is because spin-orbit effects mean that an electron on a given orbit on the
Fermi surface passes continuously between spin 1' and spin +, progressively
mixing currents as B/p(O) increases. This model predicts the saturation mag-
netoresistances of the different alloy series reasonably well.
The ordinary magnetoresistance in Fe based alloys is m o r e difficult to express in
the form of Kohler curves, because the much higher value of 4 ~ M in Fe means
that extrapolations to B = 0 are always very extended. D a t a given by Dorleijn
(1976) again indicate different Kohler curves for Fe samples containing different
impurities, but the correlation with the value of p+ (O)/p, (0) is much less clear than
in the case of Ni based alloys.
There is an additional effect that appears under similar experimental conditions
as the Lorentz force ordinary magnetoresistance, but which is due to the high field
susceptibility of the ferromagnetic metal. This high field susceptibility can have
two origins. First, there is an increasing magnetic order in an applied field which
can also be thought of as a reduction in the n u m b e r of magnons with increasing
field. This term is m a x i m u m around Tc and goes to zero as T goes to zero.
Secondly, for a band ferromagnet, the local magnetic m o m e n t s can be altered by
an applied field at any temperature, even T = 0 (Van Elst 1959).
778 I.A. CAMPBELL AND A. FERT

Insofar, as an increasing field produces increasing magnetic order and hence


lower spin disorder scattering, dp/dH due to the first term will be negative. The
second type of effect can in principle give either positive or negative mag-
netoresistance depending on the electronic structure of the system. Van Elst
(1959) measured at 300 K (1/p)(dp/dH)l I~- (1/p)(dp/dH)l with effects of the order
of 10-4/kG and with significant variations from one alloy to another. This
behaviour is due to the first effect. At low temperatures the Lorentz-force
magnetoresistance dominated except for NiMn alloys which showed negative
dp/dH even at low temperature; this is probably due to an unusual band
susceptibility in these alloys.

3.2. Ordinary Hall coefficient

In non-magnetic metals it is known that the ordinary Hall coefficient R0 behaves


to a rough approximation as Ro oc 1/en* where n* is the effective density of
current carriers and e is their charge (e is negative for electron-like carriers and
positive for hole-like carriers). The actual values of R0 can be considerably
modified by Fermi surface and scattering anisotropy effects (Hurd 1972); for the
high field condition wc >> 1, R0 depends only on the Fermi surface geometry and
can be highly anisotropic in single crystals.
In ferromagnetic metals the ordinary Hall effect can be separated from the
extraordinary Hall effect by measurements above technical saturation, as long as
the susceptibility of the sample in high fields is negligible so that there is no
paramagnetic extraordinary Hall effect correction (see section 3.4).
The ordinary Hall coefficient in Ni at room temperature is R0-~
- 6 x 10 1312cm/G (Lavine 1961), which corresponds to conduction by electron-
like carriers with an effective electron density n* of about 1 electron per atom. R0
varies by about 20% between room temperature and 50 K; at lower temperatures
the low field condition ~0c~-'~ 1 no longer holds for high purity Ni samples so R0
tends towards the high field value (Reed and Fawcett 1964).
Pugh and coworkers (Pugh et al. 1955, Sandford et al. 1961, Ehrlich et al. 1964)
and Smit (1955) showed that for a number of Ni based alloys, in particular NiFe
and NiCu, the low temperature Hall coefficients in concentrated samples cor-
respond to much lower effective carrier concentrations, n * - 0 . 3 electrons per
atom. They pointed out that this low number of carriers was probably, associated
with a regime where only the conduction band for one direction of spin was
carrying the current. Later work on Ni and NiCu alloys (Dutta Roy and Subrah-
manyam 1969) showed that R0 is very temperature dependent in the alloys, and
that above the Curie point n* returns to a value of about 1 electron-atom, i.e., to
a situation where both spin directions carry current.
This would seem to fit in well with other data on the two current model.
However, careful measurements by Huguenin and Rivier (1965) and by Miedema
and Dorleijn (1977) on a wide range of Ni based alloys have shown that the
situation is more complicated. The data can be summarized as follows: the low
temperature R0 is very close to zero in dilute alloys (concentration - 0 . 5 % ) for
TRANSPORT PROPERTIES OF FERROMAGNETS 779

which Po+/Pot > 1 (i.e. NiFe, N__iiCu, N i C o . . . ) but then increases rapidly with
impurity concentration to a value corresponding to n* - 0.3 in samples where the
impurity resistivity is greater than about 5 txf~cm. For alloys for which po ~/po t < 1,
R0 is essentially independent of impurity concentration at about - 6 x 10-13 l~cm/G
(note that only samples of this type having p > 2 ix~cm were studied).
Now in a two current model R0 is given by

Ro= p2 Rot/p2 + Ro+/p~ , (31)

where Rot, R0+ are the ordinary Hall coefficients for the two spin directions
taken separately. From the experimental data it can be concluded that R0+ is
reasonably constant, while Rot varies strongly with p~. Dorleijn and Miedema
suggested that the effect is due to a "smudging out" of the details of the spin 1'
Fermi surface of Ni with increasing Pt and they associated this with the observed
changes of the magnetocrystalline anisotropy with alloy concentration (Miedema
and Dorleijn 1977). As we will see in section 3.3, the resistivity anisotropy of the
same alloys changes similarly with impurity concentration until a certain residual
resistivity value is reached. The R0 data suggest that the "smudged out" Fermi
surface situation corresponds more closely to the extreme s-d model with con-
duction entirely by an s t like band containing about 0.3 electrons per atom.
The results on R0 in Fe based alloys are less clear, partly because the separation
into ordinary and extraordinary Hall components is more difficult because of the
large value of 4~-M. Fe has a positive ordinary Hall coefficient, as have the dilute
Fe based alloys except for FeCo (Beitel and Pugh 1958) although R0 for __FeNi
alloys changes sign with temperature and with concentration (Softer et al. 1965).
There appears to be evidence (Carter and Pugh 1966) that alloys for which
pt(O)/p+(O)> 1 such as FeCr, behave similarly to Ni in that R0 is high at low
temperatures and drops considerably at higher temperatures as both spin direc-
tions begin to participate in the conduction.

3.3. Spontaneous resistivity anisotropy

This was defined in section 1 and is a spin orbit effect. The mechanism can vary
from system to system. The simplest case to understand, at least in principle, is
that of dilute rare earth impurities (Fert et al. 1977). Because of the unclosed f
shell, the magnetic rare earths can be regarded as ion-like with a non-spherical
distribution of charge (apart from the spherical ion Gd3+). A conduction electron
plane wave encounters an object with a different cross section depending on
whether it arrives with its k vector parallel or perpendicular to the rare earth
moment, which provides an axis for the non-spherical charge distribution. The
anisotropy of the resistivity is proportional to the electronic quadrupole moment of
the particular rare earth. The theory of this effect has been worked out in detail (Fert
et al. 1977).
In transition metals, the spin orbit coupling is usually a weak perturbation on
the spin magnetization. The lowest order terms leading to a resistivity anisotropy
780 I.A. CAMPBELLAND A. FERT

will be either mixing terms of the type ( L + S - ) 2 o r polarization terms of the type
(LzSz)2. Smit (1951) calculated the resistivity anisotropy to be expected on an s-d
model from the mixing terms acting between spin 1' and spin ,~ d bands. When
data became available for both the anisotropy and the p ~/p t ratios in various Ni
alloys, it was found that there was good agreement between the results and
predictions which could be made using the Smit approach (Campbell et al. 1970).
Agreement is however less good for impurities having a virtual bound d state near
the Fermi surface, and an additional (LzS~) 2 mechanism was suggested for these
cases (Jaoul et al. 1977).
The relative anisotropy of the resistivity (Pll- P±)/P defined in section 1 has been
measured for Ni and a large number of Ni alloys as a function of concentration
and temperature (Smit 1951, Van Elst 1959, Berger and Friedberg 1968, Campbell
et al. 1970, Vasilyev 1970, Campbell 1974, Dedi6 1975, Dorleijn 1976, Dorleijn
and Miedema 1976, Kaul 1977, Jaoul et al. 1977) and for many dilute Fe based
alloys, mainly at He temperature (Dorleijn and Miedema 1976). We will first
discuss the Ni data.
The anisotropy ratio for pure Ni is near +2% from nitrogen temperature up to
room temperature, and then gradually drops as the temperature is increased up to
the Curie point (Smit 1951, Van Elst 1959, Kaul 1977). Below nitrogen tem-
perature the anisotropy is difficult to estimate for pure samples because of the
rapidly increasing ordinary magnetoresistance, but it appears t o remain fairly
constant.
For most dilute N__iiXalloy series the limiting low temperature anisotropy ratio is
relatively concentration independent for a given type of impurity X over a fairly
wide concentration range but the value depends strongly on the type of impurity,
table 4. For NiCo, NiFe and N__iiCu (fig. 13) the anisotropy ratio increases
continuously with concentration up to concentrations corresponding to residual
resistivities of about 2 p~cm. It is a disputed point as to whether the appropriate
characteristic value of the anisotropy ratio for these alloys is the plateau value
(Jaoul et al. 1977) or a value at some lower concentration (Dorleijn and Miedema
1975b, 1976).
When the temperature is increased, the anisotropy ratio of a given sample tends
towards the pure Ni value and finally becomes zero at the Curie point of the alloy
(Vasilyev 1970, Kaul 1977).
There is a clear correlation between the value of a and the low temperature
a n i s o t r o p y ratio (Campbell et al. 1970). Alloys having high values of a N(~Co,
NiFe, N i M n . . . ) have high positive resistivity anisotropies while alloys with c~ ~ 1
have small positive or negative anisotropies. A spin-orbit mixing model originally
suggested by Smit (1951) gives a convincing explanation of the overall variation of
the anisotropy ratio with the value of a. As Ni metal has a fully polarized d band,
there are no d 1' states at the Fermi surface for the conduction electrons to be
scattered to. However because of the spin-orbit mixing by the matrix element
AL+S- some d 1' character is mixed into the d $ band. The resulting weak s ]' to
d $ scattering can be shown to depend strongly on the relative orientation of the k
vector of the s conduction electron and the sample magnetization. This leads to a
resistivity anisotropy of the form
T R A N S P O R T P R O P E R T I E S OF F E R R O M A G N E T S 781

TABLE 4
Anisotropy of the residual resistivity of dilute nickel based alloys*.

Impurity Co Fe Mn Cr V Pd

PJ - P± x 102 20 Ca) 13.6 o') 9.9 °) --0.3500) 0.6 C") 2 C~)


t~
14.8 Cb) 14 Cd) 7.8 Cb) --0.28 (a) O. 15 c°)

28 (c) 19.5 Cc) 9.5 Co) -0.23 Cd)

Impurity Rh Ru Mo Nb Pt Ir

P_JI- P± x 102 0.05 C°) -0.600') 0.1 Ca) 0.15 (~) 0.4 C°) - 1 . 5 2 C~)

0.05 C~) - 0 . 8 2 C~) 0.05 (0) 0.4 C~)

Impurity Re W Cu Au A1 Si Zn Sn

pp!- Pa x 102 - 0 . 5 0 C°) 0.4 Ca) 6.8 C°) 7.5 C°) 4.7 (a) 2.5 (") 5.7 Ca) 3.4 Ca)

- 0 . 4 5 Cc) 0.8 ¢ ) 7.8 Co) 7.9 Co) 3.9 Co) 2.1 co) 4..7 (b) 2.9 Co)

4.6 Co) 2.8 (c) 6.5 ¢) 3.5 Co)

*After: ca)Van Elst 1959, C°)Dorleijn and Miedema 1974, Dorleijn 1976, e)Jaoul et al. 1977,
ca)Schwerer and Silcox 1970. We indicate - when this is possible - the resistivity anisotropy of alloys in
the concentration range where the concentration dependence is weak (see fig. 13). The experimental
data on (P/I- P±)/Pll has been re-expressed in terms of (PH- P±)/P.

aOtf (%)
3C

25

20 I
oo ooo

i
15

10 \

5 x ~\
x ""x %,

x
i "x L
0 io 40 60
Impur i ty concentrat on, at °/o

Fig. 13. Concentration dependence of the resistivity anisotropy at 4.2 K for several nickel based alloys.
AA: NiCo, OC): N iFe, ×: N_iiCu (after Jaoul et al. 1977).
782 I.A. CAMPBELL AND A. FERT

P l l - P . / P = ~(o~ - 1 ) , (32)

where y is a spin-orbit constant which can be estimated to be about 0.01 from the
Ni g factor. This model explains the sign, the magnitude and the general variation
of the anisotropy with a (fig. 14). In addition, it has been shown (Ehrlich et al.
1964, Dorleijn and Miedema 1976, Jaoul et al. 1977) that an analysis of the
anisotropy ratio of ternary alloys can lead to estimates of the individual anisotro-
pies for the spin 1' and spin $ currents and that for alloys with a > 1 the results
are in agreement with the predictions of the Smit mechanism.
However, for a number of alloys of Ni for which c~ < 1, although the resistivity
anisotropies remain small as would be expected from the Smit mechanism, eq.
(32) is not accurately obeyed and the anisotropies of the two spin currents do not
obey the Smit rules (Ehrlich et at. 1955, Jaoul et al. 1977). A further mechanism
needs to be invoked for these systems, which are characterized by virtual bound
states at the spin I' Fermi level. A mechanism has been proposed involving the
)tLzSz spin-orbit interaction on the impurity site, particularly for impurities which
have strong spin-orbit interactions (Jaoul et al. 1977). Dorleijn and Miedema
(1976) pointed out that for most impurities, whatever the value of c~, (Ap/p)t > 1
and (Ap/p)$ < 1 but they did not explain this regularity.
The temperature variation of the anisotropy ratio can also be understood using
the Smit model (Campbell et al. 1970). As phonon and magnon scattering
increases with increasing temperature, the effective value of a for an alloy tends
to approach the pure metal value. Data on NiCu alloys have been analyzed in this

30

l //
20

I0

I
1o 20 3'0
Fig. 14. Resistivity anisotropy of Ni based alloys at 4.2 K as a function of a = P0J,/P0t- The straight
line is Ap/~ = 0.01 (a - 1) (after Jaoul et al. 1977).
TRANSPORT PROPERTIES OF FERROMAGNETS 783

way over a wide temperature and concentration range (Kaul 1977) so as to estimate
pit(T), pi~(T) and p~ ~(T).
High concentration effects in certain alloy series have been interpreted as due
to characteristic changes in the electronic structure with concentration (Campbell
1974).
The resistivity anisotropy of a large number of Fe based alloys has also been
studied (Dorleijn and Miedema 1976). Here, the alloys having p,(O)/p+(O),> 1
have strong positive resistivity anisotropies while those with Pl (O)/p+(0) < 1 have
small anisotropy ratios (table 4). Again, an analysis in terms of the anisotropies of
the spin ]' and spin $ currents has been carried out and the predictions of the
Smit approach seem well borne out (Dorleijn 1976).
As we have seen in section 1 the resistivity anisotropy in cubic ferromagnetic
monocrystals can be expanded in a series of D6ring coefficients k l . . . ks. Once
again, Ni and Ni alloys have been the most studied [pure Ni (Bozorth 1951), Ni
15% Fe (Berger and Friedberg 1968), Ni 1.6% Cr and N_j 3% Fe (Jaoul 1974), N__ii
0.5% Fe, Ni 0.55% Pt and Ni 4% Pd (Dedi6 1975)]. Very roughly the individual ki
coefficients are simply proportional to the average polycrystal anisotropy with the
exception of k3 (table 5). This coefficient may behave differently from the others
because it does not strictly represent an a n i s o t r o p y - i t corresponds to an average
change of the sample resistivity with the moment direction which is independent
of the current direction.

TABLE 5
Magnetoresistance anisotropy in Ni and Ni alloy single crystals. D6ring
coefficients ki are givenin percent. References: (a~D6ring 1938,Co)Berger and
Friedberg 1968, (c~Jaoul 1974, cd~Dedi6 1975.

kl ke k3 k4 k5

Ni, 300 K (a)'(d) -3.4 -5.2 + 1.7


NiFe 15% 4.2 K(b~ 55.0 14.5 -26.3 -37.8 +24.7
NiCr 1% 4,2K(c~ -3.0 -0.3 -1.2 +2.3 0_+0.7
NiPd 4% 4.2 K(d~ 4.0 1.0 -5.5 -4.0 -3.0

There are also measurements of the D6ring coefficients for Fe at room


temperature (Bozorth 1951).
No convincing model has been proposed to explain the monocrystal anisotropy
coefficients which presumably depend on the detailed band structure of the metal.
The fact that the terms which are fourth order in the direction cosines of the
magnetization (k3, k4, k5) are as large as the second order terms (kl, k2) is
remarkable.

3.4. Extraordinary Hall effect

Apart from the resistivity, the property of ferromagnetic metals which has
attracted the greatest theoretical attention is the extraordinary Hall effect, Rs; the
extraordinary Hall voltage is remarkable in being both strong and rapidly varying
784 I.A. CAMPBELL AND A. FERT

with temperature and impurity concentration. The fundamental mechanisms


which are believed to produce this effect were proposed some years ago by Smit
(1955) and Luttinger (1958) but the physical understanding of these effects has
been considerably improved quite recently (Berger 1970, Lyo and Holstein 1972,
Nozibres and Lewiner 1973). We will outline the discussion given byNozibres and
Lewiner (1973); although this theory was developed specifically for semiconduc-
tors the same physics can broadly be used for ferromagnetic metals.
An electron in a band submitted to the spin-orbit interaction acquires an
effective electric dipole moment

p= -Akxs,

where A is a spin-orbit parameter, k is the k vector and s the spin of the electron.
If there were no scattering centres, the effective Hamiltonian would he

Ygen = k 2 / 2 m - e E . (r + p )

(where r is the centre of the electron wave packet) for a metal in a uniform
electric field E. Local scattering potentials give local terms in the Hamiltonian

V ( r ) - A(k × s). v v .

Here, the second term arises from spin-orbit coupling in the lattice. An additional
contribution to A can also arise from a local spin-orbit interaction.
There are two distinct effects:
(a) the scattering matrix elements between plane wave states are expressed as

(k'] V - A(k x s ) . V V I k ) : Vu,[1 - iA(k x k')- s]

(by applying the general commutator rule If(x), kx] ~ i 0 f ( x ) / 0 x to V(r)). This
means that the probability of scattering k ~ k' is not the same as the probability
k'~k because of interference between the spin-orbit term and the potential
scattering. For a weak 3 function potential,

Wkk' = V2[1 + 2A V r r n ( k x k ' ) . s] ,

where n is the density of states at the Fermi level.


This "skew scattering" leads to a Hall current such that the Hall angle ~bn oc A V,
which is independent of scattering centre concentration, but which can be of
either sign, depending on the sign of V.
(b) Now we come to the "side jump" term.
The total Hamiltonian is

= k2/2m + V(r)- eE. r+p • [VV- eE],


TRANSPORT PROPERTIES OF FERROMAGNETS 785

and the total velocity is:

v : / ~ - i[r, gel = k/m + A [ V V - eE] x s + p .

Without scattering, p changes as k increases under the influence of E, and


secondly, the energies of the different k states are altered by the second term in v.
However, when scattering is introduced, both of these currents are exactly
cancelled out in static conditions; the first because

( p ) = - A ( t ~ ) × s --- o ,

and the second because the electron distribution readjusts itself to minimize
energy, and this new distribution automatically has an average velocity per-
pendicular to E equal to zero.
It would thus appear that the spin orbit terms do not lead to any extra current.
But, during each scattering event there is also a "side j u m p " or shift of the centre
of gravity of a scattered wave packet

8r = f 6v d t = - A A k × s

(as 6v --- A V V × s = -A/¢ x s during the scattering event).


Now, there are two side j u m p contributions:
(i) electrons travelling with a c o m p o n e n t of k parallel to E jump sideways on
being scattered; the resultant of these jumps is a current.
(ii) electrons with a c o m p o n e n t of k perpendicular to E gain or lose an energy
- e , 3 r . E on scattering. This shifts the total electron distribution to provide a
second current.
These terms are not cancelled out by any compensating terms. They lead to a
total Hall current of 2ANe2E x (s), which is proportional to the electric field E but
independent of the scattering rate. The definition of R, is Vy/IxMz, where y is the
Hall p r o b e direction, x the current direction and z the m o m e n t direction. Putting
Vy = ply and E = pIx, with the Hall current just given we clearly obtain R, ~ Ap2.
Note that the p a r a m e t e r h represents the rate of change of the spin-orbit dipole
- A k x s with k. This is a band property. However, local spin-orbit interactions on
scattering centres can give an additional contribution to A and complicate the
picture.
We can now turn to the experimental data. The skew scattering term can be
expected to dominate in dilute alloys at low temperatures, and indeed in Ni based
alloys for which p(0) ~< 1 tzf~cm at helium temperatures (fig. 15) it has been shown
that the Hall angle ~bH is independent of impurity concentration, but depends
strongly on the type of impurity (Jaoul 1974, Fert and Jaoul 1972, Dorleijn 1976).
It is possible to define Hall angles for each direction of spin, ~bHt and 4~H~ and
experiments on ternary alloys (Dorleijn 1976) or on the t e m p e r a t u r e dependence
786 I.A. CAMPBELL AND A. FERT

- ? H ('1"1.~. cm)
"XTRAORDINARY HALL RESISTIVITY AT /-,.2°K

10 Cu

5 Mn~Fe

Co
2
0
1 c tat*l,)

-5 Cr
lr Os

Fig. 15. Extraordinary Hall resistivity of several types of Ni based alloys as a function of their impurity
concentration. The data are limited to alloys having a resistivity smaller than about 1 Ixf~cm; in more
concentrated alloys, a side-jump contribution progressively appears and becomes predominant for
p = 10 ixf~cm (see fig. 16) (Jaoul 1974).

of t h e H a l l angle (Jaoul 1974) allow o n e to e s t i m a t e t h e s e two H a l l angles for each


i m p u r i t y . R e s u l t s a r e given in t a b l e 6. T h e v a l u e s of t h e s k e w s c a t t e r i n g H a l l
angles can b e discussed in t e r m s of t h e e l e c t r o n i c s t r u c t u r e of t h e v a r i o u s
i m p u r i t i e s (Fert a n d J a o u l 1972, J a o u l 1974).
F o r s a m p l e s with h i g h e r resistivities (either b e c a u s e of h i g h e r i m p u r i t y c o n c e n -
t r a t i o n o r b e c a u s e t h e y are m e a s u r e d at h i g h e r t e m p e r a t u r e s ) t h e side j u m p t e r m
b e c o m e s i m p o r t a n t . C o n s i d e r i n g only d a t a t a k e n at low t e m p e r a t u r e s , results for a
given alloy series can g e n e r a l l y b e fitted (Jaoul 1974, D o r l e i j n 1976) b y t h e

TABLE 6
Skew scattering Hall effect in dilute Ni based alloys. For each impurity, qSH is the dilute limit Hall
angle in millirad, and ~bHt, ~bH~ are the corresponding spin 1" and spin $ Hall angles. References:
* Dorleijn 1976, *Jaoul 1974.

Impurity Ti V Cr Mn

42H +1.5", -4.5 -3", -2.5* +2.8", +2 t -6.5", -9.5 t


~bH1' -3.4* -4*, -79 - 3*, - f - 10*
qSH~ +5.5* +6", -3* +4", 3t +1.5 t

Impurity Fe Co Cu Ru Rh

~bH --6.2, --10t --6.2*, --10.5' --10", --23t +2.5*, +3 t 0", --4t
~bat -7", -10 t -6", -10 t -14", -24* -4.7", +3 t -1.4", - 3 t
~bH; +6", +10 t +2.5", +7 t +3.5", +10' +3", +3 t +1.3", -5*
TRANSPORT PROPERTIES OF FERROMAGNETS 787

expression

Rs = ap + bp :2, (33)

or alternatively (fig. 16)

c/:,H = ~b° + B p , (34)

if the variation of the magnetization with impurity concentration is neglected. It is


usually assumed that this represents a separation into the skew scattering term
and the side jump term. For most Ni based alloy series, as we have seen the
values of 4~° vary considerably, but the values of B hardly vary from one impurity
to another, with B -~ - i m i l l i r a d / ~ c m . However, for those Ni based alloys with
p+ (O)/pt (0)>> 1, the data as a function of concentration cannot be represented by
eq. (33) unless only a very restricted range of concentration is considered. It is
interesting to note that these particular alloys are those which also show
anomalous R0 and resistivity anisotropy behaviour as a function of concentration.
At room temperature, p in Ni and Ni alloys is always "high" so that the side
jump mechanism can be assumed to dominate. The experimental value of the
ratio R d p 2 increases from the pure Ni value, R s / p 2 ~ 0.1 (~cmG) -1, as a function
of impurity concentration and rapidly saturates at a plateau value of about
0.15 (f~cmG) -~ for a wide range of Ni alloys (K6ster and Gm6hling 1961, K6ster
and R o m e r 1964), (fig. 17). The room temperature R s / p 2 values for the alloys are
close to the values at low temperatures for the same alloys (Dorleijn 1976).
However, for certain alloy systems R s / p 2 measured at room temperature changes
steadily with impurity concentration. Thus for NiFe, R~ changes sign at about
15% Fe (Smit 1955, Kondorskii 1964). Alloys with this concentration of Fe show
low values of R~/p: even if a second high resistivity impurity is introduced (Levine
1961).
In pure Fe and FeSi alloys, R s / p 2 is remarkably constant over a wide range of
concentrations and temperatures (Kooi 1954, Okamoto et al. 1962, where this
ratio remains constant although R~ varies over three decades) (fig. 18). For other
Fe based alloys the ratio generally approaches the pure Fe value at moderate or

-r-
20

10 ./"
EL°
-r 1C
//
t / to
1'0 2~0 ¢0 20
. ~ (p~cm) 9±(#.O_cm)
Fig. 16. The extraordinary Hall. angle at 4.2 K as a function of the residual resistivity of FeA1 and
NiRu alloys (Dorleijn 1976).
788 I.A. C A M P B E L L A N D A. F E R T

Rs/ )2 (~.cm9)._i

o ~
-0.15
o

i
/~ .0
/t

-0.1
o CF
Ru
Mo
oNb
-0.05 • Ti
x V

Concentration, %
I I I I I / •
0 1 2 3 4 5 6
Fig. 17. The ratio Rs/p 2 in Ni and Ni alloys at room temperature (after K6ster et al. 1961 and 1964).

l
o Fe #
A 2.04% Si-Fe /
* 3.83% Si-Fe
xO,

109

g
b.
E
o
- fc~0
E
o

1611

, , ,,I , ,, I
156 lO-5 lo-~
Resistiv'lty ~o (D.cm)
Fig. 18. Log-log plot of R5 against p for Fe and some Fe alloys above nitrogen temperature (after
Okamoto et al. 1962).
TRANSPORT PROPERTIES OF FERROMAGNETS 789

I I I [ A
I

50 A12.7% Cr IN Fe /
/
o 5.1% CF IN Fe /~/~
/ /
• 0.75 °/o CFIN Fe //
/ t

20 x 2.3 % CFIN Fe #j/,y'

~" 10
o.~~ ,
t

N s

0.5
! I I I I
2 5 10 20 50
~ (10-8OHMM)
Fig. 19. Log-log plot of Rs against p for FeCr alloys, with temperature as an implicit variable (after
Carter and Pugh 1966).

high temperatures (Softer et al. 1965, Carter and Pugh 1966). However, at low
temperatures where skew scattering can be important, the behaviour can be
completely different (fig. 19) (Carter and Pugh 1966). It seems that in the F__~eCr
case, there is a strong skew scattering effect at low temperatures which has
disappeared by room temperature (but see Majumdar and Berger 1973). Dorleijn
(1976) has made an analysis in terms of skew scattering, side jump and ordinary
Hall effect in Fe alloys at helium temperatures, but the interpretation is tricky,
particularly because samples frequently show a field dependent Hall coefficient.
The extraordinary Hall coefficient has been measured as a function of tem-
perature in pure Co (Cheremushkina and Vasileva 1966).
Kondorskii (1969) suggested that the sign of the side jump effect was related to
the charge and polarization of the dominant carriers, which can be compared with
the model outlined above. No satisfactory quantitative estimates of the size of the
effect seem to have been made for ferromagnetic metals, and other basic questions
concerning this mechanism remain open.
The anisotropy of the Hall effect in single crystals is technically difficult to
study, and, as a result, the existence of an anisotropy in the extraordinary Hall
coefficient of cubic metals has been uncertain. Now evidence has been provided
for the anisotropy in Rs for Fe (Hirsch and Weissmann 1973) and for Ni (Hiraoka
1968) at room temperature. In hexagonal Co both R0 and Rs are highly anisotro-
790 I.A. C A M P B E L L A N D A. F E R T

pic (Volkenshtein et al. 1961) which means that measurements on hcp Co


polycrystals are subject to severe texture problems.

3.5. Thermoelectric power

In non-magnetic metals under elastic scattering conditions, the thermoelectric


power (TEP) coefficient depends on the differential of the resistivity at the Fermi
surface through the Mott formula:

dp
s= 3 lel p

In ferromagnets the situation is complicated by the existence of the two spin


currents at low temperatures and by magnetic scattering at higher temperatures.
The TEP curves as a function of temperatures for Fe, Co and Ni metals show
effects which are clearly due to ferromagnetic ordering (fig. 20). For Co and Ni,
the curve of S(T) shows a bulge towards negative values of S in the ferromagnetic
temperature range, and a distinct charge of slope at To. For Fe, the behaviour is
similar but complicated by a positive hump in S(T) just below room temperature.
The critical behaviour of S(T) has attracted considerable attention. In Ni, the
curve for dS/dT near Tc resembles the specific heat curve in the same way as does
dp/dT (Tang et al. 1971). Although it has been argued that the TEP anomaly
represents strictly the specific heat of the itinerant electrons (Tang et al. 1972) a
more reasonable interpretation is in terms of the critical behaviour of the elastic
scattering (Thomas et al. 1972). Combining the Mott formula and the expression

20

10

0
_10 ¸
iI
"T
x,¢

> -20

(13
-31

-41

Tc
-50
400 Tc(Ni) 8 00 1200
T (K)

Fig. 20. T h e absolute thermoelectric power of Ni, Fe, Pd and Co (Laubitz et al. 1976).
TRANSPORT PROPERTIES OF FERROMAGNETS 791

for the resistivity as a function of k near Tc leads to

S = Sp - 1 A o T ( 1 + Pn/P),
where AQ = 27r2k~/3[elEv, and Sp is the background non-magnetic TEP. Results
on GdNi2 were discussed in terms of this approach (Zoric et al. 1973).
The systematics of S(T) were studied at room temperature and above in a
number of Ni based alloys (Vedernikov and Kolmets 1961, Kolmets and Veder-
nikov 1962, K6ster and Gm6hling 1961, K6ster and R o m e r 1964). S at room
temperature becomes rapidly more positive with impurity concentration for those
alloys for which p;(O)/p~(O)~<1 ( ~ V , N i C r . . . ) while S becomes more negative
for alloys with p+(O)/pt (0) ~> 1 (fig. 21). The negative bulge in S(T) remains very
strong for a wide range of NiFe alloys measured up to Tc (Basargin and Zakharov
1974), but tends to disappear in NiV alloys (Vedemikov and Kolmets 1961).
The low temperature T E P of Ni based alloys has been analyzed using the two
current model (Farrell and Greig 1969, 1970, Cadeville and Roussel 1971). If the
intrinsic T E P coefficients for the two spin directions are S t and S+ then the
observed value of S should be S = (p; S t + Pt S;)/(p~ + p+) at low temperatures;
at high temperatures where the two currents are mixed, the impurity diffusion
thermopower becomes S = ½(St + S+). Using these two expressions, Farrell and
Greig (1969) extracted S t , S , for a number of impurities in Ni and similar
analyses have been done in Ni and Co based alloys (Cadeville et al. 1968,
Cadeville 1970, Cadeville and Roussel 1971). A detailed discussion has been given

T (K)
26 8.8%Cr
(a) 2O 40 60 80 100
24 ' i

22 11. -2
20 5.2
-4
18 .8
16 -6
14
-Q 12
::k
~'4C
8
6 -1;

4 Ni Cr
--ld

2
r i , i , i , i , i . i . i

40 80 120 160 200 2z,0 280


J (K)
Fig. 21. The absolute thermoelectric power of some nickel based alloys as a function of temperature
(after Beilin et al. 1974 and Farrell and Greig 1970). (a) NiCr; (b) Ni alloys.
792 I.A. CAMPBELL AND A. FERT

of the relationship between the electronic structure of the impurity and the T E P
coefficients (Cadeville and Roussel 1971).
Another aspect of the two current situation is the influence of magnon-electron
scattering (Korenblit and Lazarenko 1971). Scattering of a spin $ electron to a
spin I' state involves the creation of a magnon, which needs positive energy,
while spin 1' to spin + scattering is through the destruction of a magnon. The
electron-magnon scattering will then lead to a positive term in S at moderate
temperatures in alloys where the spin $ current dominates, and a negative term
in alloys where the spin 1' current dominates. The T E P due to this effect will be
superimposed on the elastic electron-impurity term except at very low tem-
peratures, and will complicate the analysis of the diffusion terms. Results on Ni
alloys have been interpreted with this mechanism (Beilin et al. 1974).
A magnon drag effect has been suggested (Bailyn 1962, Gurevich and Korenblit
1964, Blatt et al. 1967). Measurements on the T E P in a NiCu and a NiFe alloy in
applied fields appear to be consistent with this mechanism (Granneman and
Berger 1976). However, the strong positive T E P hump in pure Fe does not have
this origin (Blatt 1972).
The value of S is anisotropic with respect to the magnetization direction in a
ferromagnet. Measurements on Fe and Ni single crystals at room temperature
(Miyata and Funatogawa 1954) gave
AS100 = + 0.70 IxV/K, ASm -- - 0.13 txV/K in F e ,
and
AS~00 = +0.57 ~xV/K, ASm = +0.69 txV/K in Ni.
The Fe result was confirmed by Blatt (1972).

3.6. Nernst-Ettingshausen effect

This is the thermoelectric analogue of the Hall effect. It has been studied in the
pure ferromagnetic metals and in a number of alloys (Ivanova 1959, Kondorskii
and Vasileva 1964, Cheremushkina and Vasileva 1966, Kondorskii et al. 1972,
Vasileva and Kadyrov 1975). Like Rs, this coefficient varies strongly with tem-
perature in ferromagnets. Kondorskii (1964) proposed the phenomenological
relationship
Q = - (a + jgp)T,

and the origin of the effect was discussed in terms of the side jump mechanism by
Berger (1972) and Campbell (1979).

3. 7. Thermal conductivity

This is not a purely electron transport effect, as heat can be carried also by
phonons and even magnons, and separating out the different contributions is
difficult. Farrell and Greig (1969) in careful measurements on Ni and Ni alloys
have shown that a coherent analysis of the alloy data needs to take into account
TRANSPORT PROPERTIES OF FERROMAGNETS 793

the two current character of the conduction. They found that it was not possible
to decide for or against the presence of any electron-electron term in pure Ni at
low temperatures (White and Tainsh 1967).
At higher temperatures, Tursky and Koch (1970) have shown that it is possible
to use the spontaneous resistivity anisotropy to separate out phonon and electron
thermal conductivity.
By measurements in strong fields, Yelon and Berger (1972) identified a magnon
contribution to the low temperature thermal conductivity in N_iiFe.
The thermal conductivity of Ni shows an abrupt change of slope at Tc (Laubitz
et al. 1976). This property is very difficult to measure with high precision.

4. Dilute ferromagnetic alloys

4.1. Palladium based alloys

It has been known for some time that P dFe, PdCo, PdMn and P__ddNialloys are
"giant moment" ferromagnets at low concentrations; the transport properties of
these systems have been well studied.

4.1.1. Resistivity and isotropic m agnetoresistance


PdFe alloys are soft ferromagnets down to at least 0.15% Fe. The Fe mag-
netization at T ~ Tc saturates completely in small applied fields (Chouteau and
Tournier 1972, Howarth 1979). The magnetic disorder at relatively low tem-
peratures is in the form of magnons; for the dilute alloys (C < 2% Fe), it appears
that the magnon-electron scattering is essentially incoherent so the magnetic
resistivity is proportional to the number of magnons present, leading to a
temperature dependent resistivity proportional to T 3/2 for T ~ Tc and a charac-
teristic temperature dependent negative magnetoresistance (Long and Turner
1970, Williams and Loram 1969, Williams et al. 1971, Hamzi6 and Campbell
1978). At higher concentrations a T 2 resistivity variation replaces the T 3/2
behaviour (Skalski et al. 1970). At the Curie temperature there is a change in
slope of the p ( T ) curve but it is difficult to analyze the results in terms of critical
scattering behaviour because of smearing due to the spread of Tc values in the
samples (Kawatra et al. 1969).
PdMn alloys are "ferromagnets" below 4% Mn concentration in that they show
a high initial susceptibility below a well defined ordering temperature (Rault and
Burger 1969, Coles et al. 1975). In fact, high field magnetization measurements
(Star et al. 1975) show that the Mn magnetization only becomes truely saturated
when very strong magnetic fields are applied. The temperature dependence of the
resistivity of these alloys is qualitatively similar to that observed in PdFe, with a
change of slope in p ( T ) at Tc and a T 3/2 variation of the resistivity at low
temperatures (Williams and Loram 1969). In contrast to the PdFe alloys the
magnetoresistance remains strongly negative even when T tends to zero (Williams
et al. 1973).
794 I.A. C A M P B E L L A N D A. F E R T

PdCo alloys have very similar ordering temperatures and total magnetic
moments per atom as the PdFe alloys (Nieuwenhuys 1975), and the temperature
dependence of the resistivity is again of the same type (Williams 1970). However
the paramagnetic resistivity at T > Tc is proportional to the Co concentration
(Colp and Williams 1972) whereas in P__ddFealloys it increases as the square of the
Fe concentration (Skalski et al. 1970). The PdCo alloys below 5% Co show a
negative magnetoresistance at T ~ Tc which indicates that they are not true
ferromagnets (Hamzi6 et al. 1978a)*.
PdNi alloys are ferromagnets above a critical concentration of 2.3% Ni (Tari
and Coles 1971). Near this concentration the low temperature variation of the
resistivity of the alloys becomes particularly strong (Tari and Coles 1971). Both
the paramagnetic and ferromagnetic alloys show a large positive magnetoresis-
tance due to an increase in the local moments at the Ni sites with the applied field
(Genicon et al. 1974, Hamzi6 et al. 1978a).

4.1.2. Magnetoresistance anisotropy


PdFe, P__d_dCoand PdNi alloys all show positive anisotropies Pll > P± at moderate
magnetic impurity concentrations. At low concentrations P_ddFe samples show
vanishingly small anisotropies (Hamzi6 et al. 1978a). From this and other evidence
it has been concluded that the Co and Ni impurities carry local orbital moments.

4.1.3. Extraordinary Hall effect


Over a broad concentration range the Hall coefficient in PdFe alloys behaves
similarly to that in concentrated NiFe alloys, changing sign near 20% Fe (Matveev
et al. 1977, Dreesen and Pugh 1960). At low concentrations the Hall angle tends
to zero for PdFe and P__d_dMnbut takes on a concentration independent value for
P dNi and PdCo (Hamzi6 et al. 1978b, Abramova et al. 1974). This should be
related to the local orbital moments of Co and Ni impurities.

4.1.4. Thermoelectric power


In the concentrated ferromagnets, features clearly associated with the ferro-
magnetic ordering are visible in the temperature dependence of the TEP. For the
Pd based alloys this does not seem to be the case except perhaps when the
magnetic impurity concentration is greater than 5% (Gainon and Sierro 1970). At
1%, or lower, concentrations PdFe and PdMn show weak negative or positive
TEP below 20 K varying in a rather co'--mplex way with concentration and
temperature (Gainon and Sierro 1970, Macdonald et al. 1962, Schroeder and Uher
1978). P_dd1% Co shows a negative TEP hump at 20 K (Gainon and Sierro 1970);
this hump becomes more pronounced and goes to lower temperatures as the
concentration is decreased (Hamzi6, 1980). Below the critical concentration PdNi
alloys show a strong negative hump in the TEP around 15 K which disappears once
the concentration exceeds the critical value (Foiles 1978).
* They can he considered to be "quasiferromagnets", i.e., systems having an overall magnetic m o m e n t
but where the local m o m e n t s are each somewhat disoriented with respect to the average m o m e n t
direction.
TRANSPORT PROPERTIES OF FERROMAGNETS 795

4.2. Platinum based alloys

Again, Pt__Fe and Pt___Coare giant moment ferromagnets at concentrations of a few


percent, but at lower concentrations the behaviour is more complicated. For Pt__Fe
below about 0.8% spin glass order sets in (Ododo 1979).
In the ferromagnetic concentration range there is the usual step in p(T) at the
ordering temperature, but below 0.8% Fe this step disappears (Loram et al. 1972).
The isotropic magnetoresistance is strongly negative at concentrations less than
about 5% Fe (Hamzi6 et al. 1981).
PtCo alloys below 1% Co show resistivity variations which are complex because
of competing tendencies to Kondo condensation and to magnetic ordering (Rao et
al. 1975, Williams et al. 1975). At concentrations above about 1% Co a step can
be seen in p(T) at To. The isotropic magnetoresistance is positive at low
concentrations, becoming negative by 2% Co (Lee et al. 1978, Hamzi6 et al. 1980).
Both Pt__Fe and PtCo alloys show concentration independent resistivity aniso-
tropies and extraordinary Hall angles at low concentrations (Hamzi6 et al. 1979).
The low temperature thermoelectric power of PtCo alloys becomes strongly
negative below about 2% Co concentration (Lee et al. 1978). This TEP is sensitive
to applied magnetic fields.
PtMn alloys are spin glasses (Sarkissian and Taylor 1974), and Pt___Nialloys are
not magnetically ordered below 42% Ni.

5. Amorphous alloys

Since the early 1970s considerable effort has been devoted to the study of the
electrical and magnetic properties of amorphous alloys. The resitivity minimum
observed in many systems has been subject to much controversy.

5.1. Resistivity of amorphous alloys

The amorphous alloys have a very high resistivity (p ~ 100 Ixllcm) which changes
relatively little as a function of temperature. Figure 22 shows that, in series of NiP
alloys, the temperature coefficient changes from positive to negative as the
concentration of P increases. This behaviour is well explained in the Ziman model
of the resistivity of liquid metals (Ziman 1961) and its extension to amorphous
alloys (Nagel 1977). In the Ziman model the resistivity turns out to be propor-
tional to a(2kv) where kv is the Fermi wave vector and a(q) the atomic structure
factor. If 2kv is close to the first peak of a(q), the resistivity is high and decreases
as a function of T owing to the thermal broadening of the p~ak. In contrast, if 2kv
lies well below (or well above) the peak, the resistivity is relatively low and
increases as a function of T. In the NiP alloys (fig. 22) the additional conduction
electrons provided by the higher concentrations of P raise 2kF to the first peak of
a(q), which accounts for the experimental behaviour (Cote 1976). On the other
hand, the small resistivity upturns observed in NiP at low temperature (fig. 22)
796 I.A. C A M P B E L L A N D A. F E R T

z 1.010 ',.' ...' ' ' ' ~i ' P' ' ~ ' ' ' ' '
fl_
03 "'' ...% 26 .m
• e•l • leeal z
o
-i-- P2s " • "-.. LU
13.. • e ••e ••.e I~e e e e,, •• eo
• °% • e %°'•• ~"'4' _J
1.00I] _J
Z N 176 P24 " " <
< °° ,,~ --1.0
LL
O • • 2e I
i• -- 09
~ m 0.990 •e la_
e•e • o
" Ni P - OB
- .. • 80 20 • ~"
Q..
- 0.7
>- O.9B(]
I-- • crystaLLine
> - 0.6 >-
NiP i---
I% 80 20
- 0,5
m
¢t"
0.97C
NsP15 ".
o3

LU • 200 . . . . . . 20 ~ ~ 0.4 r¢

• ~, . . . . ~ ~ 03 >

,. 0.96C ~_ ~2o . . . . . . . . o~

az , ~, ~oo ~ i °~ 02
• " is 20 25 £t"
~ m ATOM]C PER CENT PHOSPHOROUS
0.950 " t , , ~) ' , I I ' I I J I I ' I -01
40 8 120 160 200 240 280
TEMPERATURE (K)
Fig. 22. Relative resistivities of a m o r p h o u s and crystalline Ni-P alloys. Nominal compositions are
indicated; (Pam/pcryst)293 ~ 3. T h e inset shows p and dp/dT as a function of the composition (after Cote
1976). Similar results are given by Boucher (1973) for N i - P d - P a m o r p h o u s alloys.

cannot be explained by the Ziman model. Such resistivity upturns, which


generally give rise to a resistivity minimum, have been found in many amorphous
systems. They have been found in both ferromagnetic and non-ferromagnetic
amorphous alloys and, up to now, only in alloys containing transition (or rare-
earth) metals. Their origin has been subject to much controversy.
Resistivity minima have been first found by Hasegawa and Tsui (1971a, b) in
amorphous PdSi containing Cr, Mn, Fe or Co impurities (fig. 23). The classical
features of the Kondo effect are observed: the resistivity varies logarithmically over
a large temperature range and becomes constant in the low temperature limit; at
low concentration of magnetic impurities the logarithmic term increases with the
concentration, there is a negative magnetoresistance. But, surprisingly, the resis-
tivity minimum still exists in the most concentrated alloys which are ferro-
magnetic. These results seem to indicate that weakly coupled moments subsist in
amorphous ferromagnets and can give rise to Kondo scattering. Results on many
other systems have suggested that the coexistence of ferromagnetism and Kondo
effect is quite general in amorphous alloys; thus large logarithmic upturns have
been observed (fig. 24) in ferromagnets of the series FeN•B, FeNiPB, FeN•PC,
FeNPBS (Cochrane et al. 1978, Babi6 et al. 1978, Steward and Phillips 1978),
FeNiPBA1, FeMnPBA1, CoPBA1 (Rao et al. 1979), PdCoP (Marzwell 1977); in
TRANSPORT PROPERTIES OF FERROMAGNETS 797

94.40 oO°°
o
o 94.00
oooo° °
93.60 ss
.93.20
ooO°°°ge 1Pd?9Si20
92.80 oO i34.50

o~ooO°°°° Fe 3 R:177Si20
92.00 o oo o oo 133.70
o co° ° ° ° o o

oooeo °° ooOO o ° ~18,00


o° ° ° cooo o
£0 91,20.~ ~'~ ~o °°° o m°

Cl ? ° ~ °° o°°°° Fe Pd?$Si20 117. 20


ooo ° ° 5
132,10 oOO° oOo
O_.- ~ o° 135.40
o o°

_ 116.00 o~o~ o°°°° oooo °° Fe 7 Pd73Si20


134.60
~n 134.20 o°°°
03 ~°
UJ
r,,-' g 133.'~0
133,40 o
o 133.00
o
132.60 oo
69
1322O
131.80 i , i i , , J , i ~ ,
0 40 80 120 160 200 240
TEHPERATURE (OK)

Fig. 23. Resistivity versus temperature for FexPds0-xSi20 alloys. Tc = 28 K for x = 7 (after Hasegawa
and Tsuei 1970b).

A
A a ,~AAt,~,A8 A
A
Aa 162.1
%,, A

oooooooo ~a
°o 162.0
144.3 e°o o •
A

\ A
&
t, a
1442 O % g ~' 161.9
E O

°o
o
o._, 144.1 o

%
o o
o
~ o-
1/..4.0 o
%

143.9 I I
Q01 0.1 1.0 1() 100
T(K)
Fig. 24. Resistivity versus log T plot for two ferromagnetic amorphous alloys: (A) Fe75P16B6A]3 and
(O) Fe60NilsP16B6AI3 (after Rapp et al. 1978).
798 I.A. CAMPBELLAND A. FERT

many cases the addition of small amounts of Cr strongly enhances the resistivity
upturn.
On the other hand, Cochrane et al. (1975) found that the logarithmic resistivity
upturn of several amorphous alloys was field independent, in contrast to what is
generally observed in Kondo systems. They also noticed a logarithmic upturn in
NiP alloys with high P concentration in which the Ni atoms were not supposed to
carry a magnetic moment. On the basis of these observations they ruled out the
explanation by the Kondo effect and proposed a non-magnetic mechanism. Their
model treats the electron scattering by the two level systems which are supposed
to be associated with structural instabilities in amorphous systems; a variation of
the resistivity in - l n ( T 2 + A2) is predicted, where A is a mean value of the energy
difference between the two levels. The resistivity curves of several amorphous
alloys fit rather well with such a variation law.
At the present time (1979) however the trend is in favour of an explanation of
the resistivity minima by the Kondo effect rather than by a non-magnetic
mechanism. Clear examples of logarithmic resistivity upturns in non-magnetic
systems are still lacking: alloys such as NiP or YNi c a n be suspected to contain
magnetic Ni clusters (Berrada et al. 1978). On the other hand, systematic studies
of the resistivity of FeNiPB (BaNd et al. 1978), FeNiPBAI, FeMnPBA1 (Rao et al.
1979) have shown definite correlations between the resistivity anomalies and the
magnetic properties (logarithmic term large when Tc is small, etc.); it has been also
found in several systems that the logarithmic upturn is lowered by an applied
field. Finally, M6ssbauer experiments on FeNiCrPB alloys have found very small
hyperfine fields on a significant number of Fe sites, which seems to confirm the
coexistence of ferromagnetism and Kondo effect (Chien 1979).
What we have written up to now concerned the metal-metalloid alloys which
have been the most studied amorphous alloys. Studies of metal-metal amorphous
alloys of rare-earths with transition or noble metals have been also developed
recently. Resistivity minima have been again observed in these systems but appear
to be generally due to contributions from magnetic ordering and not to Kondo
effect. In Ni3Dy (fig. 25) the resistivity increases either if a magnetic field is
applied or if the temperature is lowered below the ordering temperature To. This
suggests a positive contribution from magnetic ordering to the resistivity, in
contrast to what is observed in crystalline ferromagnetics. This has been ascribed
by Asomoza et al. (1977a, 1978) to coherent exchange scattering by the rare-earth
spins (Ni has no magnetic moment in these alloys). The model calculation predicts
a resistivity term proportional to m(2kv) where re(q) is the spin correlation
function

1
m ( q ) = NCelj( J + 1) R , ~ , exp[iq • ( R - R')IJR " JR'.

Here C1 is the concentration of magnetic ions, having local moments J and placed
at R, R ' ; the sum is over the pairs of magnetic ions.
The resistivity will depend on the magnetic order through m(2kv); for example,
TRANSPORT PROPERTIES OF FERROMAGNETS 799

I I I t 1 I 1 I I I

D y Ni 3

296 oH:O kG
~H: 8kG
H: 2OkG
o H:3OkG
295

u 294

>
I Xx "g~ •
u~
293

Q~

300~_
292
/
//
/

/ 292 I I I
291
50 100 150 200 250

t I /20I | I
40
I I
60
I I
80
I I
T ( K)

Fig. 25. T h e resistivity of a DyNi3 a m o r p h o u s alloy in several applied fields is plotted as a function of
temperature (after A s o m o z a et al. 1977).

ferromagnetic correlations will increase or decrease p according to whether the


interferences are constructive or destructive. The Ni3-RE alloys should cor-
respond to the case of ferromagnetic correlations and constructive interferences.
The A g - R E , A u - R E and A I - R E amorphous alloys also show a clear contribution
from magnetic ordering to the resistivity, but the interpretation seems to be a
little more complicated than for the Ni3-RE alloys (Asomoza et al. 1979, Fert and
Asomoza 1979). Finally, alloys of the series F e - R E and C o - R E generally show a
monotonic decrease of the resistivity from the helium range to room temperature
(Cochrane et al. 1978, Zen et al. 1979). In these alloys of high Tc the variation of
the resistivity due to magnetic ordering must be displayed over a wide tem-
perature range and is certainly difficult to separate from the normal variation due
to the phonons and the thermal variation of the structure factor. We believe that
this normal variation should be predominant, specially at low temperature.
Similarly, in the alloys such as FeNiPB discussed above, a contribution from
magnetic ordering to p(T) certainly exists but is likely covered up by other
contributions (Kondo or structural effects) at low temperature.
800 I.A. CAMPBELL AND A. FERT

5.2. Hall effect and resistivity anisotropy of amorphous alloys

The amorphous ferromagnetic alloys have a very large extraordinary Hall effect
which generally covers up the ordinary Halt effect. This is because the extra-
ordinary Hall resistivity, in contrast to the ordinary one, is an increasing function
of the scattering rate (the contributions from skew scattering and side-jump are
roughly proportional to p and p2 respectively). Thus pi~(B) is practically propor-
tional to the magnetization in many systems and, for example, is frequently used
to record hysteresis loops (McGuire et al. 1977a, b, Asomoza et al. 1977b).
The extraordinary Hall effect of ferromagnetic alloys of gold with nickel, cobalt
or iron has been studied by Bergmann and Marquardt (1979) and ascribed to skew
scattering; the change of sign of pn between Ni and Fe has been accounted for by
a model based on a virtual bound state picture of the 3d electrons. On the other
hand, the extraordinary Hall effect of FeNiPB alloys rather suggest a side-jump
mechanism (Malmhfill et al. 1978). The extraordinary Hall effect has been also
studied in amorphous alloys of transition metals with rare-earths and related to
the magnetization of the transition and rare-earth sublattices in phenomenological
models (Kobliska and Gangulee 1977, McGuire et al. 1977, Asomoza et al.
1977b).
The spontaneous resistivity anisotropy is rather large in amorphous alloys of
gold with nickel or cobalt (/911-p± -~ 1 ix12cm) and has been interpreted in a model
of virtual bound state for the 3d electrons (Bergmann and Marquardt 1979). The
resistivity anisotropy seems to be smaller in alloys of the FeNiP type (Marohnid et
al. 1977). The resistivity anisotropy has been also studied in amorphous alloys of
nickel or silver with rare-earths and turns out to be mainly due to electron
scattering by the electric quadrupole of the 4f electrons (Asomoza et al. 1979).

Reference

Abelskii, Sh. and E.A. Turov, 1960, Fiz. Met. Baber, W.G., 1937, Proc. Roy. Soc. A 158, 383.
Metalloved. 10, 801. Babic, E., Z. Marohnic and J. Ivkov, 1978,
Abromova, L.I., G.V. Fedorov and N.N. Vol- Solid State Commun. 27, 441.
kenshteyn, 1974, Fiz. Met. Metalloved. 38, Bailyn, M., 1962, Phys. Rev. 126, 2040.
90. Basargin, O.V. and A.I. Zakharov, 1974, Fiz.
Amamou, A., F. Gautier and B. Leogel, 1975, Met. Metalloved. 37, 891.
J. Phys. F 5, 1342. Beitcham, J.G., C.W. Trussel and R.V. Cole-
Anderson, J.R. and A.V. Gold, 1963, Phys. man, 1970, Phys. Rev. Lett. 25, 1970.
Rev. Lett. 10, 227. Beitel, F.P. and E.M. Pugh, 1958, Phys. Rev.
Armstrong, B.E. and R. Fletcher, 1972, Can. J. 112, 1516.
Phys. 50, 244. Beilin, V.M., T.I. Zeinalov, I.L. Rogel'berg and
Asomoza, R., A. Fert, I.A. Campbell and R. V.A. Chernenkov 1974, Fiz. Met. Metal-
Meyer, 1977a, J. Phys. F 7, L 327. loved. 38, 1315.
Asomoza, R., I.A. Campbell, H. Jouve and R. Berger, L., 1970, Phys. Rev. B 2, 4559.
Meyer, 1977b, J. Appl. Phys. 48, 3829. Berger, L., 1972, Phys. Rev. B 5, 1862.
Asomoza, R., I.A. Campbell, A. Fert, A. Berger, L., 1978, J. Appl. Phys. 49 (3), 2156.
Li6nard and J.P. Rebouillat, 1979, J. Phys. F Berger, L. and S.A. Friedberg, 1968, Phys. Rev.
9, 349. 165, 670.
TRANSPORT PROPERTIES OF FERROMAGNETS 801

Berger, L. and A.R. de Vroomen, 1965, J. Coles, B.R., H. Jamieson, R.H. Taylor and A.
Appl. Phys. 36, 2777. Tari, 1975, J. Phys. F 5, 572.
Berrada, A., N.F. Lapierre, L. Loegel, P. Colp, M.E. and G. Williams, 1972, Phys. Rev.
Panissod and C. Robert, 1978, J. Phys. F 8, B 5, 2599.
845. Connelly, D.L., J.S. Loomis and D.E. Mapo-
Bergmann, G. and P. Marquardt, 1978, Phys. ther, 1971, Phys. Rev. B 3, 924.
Rev. B 18, 326. Cote, P.J., 1976, Solid State Commun. 18, 1311.
Birss, R.R., 1964, Symmetry and magnetism Craig, P.P., W.I. Goldberg, T.A. Kitchens, and
(North-Holland, Amsterdam). J.I. Budnick, 1967, Phys. Rev. Lett. 19,
Blatt, FJ., D.J. Flood, V. Rowe, P.A. Schro- 1334.
eder and J.E. Cox, 1967, Phys. Rev. Lett. 18, Dedir, G., 1975, J. Phys. F 5, 706.
395. De Gennes, P.G. and J. Friedel, 1958, J. Phys.
Blatt, F.J., 1972, Can. J. Phys. 50, 2836. Chem. Solids, 4, 71.
Boucher, B., 1973, J. of Non-Cryst. Sol. 7, 277. Der Ruenn Su and T.M. Wu, 1975, J. Low
Bozorth, R.M., 1951, Ferromagnetism (Van Temp. Phys. 19, 481.
Nostrand, Princeton). D6ring, W., 1938, Ann. Phys. 32, 259.
Cadeville, M.C., F. Gautier, C. Robert and J. Dorleijn, J.W.F., 1976, Philips Res. Repts, 31,
Roussel, 1968, Solid State Commun. 7, 1701. 287.
Cadeville, M.C., 1970, Solid State Commun. 8, Dorleijn, J.W.F. and A.R. Miedema, 1975a, J.
847. Phys. F 5, 487.
Cadeville, M.C. and C. Lerner, 1976, Phil. Mag. Dorleijn, J.W.F. and A.R. Miedema, 1975b, J.
33, 801. Phys. F 5, 1543.
Cadeville, M.C. and B. Loegel, 1973, J~ Phys. F Dorleijn, J.W.F. and A.R. Miedema, 1976, AIP
3, L 115. Conf. Proc. 34, 50.
Cadeville, M.C. and J. Roussel, 1971, J. Phys. F Dorleijn, J.W.F. and A.R. Miedema, 1977, J.
1, 686. Phys. F 7, L 23.
Campbell, I.A., 1974, J. Phys. F 4, L 181. Dreesen, J.A. and E.M. Pugh, 1960, Phys. Rev.
Campbell, I.A., 1979, J. Magn. Magn. Mat. 12, 120, 1218.
31. Dugdale, J.S. and Z.S. Basinski, 1967, Phys.
Campbell, I.A., A. Fert and A.R. Pomeroy, Rev. 157, 552.
1967, Phil. Mag. 15, 977. Durand, J. and F. Gautier, 1970, J. Phys. Chem.
Campbell, I.A., A. Fert and O. Jaoul, 1970, J. Sol. 31, 2773.
Phys. C 3, S 95. Durand, J., 1973, Thesis (Strasbourg) unpub-
Carter, G.C. and E.M. Pugh, 1966, Phys. Rev. lished.
152, 498. Dutta Roy, S.K. and T.M. Wu, 1975, J. Low
Cheremushkina, A.V. and R.P. Vasil'eva, 1966, Temp. Phys. 19, 481.
Sov. Phys. Solid State, 8, 659. Erlich, A.C., J.A. Dreesen and E.M. Pugh,
Chien, C.L., 1979, Phys. Rev. B 19, 81. 1964, Phys. Rev. 133, A, 407.
Chouteau, G. and R. Tournier, 1972, J. de Etin Wohlman, O., G. Deutscher and R.
Phys. 32, C1-1002. Orbach, 1976, Phys. Rev. B 14, 4015.
Clark, A.L. and R.L. Powell, 1968, Phys. Rev. Farrell, T. and D. Greig, 1968, J. Phys. C, 1 sur
Lett. 21, 802. 2, 1359.
Cochrane, R.W., R. Harris, J.O. Str6m-Olsen Farrell, T. and D. Greig, 1969, J. Phys. C, 2 sur
and M.J. Zuckermann, 1975, Phys. Rev. Lett. 2, 1465.
35, 676. Farrell, T. and D. Greig, 1970, J. Phys. C, 3,
Cochrane, R.W. and J.O. Str6m-Olsen, 1977, 138.
J. Phys. F 7, 1799. Fawcett, E., 1964, Adv. Phys. 13, 139.
Cochrane, R.W., J.O. Str6m-Olsen, Gwyn Wil- Fert, A., 1969, J. Phys. C, 2, 1784.
liams, A. Lirnard and J.P. Rebouillat, 1978, Fert, A. and R. Asomoza, 1979, J. Appl. Phys.
J. Appl. Phys. 49, 1677. 50, 1886.
Coleman, R.V., 1976, AlP Conf. Proc. 29, 520. Fert, A. and I.A. Campbell, 1968, Phys. Rev.
Coleman, R.V., R.C. Morris and D.J. Sell- Lett. 21, 1190.
meyer, 1973, Phys. Rev. B 8, 317. Fert, A. and I.A. Campbell, 1971, J. de Phys.
Coles, B.R., 1958, Adv. Phys. 7, 40. (Paris) 32, Sup. no. 2-3, C1--46.
802 I.A. CAMPBELL AND A. FERT

Fert, A. and I.A. Campbell, 1976, J. Phys. F 6, Hirsch, A.A. and Y. Weissmann, 1973, Phys.
849. Lett. 44A, 230.
Fert, A. and O. Jaoul, 1972, Phys. Rev. Lett. Hodges, L., D.R. Stone and A.V. Gold, 1967,
28, 303. Phys. Rev. Lett. 19, 655.
Fert, A., I.A. Campbell and M. Ribault, 1970, Houghton, R.W. and M.P. Sarachik, 1970, Phys.
J. Appl. Phys. 41, 1428. Rev. Lett. 25, 238.
Fert, A , R. Asomoza, D. Sanchez, D. Span- Hugel, J., 1973, J. Phys. F 3, 1723.
jaard and A. Friederich, 1977, Phys. Rev. B Howarth, W., 1979, Thesis, London.
16, 5040. Huguenin, R. and D. Rivier, 1965, Helv. Phys.
Fischer, M.E. and A. Aharony, 1973, Phys. Acta, 38, 900.
Rev. Lett. 30, 559. Hurd, C.M., 1972, The Hall Effect (Plenum
Fisher, M.E. and J.S. Langer, 1968, Phys. Rev. Press, New York).
Lett. 20, 665. Hurd, C.M., 1974, Adv. Phys. 23, 315.
Foiles, C.L., 1978, J. Phys. F 8, 213. Ivanova, R.P., 1959, Fiz. Met. Metalloved. 8,
Friedel, J., 1967, Rendicanti della Scuola In- 851.
tern. di Fisica '°Enrico Fermi" XXXVII Jan, J.P., 1957, Solid State Phys. 5, 1.
Corso (Academic Press, New York). Jaoul, O., 1974, Thesis (Orsay), unpublished.
Fulkesson, W., J.P. Moore and D.L. McElroy, Jaoul, O., I.A. Campbell and A. Fert, 1977, J.
1966, J. Appl. Phys. 37, 2639. Magn. Magn. Mat. 5, 23.
Fujii, T., 1970, Nippon Kinsoku Gakkaishi Jayaraman, V. and S.K. Dutta Roy, 1975,
(Japan) 34, 456. J.P.C.S. 36, 619.
Gainon, D. and J. Sierro, 1970, Helv. Phys. Kasuya, T., 1956, Progr. Theor. Phys. 16, 58.
Acta, 43, 541. Kasuya, T., 1959, Progr. Theor. Phys. 22,
Geldart D.I.W. and T.G. Richard, 1975, Phys. 227.
Rev. B 12, 5175. Kaul, S.N., 1977, J. Phys. F 7, 2091.
Genicon, G.L., F. Lapierre and J. Soultie, 1974, Kawatra, M.P., S. Skalski, J.A. Mydosh and
Phys. Rev. B 10, 3976. J.I. Budnick, 1969, J. Appl. Phys. 41),
Goodings, D.A., 1963, Phys. Rev. 132, 542. 1202.
Grannemann, G.N. and L. Berger, 1976, Phys. Kawatra, M.P., J.I. Budnick and J.A. Mydosh,
Rev. B 13, 2072. 1970, Phys. Rev. B 2, 1587.
Greig, D. and J.P. Harrisson, 1965, Phil. Mag. Kawatra, M.P., J.A. Mydosh and J.I. Budnick,
12, 71. 1970, Phys. Rev. B 2, 665.
Gurevich, L.E. and I.Y. Korenblit, 1964, Sov. Kittel, C., 1963, Phys. Rev. Lett. 10, 339.
Phys. Solid State, 6, 1960. Kittel, C. and J.H. Van Vleck, 1960, Phys. Rev.
Guenault, A.M., 1974, Phil. Mag. 30, 641. 118, 1231.
Hamzid, A., 1980, Thesis (Orsay). Klaffky, R.W. and R.V. Coleman, 1974, Phys.
Hamzid, A. and I.A. Campbell, 1978, J. Phys. F Rev. B 10, 2915.
8, L33. Kobliska, R.J. and A. Gangulec, 1977, Amor-
HamziG A. and I.A. Campbell, J. Phys. (Paris) phous Magnetism II, eds., R.A. Levy and R.
42, L17. Hasegawa (Plenum, New York).
Hamzid, A., S. Senoussi, I.A. Campbell and A. Kolmets, N.V. and M.V. Vedernikov, 1962,
Fert, 1978a, J. Phys. F 8, 1947. Sov. Phys. Sol. St. 3, 1996.
Hamzid, A., S. Senoussi, I.A. Campbell and A. Kondorskii, E.I., 1964, Sov. Phys. JETP, 18,
Fert, 1978b, Solid State Commun. 26, 617. 351.
Hamzid, A., S. Senoussi, I.A. Campbell and A. Kondorskii, E.I., 1969, Sov. Phys. JETP, 28,
Fert, 1980, J. Magn. Magn. Mat. 15-18, 921. 291.
Hasegawa, R. and C.C. Tsuei, 1971a, Phys. Kondorskii, E.I. and R.P. Vasil'eva, 1964, Sov.
Rev. B 2, 1631. Phys. JETP, 18, 277.
Hasegawa, R. and C.C. Tsuei, 1971b, Phys. Kondorskii, E.I., A.V. Cheremushkina and N.
Rev. B 3, 214. Kurbaniyazov, 1964, Soy. Phys. Sol. St. 6,
Hayakawa, H. and J. Yamashita, 1976, Progr. 422.
Theor. Phys. 54, 952. Kondorskii, E.I., A.V. Cheremusbkina, R.P.
Herring, C., 1967, Phys. Rev. Lett. 19, 1131. Vasil'eva and Y.N. Arkipov, 1972, Fiz. Met.
Hiraoka, T., 1968, J. Sci. Hiroshima Univ. 32, Metalloved. 34, 675.
153. Kooi, C., 1954, Phys. Rev. 95, 843.
TRANSPORT PROPERTIES OF FERROMAGNETS 803

Korenblit, I.Y. and Y.P. Lazarenko, 1971, Sov. Mott, N.F., 1936b, Proc. Roy. Soc. 156, 368.
Phys. JETP, 33, 837. Mott, N.F., 1964, Adv. Phys. 13, 325.
K6ster, W. and W. Gm6hling, 1961, Zeit. Met. Nagel, S.R., 1977, Phys. Rev. B 16, 1694.
52, 713. Nieuwenhuys, G.J., 1975, Adv. Phys. 24, 515.
K6ster, W. and O. Romer, 1964, Zeit. Met. 55, Nozi6res, P. and C. Lewiner, 1973, J. de Phys.
805. 34, 901.
Kraftmakher, Y.A. and T.Y. Pinegina, 1974, Okamoto, T., H. Tange, A. Nishimura and E.
Sov. Phys. Sol. St. 16, 78. Tatsumoto, 1962, J. Phys. Soc. Japan, 17, 717.
Laubitz, M.J., T. Matsumara, 1973, Can. J. Ododo, J.C., 1979, J. Phys. F 9, 1441.
Phys. 51, 1247. Parks, R.D., 1972, AIP Conference, 5, 630.
Laubitz, M.J., T. Matsumara and P.J. Kelly, Pugh, E.M., 1955, Phys. Rev. 97, 647.
1976, Can. J. Phys. 54, 92. Ramaman, R.V. and L. Berger, 1978, Proc. Int.
Lavine, J.M., 1961, Phys. Rev. 123, 1273. Conf. Physics of Transition Metals (Toronto
Lee, C.W., C.L. Foiles, J. Bass and J.R. Cle- 1977), Institute of Physics, Conf. Ser. No. 39.
veland, 1978, J. App. Phys. 49, 217. Rao, K.V., O. Rapp, C. Johannesson, J.I.
L6onard, P., M.C. Cadeville and J. Durand, Budnick, T.J. Burch and V. Canella, 1975,
1969, J. Phys. Chem. Sol. 30, 2169. AIP Conf. Proc. 29, 346.
Loegel, B. and F. Gautier, 1971, J. Phys. Chem. Rao, K.W., H. Gudmundsson, H.U. Astr6n and
Sol. 32, 2723. H.S. Chen, 1979, J. Appl. Phys. 50 (3), 1592.
Loram, J.W., R.J. White and A.D.C. Grassie, Rapp, O., J.E. Grindberg and K.V. Rao, 1978,
1972, Phys. Rev. B 5, 3659. J. Appl. Phys. 49, 1733.
Luttinger, J.M., 1958, Phys. Rev. 112, 739. Rault, J. and J.P. Burger, 1969, C.R.A.S., 269,
Lyo, S.K. and T. Holstein, 1972, Phys. Rev. 1085.
Lett. 29, 423. Reed, W.A. and E. Fawcett, 1964a, J. Appl.
MacDonald, D.K.C., W.B. Pearson and I.M. Phys. 35, 754.
Templeton, 1962, Proc. Roy. Soc. A 266, 161. Reed, W.A. and E. Fawcett, 1964b, Phys. Rev.
McGuire, T.R. and R.I. Potter, 1975, IEET 136 A, 422.
Transactions on Magnetics, Vol. Mag. 11, Richard, T.G. and G.J.W. Geldart, 1973, Phys.
1018. Rev. Lett. 30, 290.
McGuire, T.R., R.J. Gambino and R.C. Taylor, Ross, R.N., D.C. Price and Gwyn Williams,
1977a, J. Appl. Phys. 48, 2965. 1978, J. Phys. F, 8, 2367.
McGuire, T.R., R.J. Gambino and R.C. Taylor, Ross, R.N., D.C. Price and Gwyn Williams,
1977b, I.E.E.E. Transactions on Magnetism 1979, J. Mag. Mag. Mat. 10, 59.
M A G 13, 1977. Ruvalds, J. and L.M. Falicov, 1968, Phys. Rev.
Majumdar, A.K. and L. Berger, 1973, Phys. 172, 508.
Rev. B 7, 4203. Sakissian, B.V.B. and R.H. Taylor, 1974, J.
Malmh~ill, R., G. B~ickstr6m, K. Rao, S. Phys. F 4, L 243.
Bhagat, M. Meichle and M.B. Salamen, 1978, Sandford, E.R., A.C. Erlich and E.M. Pugh,
J. Appl. Phys. 49, 1727. 1961, Phys. Rev. 123, 1947.
Mannari, J., 1959, Prog. Theor. Phys. 22, 335. Schroeder, P.A. and C. Uher, 1978, Phys. Rev.
Marohnic, Z., E. Babic and D. Pavuna, 1977, B 18, 3884.
Phys. Lett. 63A, 348. Schwerer, F.C. 1969, J. Appl. Phys. 40, 2705.
Marzwell, N.I., 1977, J. Mag. Mag. Mat. 5, 67. Schwerer, F.C., 1974, Phys. Rev. B 9, 958.
Matveev, V.A., G.V. Fedorov and N.N. Vol- Schwerer, F.C. and L.J. Cuddy, 1970, Phys.
tenshteyn, 1977, Fiz. Met. Metalloved. 43, Rev. B 2, 1575.
1192. Schwerer, F.C. and J. Silcox, 1968, Phys. Rev.
Matsumoto, H., H. Saito, M. Kikuchi, 1966, J.J. Lett. 20, 101.
Inst. Meta. 30, 885. Schwerer, F.C. and J. Silcox, 1970, Phys. Rev.
Miedema, A.R. and J.W.F. Dorleijn, 1977, J. B 1, 2391.
Phys. F 7, L 27. Skalski, S., M.P. Kawatra, J.A. Mydosh and J.I.
Mills, D.L., A. Fert and I.A. Campbell, 1971, Budnick, 1970, Phys. Rev. B 2, 3613.
Phys. Rev. B 4, 196. Shacklette, L.W., 1974, Phys. Rev. B 9,
Miyata, N. and Z. Funatogawa, 1954, J. Phys. 3789.
Soc. Japan, 9, 967. Shumate, P.W., R.V. Coleman and R.C.
Mott, N.F., 1936a, Proc. Roy. Soc. 153, 699. Eiwaz, 1970, Phys. Rev. B 1, 394.
804 I.A. CAMPBELL AND A. FERT

Simons, D.S. and M.B. Salomon, 1974, Phys. Shirakovskii, 1961, Fiz. Met. Metalloved. 11,
Rev. B 10, 4680. 152.
Smit, J., 1951, Physics, 17, 612. Volkenshtein, N.V. and V.P. Dyakina, 1971,
Smit, J., 1955, Physica, 21, 877. Fiz. Met. Metalloved. 31, 773; The Phys. of
Smith, T.R., R.J. Jainsh, R.N. Shelton and Met. and Metallog. 31, no. 4, 101.
W.E. Gardner, 1975, J. Phys. F 5, L 96. Volkenshtein, N.V., V.P. Dyakina and V.C.
Softer, S., J.A. Dreesen and E.M. Pugh, 1965, Startsev, 1973, Phys. St. Sol. (b) 57, 9.
Phys. Rev. 140, A 668. Vonsovskii, S.V., 1948, Zh. Eksper. Teor. Fiz.
Sousa, J.B., M.R. Chaves, M.F. Pinheiro and 18, 219.
R.S. Pinto, 1975, J. Low Temp. Phys. 18, 125. Vonsovskii, S.V., 1955, Izv. Akad. Nauk 555 B,
Souza, J.B., M.M. Amado, R.P. Pinto, J.M. Ser. fiz. 19, 447. Bull. Acad. Sc. USSR, 19,
Moreira, M.E. Brago, M. Ausloos, J.P., 399.
Leburton, P. Clippe, J.C. van Hay and P. Weiss, R.J., A.S. Marotta, 1959, J. Phys. Chem.
Morin, 1979, J. de Phys. (Paris) 40, sup. no. 5, Sol. 9, 3202.
*C5--42. Weiser, O. and K.M. Koch, 1970, Zeit. Nat.
Star, W.M., S. Foner and E.J. McNift, 1975, 25A, 1993.
Phys. Rev. B 12, 2690. White, G.K. and S.B. Woods, 1959, Phil. Trans.
Steward, A.M. and W.A. Phillips, 1978, Phil. Roy. Soc. (London) A 251, 273.
Mag. B 37, 561. White, G.K. and R.J. Tainsh, 1967, Phys. Rev.
Su, D.R., 1976, J. Low Temp. Phys. 24, 701. Lett. 19, 105.
Swartz, J.C., 1971, J. Appl. Phys. 42, 1334. Williams, G., 1970, J. Phys. Chem. Solids, 31,
Tang, S.H., F.J. Cadieu, T.A. Kitchens and P.P. 529.
Craig, 1972, AIP Conf. Proc. 5, 1265. Williams, G. and J.W. Loram, 1969a, J. Phys.
Tang, S.H., T.A. Kitchens, F.J. Cadieu, P.P. Chem. Solids, 30, 1827.
Craig, 1974, Proceedings LT 13 (Plenum Williams, G. and J.W. Loram, 1969b, Solid
Press, New York) 385. State Commun. 7, 1261.
Tari, A. and B.R. Coles, 1971, J. Phys. F 1, L 69. Williams, G., G.A. Swallow and J.W. Loram,
Taylor, G.R., Acar Isin and R.W. Coleman, 1971, Phys. Rev. B 3, 3863.
1968, Phys. Rev. 165, 621. Williams, G., G.A. Swallow and J.W. Loram,
Thomas, G.A., K. Levin and R.D. Parks, 1972, 1973, Phys. Rev. B 7, 257.
Phys. Rev. Lett. 29, 1321. Williams, G., G.A. Swallow and J.W. Loram,
Tsui, D.C., 1967, Phys. Rev. 164, 669. 1975, Phys. Rev. B 11, 344.
Turner, R.E. and P.D. Lond, 1970, J. Phys. C 3, Yamashita, J. and H. Hayakawa, 1976, Progr.
S 127. Theor. Phys. 56, 361.
Turov, E.A., 1955, Isv. Akad. Nauk SSSR, Ser. Yamashita, J., S. Wakoh and S. Asano, 1975, J.
fiz. 19, 474. Phys. Soc. Jap. 39, 344.
Turov, E.A. and A.N. Volshinskii, 1967, Proc. Yao, Y.D., S. Arajs and E.E. Anderson, 1975,
10th Intern. Conf. Low Temperature Phys., J. Low Temp. Phys. 21, 369.
Izd. Viniti, Moscow. Yelon, W.B. and L. Berger, 1970, Phys. Rev.
Tursky, W. and K.M. Koch, 1970, Zeit. Nat. Lett. 25, 1207.
25A, 1991. Yelon, W.B. and L. Berger, 1972, Phys. Rev. B
Tyler, E.H., J.R. Clinton, H.L. Luo, 1973, Phys. 6, 1974.
Lett. 45A, 10. Yoshida, K., 1957, Phys. Rev. 107, 396.
Ueda, K. and T. Moriya, 1975, J. Phys. Soc. Zen, D.Z., T.F. Wang, L.F. Liu, J.W. Zai, K.T.
Japan, 39, 605. Sha, 1979, J. de Phys. 40, C5-243.
Van Elst, H.C., 1959, Physics, 25, 708. Ziman, J.M., 1960, ELectrons and Phonons
Van Peski Tinbergen, T. and A.J. Dekker, (Clarendon Press, Oxford) p. 275.
1963, Physica, 29, 917. Ziman, J.M., 1961, Phil. Mag. 6, 1013.
Vassilyev, Y.V., 1970, Phys. St. Sol. 38, 479. Zoric, I., G.A. Thomas and R.D. Parks, 1973,
Valil'eva, R.P. and Y. Kadyrov, 1975, Fiz. Met. Phys. Rev. Lett. 30, 22.
Metalloved. 39, 66. Zumsteg, F.C. and R.D. Parks, 1970, Phys. Rev.
Vedernikov, M.V. and N.V. Kolmets, 1961, Lett. 24, 520.
Sov. Phys. Sol. St. 2, 2420. Zumsteg, F.C. and R.D. Parks, 1971, J. de Phys.
Visscher, P.B. and L.M. Falicov, 1972, Phys. St. 32, C1-534.
Sol. B 54, 9. Zumsteg, F.C., F.J. Cadieu, S. Marcelja and
Volkenshtein, N.V., G.V. Fedorov and V.P. R.D. Parks, 1970, Phys. Rev. Lett. 15, 1204.
SUBJECT INDEX

abbreviation, see phases temperature dependence of


absorption coefficient, a, of BaO-6Fe203 as a BaScxFe12-xO19 375
function of wavelength 359 magnetocrystalline, formulae 233
AC magnetization of the alnicos 18l shape 50, 52ff
activation energy 101, 103 alnicos 1-4, 113-120
of pinned wall 103 alnico 5 and alnico 5 D G (alnico 5-7)
of thin wall 104 121-133, 149
additives 462-464, 468, 476, 499, 500, 514, 522 alnicos 8 and 9 145, 149
after-effects Fe2NiA1 120, 133
dielectric relaxation 276 single ion, spinels 235ff
elastic relaxation 286 surface 52
magnetic, spinels 249ff uniaxial 50
aging, see stability anisotropy constant 67
air gap 44, 46 anisotropy energy 49, 62, 85, 96
amorphous anisotropy field 49ff, 50, 446, 493, 499, 537
alloys 795ff BaAlxFe12-xO19 373
ribbons 524 BaCox/2Tix/2Fe12-x019 378
Anderson localization 268 BaCrxFe12-xO19 373
anhysteretic magnetization of the alnicos 179 BaGaxFe12-xO19 373
anilin process 465 Ba(TiCo)xFe12_xO19 379
anisotropy BaO.x(TiCoO3)-(6- x)Fe203 347
C O 2+ contribution 236, 240, 247, 251 BaO.6Fe~O3 332-335, 349, 381
exchange 52 BaZnx/2Ge~/zFe12-xO19 380
induced 246ff B aZn:,/2Irx/2Fe12-~O19 381
experimental data for spinels 251ff BaZn2x/3Nbx/3Fe12_xO19 380
magnetic 49ff, 55 BaZn2x/3Tax/3Fe12-x019 380
magnetocrystalline 50, 53ff, 56, 57, 102, 120, BaZnx/2Tix/2Fe12-x019 347, 381
133 BaZr~Tir Mnz Fea2-x-r_z O19 380
magnetocrystalline constant BaZn2x/3Vx/3Fe12-x019 380
BaCox/2Tix/2Fe12-x019 379 CaO--A1203-Fe203 375
BaCu~Fe12-~O19-xFx 386 PbO.6Fe203 332-334
BaInxFel2-xO~9 376 SrAlxFe12-xOt9 373
BaNixFe12-xOa9-xFx 386 SrO-xA1203.(6- x)Fe203 347
BaO.6Fe203 329-332, 376, 386 SrO-6Fe203 332-334, 374
LaFe12Oi9 367 temperature dependence of
LaFeZ+Fe~-O19 367 BaO.6Fe203 333
Na0.sLa0.sFe12019 367 PbO-6Fe203 333
PbO.6Fe203 329-332 SrO.6Fe203 333
SrO-6Fe203 329-332 W-type compounds 433

833
834 SUBJECT INDEX

Y-type compounds 432 magnetocrystalline anisotropy 66ff


Z-type compounds 433 of antiphase boundary 92-94
antiferromagnetic 52, 89 of buckling mode 63
antiferromagnetism of curling mode 61, 63
oxide spinels 224 of discrete sites 100
antiparallel coupling 89 of fanning mode 65
application of of planar defect 87 , 88
hard ferrites 535, 581, 582 of powder 486, 553, 566
plastoferrites 585, 586, 592 of thin wall 94ff
assemblage 463, 484 of twisting mode 63
atomizer 469, 483 of wall pinned by cavities 76, 77, 79
attritor 469, 481, 483, 485 of whiskers 64
particle packing density 62
(BH)m,x value shape anisotropy 60ff
angular dependence 561, 564 temperature dependence 447, 551-554, 579,
(BpH)max value 579, 580 580
binder 463, 471, 483, 484, 527 uniform rotation 61, 63
blending, see raw materials, mixing columnar crystallization of alnico 9, 142
Bloch wall thickness of PbO.6Fe203 358 compass 4
Boltzmann constant 101 compaction anisotropy 492, 540, 561
Brown's paradox 60 compensation material 560
bubble memory 23 complex permeability, temperature dependence
bubbles 21 of BaO.6FeaO3 344
bulk modulus K, variations with temperature compressive strength
BaO.6Fe203 360, 361 BaO-6Fe203 361
core loss 12
calcination, see reaction sintering creep, by compressive deformation
calendering, see rolling BaO.6FezO3 361
calenders 585 critical
calibration 510, 517 behaviour 773
charge-density waves (in CuV2S4) 624 stress intensity factor 573
chemical analysis volume for thermal stability 102
BazZn2Fe12022 (Zn2Y) 436 critical radius 77
chemical vapour deposition 534 cylinder 59, 63
cobalt, effects of fine particles 55ff
in alnicos 154 resistivity phenomena 773ff
coercive force 334-342 sphere 62
angular dependence of BaO.6Fe203 361 upper and lower bounds 55, 59
effect of milling time on BaO.6Fe603 337 crystal field splitting
effect of packing factor on BaO.6FeeO3 336 of 3d n ions in oxide spinels 202ff
effect of particle size on BaO-6Fe203 338 crystal field transitions
temperature dependence of of Cr 3+ (in CdCr2S4) 659ff
(Ba or Sr)(Cu+ Ge)xFe~2-xOx9 383 of Co 2+ (in CoCr2S4) 713, 718
(Ba or Sr)(Cu + Si)xFe12-xO19 383 crystal growth, see grain growth
(Ba or Sr)(Cu + Ta)xFe12-xO~9 383 crystal structure
(Ba or Sr)(Cu+ V)xFe12-xO19 383 U-type and other compounds 402
coercivities of W-type compounds 396
alnicos 1-4 112-114 X-type compounds 401
alnicos 5-9 127, 131, 133 Y-type compounds 397
alnico 8, extra high 141 Z-type compounds 400
coercivity 49, 50, 59, 60 crystallographic texture of alnicos and effects
angular dependence 561, 564-566 on magnetic properties 161
flux 42 Curie constant 343
magnetization 41 Curie temperature 74, 113, 126, 155, 156, 550
SUBJECT INDEX 835

Ag0.sLa0.sFe12019 367 die 489


BaAlxFe12-xO19 371, 372 pressing 489, 513, 584, 585
BaCoxFea2-xO19-xFx 386 dielectric behaviour 569
BaCrxFelz-xO19 372 dielectric constant
BaCuxFe12-xO19-xFx 386 BaNizAlxFe16-xO27 (NiA1)W 435
BaGaxFe12-~O19 371, 372 oxide spinels 275ff
BalnxFe12-xO19 376 dielectric constant, real part
Bal-~(K or Bi)x(Cu, Ni or Mn)xFe12-xO~9 temperature dependence of BaO-6Fe203
383 365
BaNixFe12-x O19Fx 386 diffusion couples 472, 534
BaO.6Fe203 326, 327, 386 dilute ferromagnets 793
Ba(TiCo)xFe12 xO19 379 dipole 55
BaZnx/2Irx/2Fe12 xO19 381 directional ordering 54
BaZnx/2Tix/2Fe12-xOa9 381 dispersants 483
Ca0.88La0.14Fe12019 367 domain observation
LaFe2+Fe3~ O19 367 BaO.6Fe403 354-358
LaFe12019 367 domain wall
Na0.sLa0.sFe12019 367 detachment 92ff
Nrel theory 222 nucleation 59, 66, 67, 81
oxide spinels 225, 296 180 degrees 67ff, 84
PbO-6Fe203 326, 327 pinning 59, 67, 81
SrAI4.sFeT.2019 374 thickness 62, 67ff
SrAlxFe12-xO19 371,372 domain wall energy 59, 67ff, 68, 334, 335
SrO.6Fe203 326, 327 BaO.6Fe203 335, 355
cutting 463, 584 PbO-6FezO3 335
cyclic heat treatment of alnico 5 129 SrO.6Fe203 335
domain wall mobility
BaFeH.3AI0.7019 358
DC electrical conductivity BaO-6Fe203 358
oxide spinels 260ff domain width
deagglomeration 463 BaO.6Fe203 355, 358
Debye temperature PbO-6Fe203 358
Ba-6FezO3 361 SrO.4.2Fe203-1.8A1203 355
spinel ferrites 289 SrO.4.5Fe203.1.5A1203 355
demagnetization 510-512 domain width effect of crystal thickness on
curves 535, 539, 588, 589 BaO.6Fe203 357-358
curves, comparison of various materials PbO-6Fe203 357-358
578, 579 SrO'6Fe303 357-358
curves, temperature dependence 555-557 SrO.(6- x)Fe203-xA1203 357
factor 49, 53 domain width effect of magnetic field on
influence of method 512 BaO-6Fe203 355-356
mode 56 SrO-4.2Fe203-1.8A1203 355
demagnetizing field 42, 44, 53, 60 SrO.4.5Fe203-1.5A1203 355
at conical pit 80 DS processes 519, 525
at surface defects 59 dynamic excitation of alnicos 180
density 322, 326, 349, 384
apparent powder 464, 466, 471 easy axis 49
green 490, 514 economic aspects of magnetism 6, 7
sintered 498, 502 effects of y-phase in alnicos 138-140
tap 465, 471 elastic constants of oxide spinels 285ff
uniformity 491 elastic moduli 572
X-ray 575 BaFe12019 361
designation, see phases electric conductivity
dewatering 482, 494 ferroxplana-type compounds 434
836 SUBJECT INDEX

electric conductivity, temperature dependence feed materials, s e e raw materials


of ferrimagnetism 216ff
BaO.6Fe203 364, 365 collinear, s e e Nrel configuration
PbO-6Fe203 364 effect of diamagnetic substitution, spinels
electric steels 9-11 227ff
electrical properties effect of magnetic field 224, 229
of sulpho- and selenospinels 607ff spiral 223ff
ellipsoid 53 theory, spinels 221ff
energy triangular arrangement, s e e Yafet-Kittel
anisotropy 49, 57, 62, 85, 96 configuration
conversion capability 579 ferrite(s) 12
coupling 90 components 19
domain wall 59 electrical properties 262ff
exchange 57, 62, 68, 78, 85, 90, 96 ferromagnetic resonance data 258ff
interaction between magnetized bodies 47 magnetization data 294ff
magnetocrystalline anisotropy 56, 78 magnetocrystalline anisotropy data 292ff
magnetostatic 49, 55, 57, 59, 62, 96 magnetostriction data 294ff
energy product 27, 42 survey of intrinsic magnetic properties 296
maximum 44 ferromagnetic resonance, s e e FMR 101
of alnicos 112-114 ferromagnetism
ESR oxide spinels 224
Camll2 xFexO19 370 ferroxdure 28
Euler's equation 68, 84 fibre texture 473, 526, 562
eutectic field
composition 455, 458, 524 anisotropy 50, 53, 60, 64
temperature 455, 458 demagnetizing 42, 44, 53, 60
exchange nucleation 66, 81
biquadratic 218, 230 Weiss 60
constant of PbO.6Fe203 335 filter press 469
coupling 56 fine particles 55ff
energy 57, 62, 68, 78, 85, 90, 96 Fisher sub-sieve-sizer 485
coefficient 68, 69ff fluidized bed 466, 513, 525
of body-centred cubic lattice 70, 73 flux density, magnetic 445
of face-centred cubic lattice 71, 73 fluxes 522
of hexagonal close-packed lattice 71, 73 FMR
of simple cubic lattice 71, 73 BaO-6Fe203 345-347
integrals, values in oxide spinels 218ff BaO-x(TiCoO3).(6- x)Fe203 347
resonance frequency 256 BaZnx/2Irx/2Fe12-xO19 380-381
striction spinels 244 B aZnx/2Tix/2Fe12-xO19 380-381
exchange interaction 55, 60 SrO.xAI203(6- x)Fe203 347
between Cr 3+ ions in semiconducting sulpho- FMR effect of DC field on the frequency of
and selenospinels 647ff, 701if, 728ff BaO.6FezO3 345, 346, 348
in metallic sulpho- and selenospinels 608, SrO-6Fe203 348
632ff, 644ff formation kinetics
existence range 449 ferroxplana-type compounds 402
extrusion 513, 515, 526, 527, 532, 584 formation process
BaO.6FeaO3 315-317
fanning PbO-6Fe203 315-317
asymmetric 65 SrO-6Fe203 315-317
symmetric 64 fracture surface energy 574
far infrared absorption data free energy
oxide spinels 284 magnetic 46ff
Faraday rotation reversible change 47
BaO-6Fe203 358, 359 free drying 518
SUBJECT INDEX 837

gangue 464 high field behaviour 765


garnets 20 history of
geometric defects 509 ferrites 12
glass melt process 523 garnets 20
glassy phase 457, 499 iron silicon alloys 10
grain growth 462, 480, 497, 498, 505-507 magnetism 3
anisotropic 475 permanent magnets 24
discontinuous 472, 498, 501, 505, 506, 517 homogeneity range 451, 454, 457, 459
inhibition 457, 459, 499 homogeneous nucleation 81
granulate 462, 469-471, 478, 484, 492 honeycomb domain, stability against tern-
granulation 463, 468, 471, 472 perature
grinding 463, 508, 513, 573, 579, 580, 584 BaO-6Fe203 358
mechanism 484, 509 PbO'6Fe203 358
gyromagnetic factor g hot
BaO-6Fe203 348 deformation techniques 525, 526, 530
SrO-6Fe203 348 pressing 489, 513, 518, 525, 528
gyrator 15 hydrothermal method 518
gyromagnetic ratio 102 hyperfine fields
effective 256 Fe 57 in oxide spinels 298
hyperfine magnetic field
BaMg2Fe16027 423
Hall coefficient hysteresis loop(s) 41, 44, 48, 50, 51, 67, 541-
extraordinary 752, 754, 783ff, 793 543
ordinary 752, 778 after annealing 81, 82
Hall effect 751ff after polishing 81, 82
extraordinary
definition 754
ilmenite 467
of dilute ferromagnets 794
impurities 464, 499
theory and experimental data 793
indirect shaping 463, 510
ordinary 778
induction (magnetic) 445
planar 754
injection moulding 484, 585
hard ferrite particles
interaction
aligning 462, 463, 479, 482, 492-494, 497,
domain wall with cavities 74ff
583
domain wall with crystal lattice 81ff
grain size determination 485
exchange 55, 60
shape anisotropy 475, 485, 514, 518, 583
magnetostatic 54
hard ferrites 443
interracial energy of al and a2 phases 117
annual mass production 577, 592
intermediate product 457, 472, 473, 476, 477,
bonded 582
522
hardness 502, 574, 575
internal field, temperature dependence of
H - C process 519
LaO.6F~O3 354, 368
heat capacity
SrO.6Fe203 352
BaO.6Fe203 361, 362
ionic radii, list of several ions 317
heat treatment of alnicos
iron oxide
annealing 112, 118, 119, 121-123, 138, 141,
natural 465-467
142
synthetic 465, 468, 469
homogenization 111, 112, 121, 123, 125,
isomer shift, temperature dependence of
138, 141, 142
SrO.6Fe203 352
thermomagnetic 121, 125, 130, 137, 138,
isotropic pressing 489
141, 142
Heisenberg model 73
hematite, s e e iron oxide, natural Jahn-Teller effect
hexaferrites, s e e hard ferrites cooperative in oxide spinels 213ff
hexagonal ferrites 443, 454, 473 elastic constants 285
838 SUBJECT INDEX

far infrared spectra 284, 286 lattice defects 59, 60


in Ba2Cu2Fex2022 (Cu2Y) 435 leakage factor 46
in FeCr2S4 701, 707 light switching matrix 34
magnetic anneal 249, 255 linear thermal expansion 569-571,580
liquid phase 476, 499
Kerr rotation, as a function of wavelength epitaxy 534
BaO.6Fe203 358, 359 lodestone 4
(K1)m/M~, temperature dependence of Lorentz microscopy 94
BaO.6Fe203 339 lubricants for
kilns 470, 478-480, 507, 508 milling 483
rotary 478, 479 pressing 492
kneaders 585 Lurgi process 466
Kondo effect 796
Kopp's rule 288 machining, see grinding
Madelung constant
oxide spinels 207
lattice constant(s) 575 magnetic
Ag0.sLa0.sFe12019 366 after-effect 569
BaA112019 370, 385 of BaO-6Fe203 344
BaAlxFe12-xO19 369 anneal 245
BaCozA112 xO19-xFx 385 constant ~0 445
BaCoxFelz-xO19-xFx 385 domain of ferroxplana-type compounds 436
B aCo0.sZnzA19.5O16.5F2.5 385 hardness 443
BaCoZnAI10OlyF2 385 pressure 98
BaCrsFe4019 370 viscosity 101, 102
BaCr~Felz-xO19 369 of the alnicos 173
BaCuxFelz_xO19 xFx 385 magnetic properties
BaFe22+Ni0.58Fels.~2018.42F0.58 385 dependence on sintering temperature 503,
BaGa12019 370 504
BaGaxFelz_~O19 369 of isotropic magnets 457, 459
BalnxFea2-xO19 376 of powder 479
BaNixAl~2-xO19 xFx 385 of pressed parts 491
BaNixFex2-xO19-xFx 385 of sulpho-, seleno- and tellurospinels 607ff
BaNi0.sZn2A19.5016.sF2.5 385 primary 459, 462
BaO.6Fe203 322, 376 magnetic structure
BaScxFe~2-xO19 376 BaCoxFelz-x O19 384
Bal-xSr~Fe12019 368 BaO'6Fe203 323-325
BaZn~/zlrx/2Fe12 ~O19 380 BaScxFe12-xO19 377
Ca(AIFe)12019 369-370 magnetization (quantity) 445
CaA112019 369-370 magnetization (remagnetization) 463, 510,
Ca0.ssLa0.14Fe12019 366 512, 541, 584
LaMgGa11019 369-370 magnetization
Pbl-~Ba~Fex2019 368 buckling 62, 65
PblnLgFe10.1019 376 changes
PbO-6Fe203 322 coherent reversal 518
SrAl12Ot9 370 irreversible 550, 557-560
SrAl~Fe12-xO19 369 reversible 550, 554, 557-559, 579
SrCr6Fe6019 370 curling 62, 63
SrCr~Fe12-xO19 369 in the alnicos 169
SrGa12Ox9 369-370 curve, initial 41
SrO.6Fe203 322 fanning . 64
Sra-xPbxFe~2019 368 field dependence of
temperature dependence of BaO-6FeeO3 330
BaO-6Fe203 363 PbO.6Fe203 329, 330
SUBJECT INDEX 839

remanent 41, 42, 44 alnico 9 146


saturation 41 FezNiAI 113, 118, 120
twisting 62 microwave linewidth
uniform 57 W-type compounds 432
magnetizing process 341 Y-type compounds 431
magnetocrystalline anisotropy 120, 133 Z-type compounds 431
ferroxplana-type compounds 412 milling
magnetomotoric force 46 particle size after 484
magnetooptical effects of oxide spinels 282ff reaction with water 488
magnetooptical properties, of sulpho- and wet 463, 472, 481,485
selenospinels 663, 685, 687, 711, 713, 718 mills 469, 481-483
magnetoresistance 776ff, 793 miscibility 461
in sulpho- and selenospinels 607, 651ff, mixers 585
657ff, 679ff, 691, 692, 708ff, 712, 715, 718, mixing, s e e raw materials
727 Mott's formula for variable range hopping 264
ordinary 776 Mrssbauer effect
magnetoresistive element 32 ferroxplana-type compounds 421
magnetostatic AI, Cr, ZnTi, ZnGe, ZnSn, ZnZr, CuTi,
energy 49, 55, 57, 59, 62, 96 CoTi, CoCr and NiTi substituted M-type
of cavity 76 compounds 353354
interaction domains in the alnicos 170 BaAIxFe12 xO19 370
magnetostriction 54, 567 BaO.6Fe203 351-354
alnico 5 133, 134 BaZnxTir Mnz Fe12_x_y-zO19 380
alnico 5DG (alnico 5-7) 133, 134 CaAllz-xFexO19 370
dipole-dipole 242 LaO-6Fe203 354
linear, basic relations 234 PbO-6Fe203 351-354
single ion 238ff SrAlxFea2-xOl9 370
BaFe2FelrOz7 (Fe2W) 425 SrO-6Fe203 351-354
BaO-6Fe203 360 Tl05La0.sFelzO19 368
magnons, dispersion relation for spinels 231ff M6ssbauer spectra of
main components 464 n9I (in CuCr2Te3I) 645
manufacturing process 57Fe in sulphospinels 707, 715, 716, 720,
hard ferrite, usual 462, 463 724-726, 729, 730, 733
hard ferrite, special 513 119Sn in sulphospinels 634, 641, 711, 7'14
plastoferrite 583, 584 M6ssbauer spectroscopy of the alnicos 148
mass efficiency 579, 580 moulded alnico 149
material length l multipole 55
BaFe12019 358
PbFe12019 358
Maxwell's equations 43 Nrel configuration 222
mechanical stability conditions, spinels 222, 224
properties of BaO-6Fe203 361 temperature, oxide spinels 225
work 47 neighbouring phases 454, 455, 458, 461
melting Nernst-Ettinghausen effect 792
congruent 453 neutral zone 491
incongruent 457 neutralization 511
point 455 notched-bar impact test, s e e strength values,
techniques 519 impact
memory cores 17 NMR (nuclear magnetic resonance)
micromagnetic 55, 57, 60, 81, 84 Y-type compounds 427
microstructure frequency temperature dependence of
alnicos 1--4 113 BaO.6Fe203 350
alnico 5 134, 136 spin-echo amplitude versus resonance
alnico 8 146 frequency, BaO.6Fe203 350
840 SUBJECT INDEX

of BaO.6Fe203 347-351 permeability 565


of nl'n3Cd in semiconducting sulpho- and permanent magnet characteristics 536, 540
selenospinels 666, 691 influence of mechanical stress 565, 567
of 59Co in sulphospinels 625, 713, 736 influence of neutron irradiation 567
of 53Cr in metallic sulpho- seleno- and tel- optimum values 538, 585, 587, 588
lurospinels 632, 636, 641, 642 temperature dependence 550
of 53Cr in semiconducting sulpho- and perovskite 477
selenospinels 666, 669, 690, 694, 710, 713 phase diagram(s) 449
of 63'65Cu in metallic sulpho-, seleno- and BaO-Fe203 310-313
tellurospinels 622, 623, 625, 626, 630, 632, BaO-MeO-Fe203 308
636, 641, 642 PbO-Fe203 313-315
of 63'65Cu in semiconducting sulpho- and SrO-Fe203 312-314
selenospinels 700, 721, 733 phases
of 119'2°1Hgin sulpho- and selenospinels 669, abbreviations 450
694 designations 451
of rain in sulpho- and selenospinels 700 phenacite 198
of 77Se in selenospinels 641, 691, 694 photoluminescence
of ~2STein tellurospinels 642 BaGaa2019 375
of 51V in sulphospinels 624 LaMgGa11019 375
nucleation MgGa204 375
field 66, 81 SrGa12019 375
reverse domains 66, 67, 89 photon structure of sulpho- and selenospinels
wall 80, 93 664, 668, 675, 689, 692, 696, 705, 711, 713
photomagnetic effect
olivine 198 oxide spinels 250
operating point 43 selenospinels 678ff, 696
optical properties pinning force 75, 79
BaO-6Fe203 358-360 plasticising agents 515, 527
PbO.6Fe203 358-360 plastoferrites 443, 479, 486, 582
sulpho- and selenospinels 607, 645, 653, Poisson's
659ff, 668, 673, 682ff, 692, 696, 713, 718 number 572
optical transition in oxide spinels 277ff ratio of BaFe12019 361
polarization (magnetic) 445, 492
porosity 502
paramagnetic properties 342, 343 pot cores 14
paramagnetic susceptibility, temperature potential energy of magnetic field 44
dependence of precursor phases, s e e intermediate product
BaM 342 preferred direction 462, 473, 510
PbM 342, 343 press forging, s e e hot deformation techniques
SrM 342 pressed alnicos 149
particle(s) presses 494, 496, 497
accelerator 17 pressing 489
fine 55ff dry 463, 495, 517
interactions 164--166 orienting field 489
misalignment, effects on wet 517
coercivity 163 pressure filtration, s e e compression moulding,
remanence 161 wet
orientation determination 563, 564 pressure sintering, s e e hot pressing
Peierls force 98 P - T diagram
pellet density 471 BaO.6Fe203 311
peritectics 461 Fe304 311
peritectoid reaction temperature 455 2FeO.BaO.8Fe203 311
permanent 2FeO.SrO-8Fe203 311
magnet materials 443 pyrite 465, 467
SUBJECT INDEX 841

quadrupole splitting, temperature dependence nucleation 66, 67, 89


of SrO.6Fe203 352 of cavity 78
Righi-Leduc effect 756
rigidity modulus, temperature dependence of
raw materials 462-465, 499 BaO.6Fe203 360, 361
coprecipitation 473, 477, 518 rolling techniques 513, 515, 527, 532, 584
mixing 468, 469, 471, 472 rotational hysteresis in alnicos 177
precipitation 513, 517 rubber 584, 585
reaction Ruthner process 465, 466
kinetics 462, 476
layers 473 salt bath process 513, 523
mechanism 473, 476 saturation magnetization 41
model 475 ferroxplana-type compounds 404
product 462, 472 Ba/MxFe12-~O19 370-372, 374
sequence 472, 476, 477 BaCoxFe12-xO19-xFx 384, 386
sintering 462, 463, 472, 477 BaCrxFe12-xO19 370-372
thermal quantities 476, 477, 498 BaCuxFel2-xO19-xFx 386
recording heads, integrated 33 BaFz-2FeO-5Fe203 384
reflectance spectra BaGaxFe12-xO19 370-372
BaCo0.sGall.5Oas.sF0.5 386 BalnxFe12_xO~9 376
BaNil.sAI10.5017.sF1.5 386 Bal-x(K or Bi)x(Cu, Ni or Mn)xFe12-~Oa9
B aNi0.sGan.5018.sF0.5 386 383
relaxation, see after-effect BaNixFe12-xO19-xFx 386
resonance linewidth, oxide spinels 257 BaO-6Fe203 325-328, 335, 349, 376, 384,
slowly relaxing ions (impurities) 257 386
reluctance 46 Ba(TiCo)xFela-~ O19 379
remanence BaZnx/2Irx/2Fe12-x019 381
angular dependence 510, 511, 561-563 Ca0.88La0.14Fe12019 366
calculation from texture 540, 562, 563 CaO-AI203--Fe203 375
remanence, temperature dependence 551, LaFe2+Fe~O19 366
579, 580 LaFe12019 366
of alnicos 112 Na0.sLa0.sFe12019 366
of (Ba or Sr)(Cu+ Ge)xFe12-xO19 383 PbO.6Fe/O3 325-328
of (Ba or Sr)(Cu + Nb)xFel2-xO19 383 SrAI4.sFe7.zO19 374
of (Ba or Sr)(Cu + Si)~Fe12-~O19 383 SrAlxFe11-~O19 370-371, 374
of (Ba or Sr)(Cu + Ta)~Fea2-xO19 383 SrCrxFelzOi9 370-372
of (Ba or Sr)(Cu + V)xF'elz-~O19 383 SrO.6FezO3 325-328, 335
remanent magnetization 41, 42, 44 saturation magnetization, temperature depen-
resistance factor 46 dence of
resistivity 751, 752ff, 762, 793, 795 BaO.6Fe203 326-328, 382
anisotropy 752, 753, 779ff, 800, 850 BaSb0.sFe 2+
1.0Fe3+
10.5019 382
high field 765 Ba(SbFe)12019 382-383
low temperature 762 BaSc~Felz-x O19 375
of alloy 766ff, 793ff B aTi0.8Fe2+6Fe3~.8019 382
of amorphous alloys 795 BaTiO3.5Fe203 382
of magnons 757 BaZnxTiyMn~Fel~-~-y-~019 380
of pure ferromagnets 762 PbO'6Fe203 326-327
of single crystals 755 Sr(AsFe)12019 382-383
minimum 7 9 5 f f saturation polarization, magentic 445, 446,
residual 764, 766 462, 499, 536
tensor 752 of powder 487, 488
reversal mode 55 temperature dependence 446, 551
reverse domain Seebeck coefficient
at surface defect 80 values for spinels 269ff
842 SUBJECT INDEX

segment magnets 494, 509, 535 description 191


self cleaning effect 517 inverse 193
shape anisotropy of ionic radii 194ff
alnicos 1-4 11%120 normal 193
alnico 5 and alnico 5DG (alnico 5-7) 131- thermodynamic properties 196ff
133, 145, 149 spinodal decomposition in
alnicos 8 and 9 120 alnicos 1-4 115, 116
FezNiA1 148 alnico 5 126
shrinkage 462, 463, 497, 507, 508, 514 alnico 8 146
ratio 502 alnico 9 146
temperature dependence 502 FezNiA1 115, 116
single crystals 755, 765, 778, 783 spontaneous resistivity anisotropy 752, 800
shunt 560 splat-cooling 522
single domain particle, critical diameter of spray
BaO.6Fe203 335, 355 drying 524
BaZnx/zGex/aFetz-x Oa9 380 wasting 464, 525
BaZn2x/3Nbx/3Fetz-x019 380 stability
BaZn2x/3Tax/3Felz-xOt9 380 chemical 445, 448, 575-577
BaZnzx/sVx/3Felz-x019 380 magnetic 510, 578, 579
PbO.6FezO3 335 natural 545, 549
SrO.6Fe203 335 structural 445, 448, 550
single sintering techniques 513, 514 thermal 453, 461, 545
sintered alnicos 148 standardization 541,545-546, 590
sintering 463, 480, 497, 513 Stoner-Wohlfarth theory of hysteresis in al-
promotion 457, 459, 514 nicos 166
slurry 469 strength values 573, 580
small-defect-width approximation 87, 88 substitution 450, 461
small-deviations approximation 85, 88 of M-type compounds with anions 384-386
small-field approximation 87, 88 suitability criterion 42ff
solid solution 458, 461 super-exchange interactions in spinels 217ff
solid state reaction 468 superconductivity in sulpho- and selenospinels
solubility range, see homogeneity range 623, 626, 627
specific heat 572 superparamagnetic crystals 487
of sulpbo- and selenospinels 624, 626, 628,
672, 700, 708 temperature
specific resistivity 568, 569, 580 compensation 560
specific surface 465 dependence of magnetic properties of al-
spheroid, prolate 60 nicos 180
spin disorder scattering 757ff influence of 100ff
spin dependent Raman scattering 665ff, 689ff tensile strength, BaFex2019 361
spin Hamiltonian thermal
3d" ions with orbital singlets 235 activation of wall displacement 102
values of parameters, oxide spinels 242 agitation 101
spin mixing 759 conductivity 572, 792
spinel crystal structure 609ff excitation 101
cation-anion distances in spinel compounds expansion, BaO-6Fe203 362, 363
613ff, 618 fluctuation 100
lattice parameters of spinel compounds hydrolysis of salts 518
610ff properties, oxide spinel data 288ff
polymorphism in spinel compounds 608 thermoelectric
spinel structure behaviour 569
cation distribution 208ff effect 756, 790
cation ordering 211ff thermomagnetic treatment
crystal energy 206ff Cahn theory of 173
SUBJECT INDEX 843

dependence of magnetic properties on field volumetric feeding 492


direction 151 vulcanization 584, 585
effects on al particles shape anisotropy 149
N6el-Zijlstra theory of 172
of alnicos 121, 125 wall creep 100, 104
of alnico 5DG (alnico 5-7) 129 wall energy
of alnico 8 137 at anti-phase boundary 90
of alnico 8 (extra high coercivity) 141 in applied field 91ff, 95
of alnico 9 142 wall nucleation
relationship between field direction and at antiphase boundary 93ff
preferred direction of magnetization 151 at surface defects 80
thermoplastics 584 wall pinning 81
thermosettings 584, 585 at antiphase boundary 88ff
thin layer techniques 534, 535 at discrete sites 98ff
ticonal 26 at large cavities 76ff
titanium, effects of at line defects 99
in alnicos 155 at planar defects 83ff
topotactic mechanism 475 at point defects 99
trade marks 541, 544, 591 at small cavities 78ff
transport 751ff wall thickness 94
trends in magnetism research 31 parameter 96
two-current model 758ff Weiss field
two-domain state 56ff energy 73
model 69
uniform rotation 49-51, 55, 60, 63, 93 wettability 583
units (SI, cgs) 443 whiskers 64
working point 44, 46
valency of copper ions in sulpho- and seleno-
spinels 618ff
variational calculus 68, 91 X-ray absorption
Verwey transition 264ff of Co ions 714, 721, 735
volume efficiency 579, 580 of Cu ions 620, 632, 642, 721, 735
volume functions of the Fe-Co rich c~l phase X-ray photoelectron spectra of sulpho- and
particles selenospinels 618, 637, 653, 719
determination of the optimum value 164
in alnico 5 135, 136 Young's modulus E, temperature dependence
in alnico 8 146 of BaO.6Fe203 360-361
MATERIALS INDEX

* Me = divalent metal ion, M = magnetoplumbite B7F2 450


type compound B203 498-500, 514, 522-524, 529, 534
B203-BaO-Fe203-GeO2 524
Agl/2All/2Cr2S4 698ff BzO3-Fe203-GeOz-PbO 524
AgxCdl xCr2Se~ 67%681 B203-FezO3-SiOz-SrO 524
AgmGamCr2S4 698ff BSF2 461
AgxHgl-~Cr2Se4 691,692. BaAll2019 370, 385
AgmlnmCr2S4 698ff BaAlo.TFelL3019 358
AgmlnmCr2Se4 698ff BaAlxFelz-xOl9 368-375
Ag0.sLa0.sFe12019 366-367 Ba(CH3COO)2 468, 519, 521
AI substituted M-type compound 354, 368- BaCO3 444, 463, 467, 471-473, 499, 501, 515,
374, 385 516, 519, 520, 522-525, 533, 534
alnico 39, 53, 54, 63, 64, 102, 445, 448, 578-- BaC12 520
580, 582 BaCoFe12-xO19-xFx 385
alnicos 1-4 111 BaCoOaal.5018.sF0.5 386
alnico 5 121 BaCox/2Tix/2Fel2-xO19 378-379
alnico 5DG (alnico 5-7) 129 BaCo0.sZn2AI9.sO~6.sFz5 385
alnico 6 137 BaCoZnAIa0017F2 385
alnico 8 137, 141 Ba-Co--Zn-W 317
alnico 9 137, 142 Ba-Co--Zn-Z 317
A1203 204, 205, 462, 464, 499, 500, 519, 522, BaCo2Fe16027 (Co2W) 403, 434-436
526, 528, 531 Ba2Co2Fe12022 (C@Y) 406, 407, 410, 411, 414,
A1203-BaO-Fe203 454 425, 427, 429
A1EO3-BaO-Fe203-SrO 461 Ba2Co2Fee80~ (Co2X) 409, 414, 418
AI203-Fe2Oa-SrO 458 Ba2CoZnFe120~ (CoZnY) 403, 413
All/2CumCrzS4 698ff Ba2CoZnFez8046 (CoZnX) 418
All/2Cul/2Cr2Se4 698ff Ba3CozFe24041 (Co2Z) 407, 408, 411, 414, 424,
A15/2Lil/204 614 425, 428, 436
ec-FeaO3, see Fe203 Ba3CoL75Zn0.~Fe24041 (Coa.75Zn0.25Z) 414
amorphous alloys 795ff BaCrsFe4019 370
BaCrxFe12-xO19 353, 354, 368-375
B, see BaO, BxFy and BaO-Fe203 Ba(Cu+Ge)xFe12_xO19 383
fl-A1203 444 Ba(Cu + Nb)xFe12-xO19 383
BF 449-456, 458, 461, 473, see also BaFe203 Ba(Cu+Si)xFe12_xO19 383
and BaO.Fe203 Ba(Cu + Ta)xFe12-xO19 383
BF2 450, 454, 455, see also BaO.2Fe203 and T Ba(Cu + V)xFei2-xO19 383
B2F 449-451, 473, see also Ba2Fe205 BaCuxFe12-xO19~ ~Fx 385
B2F3 453-455, 473 Ba2Cu2Fe12022 (Cu2Y) 435
BsF7 453 BazCu2Fe~O4~ (Cu2X) 411

845
846 MATERIALS INDEX

Ba3Cu2Fe24041 (Cu2Z) 411 BaO-6AI203 444, 535


BasCuNiTi3Fe12031 (CuNi-18H) 411,420, 421 BaO.6.6A1203 535
BaFeO3_x 308 BaO-CaO-Fe203 454
BaFe204 453, 454, 518, 534, see also BF and (BaO)x (CaO)l-x- nFe203 457
BaO.Fe203 BaO-Fe203 449, 451~454, 458, 522
BaFe22+Nlo.58Fe15.42018.42Fo.s8
- 3+ 385 BaO-FeO-Fe203 454
BaFe2Fet6027 (Fe2W) 403, 406, 407, 411, 422, BaO-FeO-7Fe203 308, 453, 454, 456, see also
425-427, 434 BaFelsO23, BaO-MeO.7Fe203 and X
BaFe42A17.8019 454 3BaO.4FeO- 14Fe203 454
BaFelo.92017.38 568 BaO.FeO.3Fe203 308
BaFe12019 433-448, 451, 452, 454, 455, 457, BaO.Fe203 (B) 308-310, 449-454, 456, 458,
479, 519, 522-524, 534, see also, 468, 472, 474, 488, 522, see also BaFe204
BaO.6Fe203, BaM and M and BF
BaFe12Olg-Na20 524 BaO.2Fe203 (T) 308, 309, 316, 321, 456, see
BaFe12.59019.89 568 also BF2 and T
BaFe15023 453, see also BaO.FeO.7Fe203, BaO.4.5Fe203 452
BaO-MeO.7Fe203 and X BaO-5Fe203 452
Ba2Fe2Os 534, see also B2F BaO-5.3Fe203 498
BaGa12019 370, 375 BaO@Fe203 453
BaGaxFe12-xO19 368-375 BaO.5.5Fe203 529
Ba-hexaferrite 55 BaO-5.6Fe203 484
BaInxFe12-xO19 375-378 BaO.5.9Fe203 498
Bal_x(K or Bi)x(Cu, Ni or Mn)xFe12-x019 383 BaO.6Fe203 30%366, 444, 449-453, 456, 461,
Bal_xLaxMnO3 616 464, 468, 472, 475, 488, 522, see also
BaM 450, 536, 537, 551-553, 556-575, see also BaFe12019 and BaM
BaFe12019, BaO-6Fe203 and M BaO-nFe203 449, 453
Ba-Me-U 307, 309 2BaO.Fe203 308
Ba-Me-W 307, 309, 311 2BaO.3Fe203 308, 312, 313
Ba-Me-X 307, 309, 312 3BaO.Fe203 308
Ba-Me-Y 307, 309 5BaO.Fe203 308
B a - M e - Z 307, 309 5BaO.TFe203 308
BaMg2Fe16027 (Mg2W) 405, 421-423 7BaO.2Fe203 308
BaMn2Fe16027 (Mn2W) 411 3BaO.4FeO. 14Fe203 308
Ba2Mg2Fe12022 (Mg2Y) 406, 407, 411, 414, 429 BaO.Fe203-Fe203-S 456
Ba2Mn2Fe12Oz~ (Mn2Y) 406, 407, 411 BaO-Fe203-MeO 454, 456
Ba2MnZnFe12022 (MnZnY) 407, 425 BaO-Fe203PbO--SrO 461
BasMg2Ti3Fex2031 (Mgz-18H) 410, 411, 420, BaO-Fe203-SiO2 457
421 BaO-Fe203-SrO 461
BasMgZnTi3Fe12031 (MgZn-18H) 420 BaO-Fe203-ZnO 454
Ba(NO3)2 468, 520, 521, 528 BaO-2Fe203-8Fe203 (Ba-Fe-W) 308, 311,312
BaNil.sAlmsO17.sF15 386 BaO2 468
BaNio 5Gan.sOls.sFo5 386 Ba(OH}2 468, 488, 517, 519, 521
BaNio.sZn2A19.5016.sFz5 385 BaO-MeO-Fe203 307, 308
BaNixAl12-xO19-~Fx 385 BaO.MeO-3Fe203 456, see also Y
BaNixFel2 xO19-xFx 385 BaO.MeO.7Fe203 456, see also X
BaNi2Fe16027 (Ni W) 405 BaO.2MeO-8Fe203 456, see also W
BaNi2Al~Fe16-xO27 (NiA1W) 435 2BaO-MeO-9Fe203 456, see also U
BaNiFeFe16027 (NiFeW) 406, 411 2BaO.MeO- i3Fe203 456
BaNio.sZno5FeFea6027 (Nio.sZno5FeW) 406 3BaO-MeO- 19Fe203 456
Ba2NiZnFe12022 (NiZnY) 413 3BaO.2MeO.12Fe203 456, see also Z
Ba2Ni2Fe120~ (Ni2Y) 406, 411, 414 BaO-PbO-Fe203 367, 368
BaO 449, 450, 454, 455, 467, 468, 472, 473, BaO-SrO-Fe203 367, 368
488, 522-524 BaSO4 519, 523
BaO.4.6AI203 535 Ba(SbFe)12019 382
MATERIALS INDEX 847

B aSbo.sFeZ+oFe3~.5Oa9 382 CaSiO3 503


BaSc~Fe~z-xO19 375 CdxCol-xCr2S4 717
Bao.2Sro.sFe12019 575 CdCr2S4 607, 612-615, 647,-650-654ff, 675
Bao.4Sro.6Fea2019 575 CdCr2Se4 697, 608, 613-615, 641, 647, 650-
Bao.6Sro.4Fe12Oa9 575 653, 675ff
Bao.vsSro.zsFe12O19 367 CdCr2(S~-xSe~)4 696
Bao.sSr0.2Fe12019 575 CdCrz-xlnxS4 656
Bal-xSr~Fea2Oa9 443, 461 CdCrl.sTi0.2S4 656
(Ba, Sr)(Fe, A1, Ga)12019 535 CdCrI.sV0.2S4 656
BaTiO3-5Fe203 380 Cdl-xCuxCr2Se4 679
• 2+
BaT10.6Fe 0.?,Fe310.8019
+
382 CdFe204 260, 286
BaZn2A12Fe12027 405 CdxFel-xCr2S4 656, 707, 715-717
BaZn2AIxFe12-~O27 405 Cdx_xFexCr2Se4 677
BaZn2Fe16027 (Zn2W) 404, 412, 422 Cd0.98Ga0.02Cr284 657
BaZnFeFe16027 (ZnFeW) 406, 411 Cdl-xGaxfr2Se4 678, 679, 682
Ba2Zn2Fe1202z (Zn2Y) 404, 407, 410, 411, 414, Cdl-xHgx Cr2S4 694
425, 427, 429 Cdl-xHg~Cr2Se4 696
Ba3Zn2Fe24041 (ZnzZ) 407, 408, 411, 429, 430, Cdln2S~: Cr3+ 659, 662
433 Cdl-xlnxCr2S4 655, 656
Ba2Zn2Fe28046 (Zn2X) 409, 411,414, 416, 418, Cdl_xlnxCr2Se4 652, 675, 680-683, 686
433 Cdo.sluo.2Cr1.80C00.2S4 656
Ba4Zn2Fe36Oeo (ZnzU) 409, 410, 414, 419, 430, Cdo.8In0.2Cra.80Ni0.2S4 656
433 CdMnzO4 215
Ba4Zn2Fe52084 401 CdxZn~-xCr2Se4 695
BaZn2GaFexsO27 405 CI 465, 466
BaZn2Ga3Fe13027 405 Co 52, 762ff, 773, 776ff, 789, 790
BasZn2Ti3Fe12031 (Zn2-18H) 410, 411, 419- Co alloys 766, 768, 771,772, 776, 799
421 Co-steel 45
BaZn~/2Gex/2Felz-x019 380 Co~ 467, 472, 473, 476, 477, 498, 501, 518, 520,
BaZn~/2Ir~/2Fe12-~O19 380, 381 523
BaZnz~/3Nb~/3Fe~z xO19 380 Co-Cr, Co-Ti and Cu-Ti substituted M-type
BaZnzx/3Tax/3Felz-xOa9 380 compounds 353, 354
BaZn~/zTi~/2Fe12 ~O19 381 COA1204 225, 261
BaZnxTiyMn~Felz-x-r-zO~9 380 CoCo2-2xMn2xOa 216
BaZn2~/3V~/3Fea~-~O19 380 CoCrz-xlnxS4 725
Ba-Zn-W 317 CoCr2-2xMnzxO4 216
Ba-Zn-Y 317 CoCr204 284, 286
Ba-Zn-Z 317 CoCrxRh2-xS4 729-731
Bi~O3 500, 519, 522 CoCr2S4 611, 613-615, 701, 703, 711ff
CoCr2S4-xSex 726
C-steel 45 ColaCrl.sSn0.1S4 714
CO 477 COl_xfuxfr2S4 721,728
CsH~2 477 Coi-x (CUl/2Fel/z)xCr2S4 718
CaAI~20~9 370 COl-xCuxRh2S4 728, 734, 735
Ca(AIFe)~20~9 370 Co~Fex-~Cr2S4 707, 714
CaFesO13 454, 459 CoFe204 196, 197, 220, 231, 242, 258, 276,
Cao.88Lao.~4Fe~zO~9 366, 367 286, 288, 289, 291,296, 298
Ca~_xLaxMnO3 616 CoxFe3-xO4 248, 250-252, 269, 273, 282, 294
CaO 461, 467, 519 CoFe204 :Ti 270, 271
CaO-6AI~O3 444, 475 COl-xFe~Rh2S4 729, 730
CaO-AI~O3-Fe203 375 CoFeVCr 578, 582
CaO-Fe~O3 462 Co2GeO4 195, 197, 199
CaO-Fe~O3-SrO 458 Coln2S4 723
(CaO)~_~(SrO)x •nFe~O3 459 CoMn204 215
848 MATERIALS INDEX

CoxMnl-xFe204 244 CuxFea-xRh2S4 608, 698, 701, 728, 732, 733


CoNiZn ferrite 251 Cul/zGamCr2S4 698ff
COl-xNixCrzS4 722 Cul/zGal/zCrzSe4 698ff
Co304 220, 614, 626 CuGa204 2O5
CoRhl.5Feo.504 284 ChaxHgl-xCrzSe4 691
CoRh2S4 608, 612, 614, 615, 728 CumInmCr2S4 622, 698ff, 715
Co354 612, 614, 728, 736 Cu0.5+xIn0.5-xCr2S4 700
CoS7 616 CumInmCr2Se4 641, 698ff
Co7/3Sb2/304 195 (C'umlnl/2)xFel_xCr2S4 715
CozSiO4 199 CuxMl-xCr204 (M = Cd, Co, Mg, Zn) 216
CozTiO4 197 CuMg0.sMnl.504 225
Co0.zZnl.sSnO4 205 CuxMnl-~Cr2S4 718
Coo.2Znl.sTiO4 205 Cul-xNixCr2S4 722
Cr substituted M-type compounds 353, 354, CuNiFe 578, 582
368-375 CuNi0.sMnl.504 195, 225
CrAI2S4 656 CuO 620
CrFeCo 578, 579, 582 Cu20 620
Crln2S4 656 CuRhMnO4 195
CrMn204 213 CuRh204 215, 225
CrO2 638 CuRh2S4 612, 614, 615, 618, 619, 625, 626ff,
CrTe 616 641
Cu 620 CuRh2Se4 613-615, 618, 619, 615~17ff, 641
CuA1Sz 618 CuRh2(Sl_xSe~)4 626
CuA1204 205 CuRh2_~SnxSe4 624, 627ff
CuCO3 620 CuSO4-5H20 620
CuC1 620 Cu2Se 620
CuC12.2H20 620 CuTi2S4 610, 614, 615, 619, 620tt, 623, 625
CuCozS4 612, 614, 615, 619, 624ff, 736 CuV2S4 610, 614, 615, 618, 619, 621, 622ff, 625
CUCOxRhz-xS4 626ff Cu0.2Zn0.4Cd0.4Al204 205
CuCoTiS4 624ff Cu~Znl_xCr2Se4 672, 673
CuCr204 215, 224, 286, 618, 620 CuZn ferrite 275
CuCr2S4 607, 608, 611-615, 618-620, 630ff, 701 CuZnGeO4 205
CuCr2S4-xClx 644
CuCr2S4_~Se~ 643 Dy 98
CuCr2Se4 607, 608, 612-615, 618-620, 630, DyNi3 799
631, 636tt Dy3A12 98
CuCr2Se4-xBr~ 645, 673
CuCr2Sea-xClx 644 F, see Fe203, BxFy, and PxFy
CuCr2Se4_~Te~ 644 Fe 53, 56, 63, 64, 77, 78, 80, 762ff, 773, 775,
CuCr2Te4 607-609, 613-615, 618-620, 630, 776tt, 788, 790, 792
631, 637, 641ff Fe alloys 766, 769, 771, 772, 779, 783, 787ff,
Cu~+~Cr2Te4 641ff 797tt
CuCr2Te4_~I~ 645 Fe(CI-I3COO}2 519, 521
CuCr2-~RhxSe4 636ff FeC204 519, 521
CuCr0.3Rhl.7-xSn~Se4 636ff Fe(CO)5 467, 534
Cufr2-x SnxS4 630ff FeCI2 464
CUfl'l.9Sn0.1Se4 641 FeCI3 520, 524
CuCr2 ~TixS4 630tt FeCo 64
CuCr2-~V~S4 630ff Fe~Col_~Cr204 216
CuFe204 213, 215, 243, 258, 259, 269, 289, FeCr2-xInxS4 724, 725
291, 296, 618 FeCr204 215, 285, 287
CuFel.vCr0.304 213 Fe3-xCrxO4 214
Cu~Fel xCr2S4 698, 710, 718ff, 728, 732, 733 FeCrzS4 607, 608, 611,613-615, 701, 702, 706ff
Cul/2Fel/2Cr2S4-~Se~ 726 Fel+xCr2-xS4 726
MATERIALS INDEX 849

FeCr2S4-xSex 726 Inl/2LimCr2S4 698ff


FeCrxRh2_x$4 730
FeL1CrI.sSn0.1S4 711 KCI 523, 524
Fel-sCusFel+sCuz~Ml-2~-804 (M= Co, Mg, KOH 524
Ni) 216 K2SO4 522, 523
Fe2GeO4 199, 286
FeIn2S4 723 La(Co, Ni)5 98
FeMnzO4 215 LaFe2+Fe3i~O19 366, 367
Fe(NO3)3 519-521 LaFe12019 354, 366, 367, 368
Fe-Ni-A1 alloy 45 LaMgGanO19 369, 370, 375
Fe2NiA1 111 La203 462
Fel-~Ni~CrzS4 722 LaxSrl_xMnO3 616
FeO 201 Lio.sAlzsO4 204, 219, 286
FeO2 459 LiCO3 500
FeOOH 515, 521, 525 LiF 476
Fe203 195, 201, 444, 448-469, 471-477, 487, Lio.5Feo.sCr204 212
499, 514-516, 519, 520, 522-525, 528, 533, LiFeO2 476
534 LiFesO8 476
Fe203-2(FeO2)-SrO 459 Lio.sFezsO4 195, 212, 220, 229, 231, 235, 238,
Fe203-MeO-SrO 458 242, 251, 257, 260, 269, 271, 277, 282, 283,
Fe203-PbO 449, 459-461, 522 287-289, 291, 296, 298
Fe203-SiO2-SrO 459 Lio.sFe2.504:Mn 255
Fe202~SrO 449, 457, 458, 522 Lio.sGazsO4 206, 219
Fe203-SrO-ZnO 458 Lio5Mnz504 195
Fe304 196, 197, 200, 201, 212, 213, 220, 231, LiV204 195, 261, 262
232, 235, 242, 258, 259, 262ff, 269, 270, 277,
278, 286, 287, 289, 291, 292, 296, 311, 451, M 449-458, 460, see also BaFel2019, SrFe12O19
452, 457, 458, 462 and PbFe12019
Fe304: Mn 255 MS 456, see also W
Fe304 : Zn 254 M2S 456, see also X
FeRhzS4 612, 729 MaS 456
FeS2 465, 467, 701 M6S 456
FeSO4 525 MY 456, see also Z
Fe2SiO4 199 M2Y 456, see also U
Fe3S4 726 2MeO-BaO-8F~O3 (2MEW) 307, 312, 316
FeTiO3 467 2MeO.2BaO.6FezO3 (3MeY) 307, 309, 316
Fe2TiO4 195, 227 2MeO.3BaO.12Fe203 (2MeZ) 307, 309, 316
Fe3-xTixO4 244 2MeO.2BaO-14Fe203 (3MeX) 307, 309, 312
FeVzO4 196, 215, 262 2MeO.4BaO.18FezO3 (MeU) 307, 309
MeO.Fe203 309, 321,456
Gal/2Li1/2Cr284 698ff MgAI204 191, 197, 204, 206, 219, 260, 261,
Gd 775 284, 286, 287, 289, 295
GeCo204 224 MgCr204 225, 245, 648
GeFe204 206, 224, 225 Mga-xFexAleO4 290
GeNi204 224, 225 Mg ferrite 252, 253, 270, 295
MgFe204 197, 209, 210, 220, 239, 242, 251,
HC1 464-466, 471 257-259, 277, 279, 283, 288, 289, 291, 296,
H2504 466 298
hexaferrite 39, 49, 55 MgFe204 : Mn 272
HgCr2S4 607, 612-615, 650, 652, 666ff MgGa204 375
HgCr2Se4 607, 613-615, 650, 652, 691ff MgxMno.6Fez4 xO4 239, 242
HgCrz-xInxSe4 691, 692 MgMn204 210, 215, 269
Hgl xInxCrzSe4 692 MgO 201, 205, 499
HgxZnl xCr2Se4 696 Mg2SiO4 199
850 MATERIALS INDEX

Mg2TiO4 196, 204 NiCrxRh2 xS4 731, 732


MgV204 225, 262 NiFeCo ferrite 251, 272
Mg2VO4 199 NiFe, Cr2_~O4 213
MnA1 55, 89, 94 NiFe204 195-197, 235, 242, 257-259, 275-277,
MnAIC 578, 580 279, 281, 286, 288, 289, 291, 296, 298
MnA1204 220, 225 NixFe3-xO4 252, 253, 258, 259, 269-271, 282,
MnBi 578, 579 292, 293
MnCr ferrite 255 NiFe2 xVxO4 228
MnCr204 196, 230 Ni2GeOa 199, 286
Mnl+2xCr2_2xO4 214, 216, 246 Niln2S4 723
IVlnCrz-xInxS4 723 NiMn204 210
MnCr2S4 607, 611, 613-615, 701ff, 724 NiO 281, 499
MnCr2S4-xSe~ 726 NiRh204 215, 225
MnCr2-~VxS4 725 NiRh2S4 612
MnFeCrO4 230 Ni2SiO4 199
MnFe204 196, 197, 209, 210, 230, 233, 242, Ni-Ti substituted M-type compound 354
258, 259, 269, 276, 286, 289, 291, 296, 298 NiZnCo ferrite 251
Mn~Fe3-xO4 213, 220, 238, 250, 253-256, 259, NixZnl-xCr204 285, 287
260, 269-271, 273, 278, 280, 282, 293, 294 NixZnl-xCr2S4 721, 722
Mnln2S4 722, 723 NiZn ferrite 243, 270, 271, 275-277, 290
MnMg ferrite 255, 259
Mnl-xNixCrzS~ 721,722 P, see PbO and PxFr
Mn203 195, 215, 466 PF2 460, 461, 477
Mn304 215, 224, 286 PzF 460, 461, 477
MnRh204 225 PbAlxFe12-xO19 374
MnSb 616 Pb(C2H5)4 534
MnV204 245, 262 PbCO3 529,
Mn0.6Zn0.4Fe204 220 PbF2 529
MnZn ferrite 255, 258, 259, 275 PbFe7.sMn3.sAlii.sTi0.sO19 444, 574
MoAg204 195 Pb2FelsMn7(A1Ti)O38 307
PbFe12019 443, 534, s e e a l s o PbO-6Fe203,
NH3 518, 520 PbM and M
NI-I40H 517, 519, 520 PbM 450, 536, 537, 551, 569, 574, 575, s e e a l s o
(NH4)2CO3 517, 519, 520 PbFe12019, PbO.6FezO3 and M
Na2CO3 517, 519, 510, 522, 523 Pb(NO3)2 520
NaC1 523, 524 PbO 449, 450, 460, 461, 468, 471, 477, 507,
NaF 476 514, 522, 529
NaFeO2 522, 523 PbO-2Fe203 313, 315
Na20 524 2PbO.2Fe203 314, 315
Na0.sLa0.sFe12019 366, 367 PbO.2.5Fe203 461
Na2Mn2Si207 204 PbO.!~Fe203 461
Na20.11AIzO3 444 PbO-5Fe203 315, 459, 461,501
NaOH 517, 520, 521, 524 PbO.6Fe203 307-366, 444, 449, 450, 461, 464,
NaSO4 522, 523 477, 527, 532
NazWO4 261 PbO-nFe203 449
Nd203 462 Pd 790
Ni 762ff, 764, 773-776ff, 783, 788, 790, 792 Pd based alloys 775, 793ff
Ni alloys 766ff, 772, 775, 777ff, 781ff, 786ff, Pt based alloys 795
791,796ff PtCo 578, 579
NiAI204 261
NiCo ferrite 251 RE-alloys 49
NiCo.eS4 612, 614, 736 RE-Co 580, 582
NiCr2-xInxS4 725 RECo5 578, s e e a l s o SmCos
NiCr204 205, 215, 285, 287 RE(Co, Cu, Fe, Mn)x 578, 579
MATERIALS INDEX 851

S 456, s e e SrO and SxFy SrO.4.2Fe2Oy 1.8AlzO3 355


SF 458, 459, 476 SrO-4.5Fe203.1.5A1203 355
SF6 459, s e e a l s o SrFe12019 and SrO.6Fe203 SrO.5.5Fe203 501
S2F 457, 459, 476 SrO.5.9FeeO3 501
S3F 457, 459 SrO-6Fe203 307, 366, 444, 449, 461, 464, 476,
$3F2 458, 459 522, s e e a l s o SrFe12019, SrM and M
$4F3 457-459, 477 SrO.nFe203 449
87F5 457-459, 461,477, 522 3SrO.2Fe203 313, 314
SO2 467 Sr0.75Pb0.25Fe12019 368
SiC 526, 531 SrO-PbO-Fe203 367, 368
Si3N4 527 SrSO4 519, 520, 523
SiO2 457, 459, 462, 464, 465, 467, 498-503, Sr(SbFe)12019 382
514, 519, 522
SmCo5 45, 55, 67, 81, 83, 88, 103, 490 T 456, s e e a l s o BaO.2Fe203 and BF2
SmCo5.3 81 Th(Co, Ni)5 98
Sm2Co17 83 Ticonal G 45
Smz(Coo.85Feo11Mn0.04)17 45 Ticonal GG 45
SnO2 499 Ticonal II 45
SrAI3.8Fes.20|9 374 Ticonal XX 45, 54
SrAl4.aFe7.2Ol9 374 TiFe204 235, 285, 287
SrAl12019 370, 458, s e e a l s o SrO-6A1203 TiO2 464, 499, 519
SrAlxFe12 1019 355, 368-375 T10.sLao.sFe12019 366-368
Sr(AsFe)12019 382
SrCH3(CH2)10COO2 477, 519 U 454, 456, 473, see 2MeO.4BaO. 18Fe203
SrCO3 463, 467, 476, 477, 515, 516, 519, 520,
522-525, 533 V205 499
SrCr6Fe6019 370
SrCrxFe12-xO19 368-375 W 450, 451, 454, 456-458, 461, see also
Sr(Cu + Ge)xFe~2 xO19 383 2MeO-BaO-8Fe203
Sr(Cu+Nb)xFe12_xO19 383 W-steel 45
Sr(Cu+ Si)~Fe12-xOa9 383
Sr(Cu+Ta)~Fe12-xOa9 383 X 450-454, 456-458, 461, see also
Sr(Cu+V),Fe12 xO~9 383 2MeO.2BaO- 14Fe203
SrFeO3 x 314, 458, 459, 476, 477
SrFesA14019 535 Y 454, 456, 461, 473, s e e a l s o
SrFel2019 443, 448, 457-459, 479, 519, 522- 2MeO-2BaO.6Fe203
524, s e e a l s o SrO-6Fe203, SrM and M Y(Co, Ni)s 98
SrFe18027 457, 458, s e e a l s o W YCos 81
Sr-Fe-W 311, 313, 314 Y3AI5012 204
Sr-Fe-X 313, 314 Y3FesO12 251, 277-279, 281-283
Sr2FeO4-/ 459
Sr3Fe207-~ 459 Z 454, 456, 473, s e e a l s o
Sr4Fe3Om-x 459 2MeO-3BaO. 12Fe203
SrGaxFe12-~O19 368-375 ZnAlxCr2 xO4 285
SrGal2019 370, 375, 535 ZnA1204 242, 260, 261
Sr-hexaferrite 55 ZnCo204 626
SrM 450, 536, 537, 551-553, 566, 570, 571, ZnCr204 196, 285, 287, 608, 614, 647
573-575, s e e a l s o SrFe12019, SrO-6Fe203 ZnCr2-2,Mn2xO4 216
and M ZnCr2S4 608, 611, 613-615, 647, 652, 653ff
Sr(NO3)2 519, 520 ZnCr2Se4 608, 613-615, 647, 649, 650, 652,
SrO 449, 450, 457-459, 467, 468, 519, 522 653, 669tt
SrO.6A1203 4 4 4 , s e e a l s o SrAl12019 ZnCr2(Sl-xSex)4 696
SrO.2FeO-8Fe203 311-314, 457, s e e a l s o W ZnCr2(Sea-xTe,)4 696
SrO-FeO.7Fe203 458, s e e a l s o X ZnFe2-2xMn2xO4 216
852 MATERIALS INDEX

ZnFe204 195--197, 225, 260, 285, 286, 288, 289, Zn~-xMnxCr2Se4 673
298 ZnMn204 215, 225, 286
ZnxFe3 xO4 270, 271 ZnMn2Te4 609
ZnGa204 204, 206, 219, 242, 284 ZnNbLiO4 195, 212
ZnGa204 : Cr 3+ 648, 651 ZnNiSnO4 205
Zn-Ge, Zn-Ti and Zn-Zr substituted M-type ZnO 534
compounds 353, 354 ZnRh204 285
Zn2GeO4 198, 199 Zn2TiO4 204
ZnLiSbO4 212 ZnV204 262
ZnxMl-xFe203 (M=Co, Fe, Li0.sFe0.5, Mn, ZrO2 526, 529, 530
Ni) 228

You might also like