Download as pdf or txt
Download as pdf or txt
You are on page 1of 164

chapter 4

MAGNETIC AMORPHOUS ALLOYS

P. HANSEN
Philips GmbH Forschungslaboratorium Aachen
D-5100 Aachen, Germany

Handbook of Magnetic Materials, Vol. 6


Edited by K. H. J. Buschow
© Elsevier Science Publishers B.V., 1991

289
CONTENTS

I. Ir t r o d u c t i o n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292
2. P r e p a r a t i o n m e t h o d s . . . . . . . . . . . . . . . . . . . . . . . . . . 293
2.1. L i q u : d q u e n c h i n g . . . . . . . . . . . . . . . . . . . . . . . . . 293
2.2. Vapor q u e n c h i n g . . . . . . . . . . . . . . . . . . . . . . . . . 295
3. Structure of a m o r p h o u s allo) s . . . . . . . . . . . . . . . . . . . . . . . 297
3.1. C o m p o s i t i o n a n d h o m o g e n e i t y . . . . . . . . . . . . . . . . . . . . 297
3.2. D e r sity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298
3.3. M i c r o s t r u c t u r e . . . . . . . . . . . . . . . . . . . . . . . . . . 301
3.4. A l o m i c structure . . . . . . . . . . . . . . . . . . . . . . . . . 303
4. S | a b i l i t y c f a m o r p h o u s al:oys . . . . . . . . . . . . . . . . . . . . . . . 307
4.1. C r y s t a l l i z a t i o n t e m p e r a t u r e s . . . . . . . . . . . . . . . . . . . . . 307
4.2. S t r t c t u r a l r e l a x a t i o n . . . . . . . . . . . . . . . . . . . . . . . . 313
5. M a g n e t ! c properties . . . . . . . . . . . . . . . . . . . . . . . . . . 315
5.1. M a g n e t i c structures . . . . . . . . . . . . . . . . . . . . . . . . 315
5.2. Influence of p r e p a r a t i o n c o n d i t i o n s . . . . . . . . . . . . . . . . . . . 319
5.3. L o w - t e m p e r a t u r e m a g n e l i c m o m e n t s . . . . . . . . . . . . . . . . . . 320
5.3.1. O n e - s u b n e t w o r k alloys . . . . . . . . . . . . . . . . . . . . . 320
5.3.1.1. Transition-metal-ba~,ed alloys . . . . . . . . . . . . . . . . 320
5.3.1.2. R a r e - e a r t h - b a s c d alloys . . . . . . . . . . . . . . . . . . 333
5.3.2. T w o - s u b n e t w o r k alloys . . . . . . . . . . . . . . . . . . . . . 339
5.4. T e m p e r a t u r e c'epcndence c f l i e m a g n e t i z a t i o n . . . . . . . . . . . . . . . 344
5.4.1. M o d e l s . . . . . . . . . . . . . . . . . . . . . . . . . . 344
5.4.1.1. Mean-field t h e o l y . . . . . . . . . . . . . . . . . . . . 344
5.4.1.2. d - b a n d m o d e l with r a n d o m axial a n i s o t r o p y . . . . . . . . . . 347
5.4.1.3. O t h e r c c n c e p t s . . . . . . . . . . . . . . . . . . . . . 348
5.4.2. O n e - s u b n e t w o r k alloys . . . . . . . . . . . . . . . . . . . . . 349
5.42.1. Transitfon-metal-ba,,ed alloys . . . . . . . . . . . . . . . . 349
5.42.2. R a r e - e a r t h - b a s e d alloys . . . . . . . . . . . . . . . . . . 354
5.4.3. T w c - s u b n e f w o l k alloys . . . . . . . . . . . . . . . . . . . . . 356
5.5. Critical ~)pex~erts . . . . . . . . . . . . . . . . . . . . . . . . . 369
5.6. U n i a x i a l m a g n e t i c an:'sctropy . . . . . . . . . . . . . . . . . . . . . 373
5.7. M a g n e t o s t r i c t i o n . . . . . . . . . . . . . . . . . . . . . . . . . 379
5.8. C o e r c i v i l y . . . . . . . . . . . . . . . . . . . . . . . . . . . . 385
5.9. M a g n e t i c e x c i l ~ t i e r s . . . . . . . . . . . . . . . . . . . . . . . . 392
5.10. A n t ealing effects . . . . . . . . . . . . . . . . . . . . . . . . . 395
6. M a g n e t o - c p t i c a l p r o p e r t i e s . . . . . . . . . . . . . . . . . . . . . . . . 398
6.1. Phenomeno~ogical t h e o r y . . . . . . . . . . . . . . . . . . . . . . 398

290
MAGNETIC AMORPHOUS ALLOYS 291

6.2. Microscopic models . . . . . . . . . . . . . . . . . . . . . . . . 399


62.1. ]nlerbal~d t r a n s i t i o n s . . . . . . . . . . . . . . . . . . . . . 399
62.2. Tntraband t i a n s i t i o r s . . . . . . . . . . . . . . . . . . . . . 401
6.3. O n e - s u b n e t w c r k alloys . . . . . . . . . . . . . . . . . . . . . . . 402
6.4. T w o - s u b n e t w o r k alloys . . . . . . . . . . . . . . . . . . . . . . . 405
7. Tran,,port p r o p e r t i e s . . . . . . . . . . . . . . . . . . . . . . . . . . 413
7.1. Resistivity a n d m a g n e l o r e s i s t a n c e . . . . . . . . . . . . . . . . . . . . 413
7.2. H a l l ~ffect . . . . . . . . . . . . . . . . . . . . . . . . . . . . 420
8. T e c h n o l o g i c a l a p p l i c a t i o n s . . . . . . . . . . . . . . . . . . . . . . . . 423
8.1. M~tallic glasses . . . . . . . . . . . . . . . . . . . . . . . . . . 423
8.2. M a g n e t o - c p t i c a l r e c o r d i n g . . . . . . . . . . . . . . . . . . . . . . 425
8.2.1. Storage ~ri~ciple . . . . . . . . . . . . . . . . . . . . . . . 425
8.2.2. M a t e r i a l ~,e?ection . . . . . . . . . . . . . . . . . . . . . . . 428
8.2.3. T h e r m c m a g n e t i c ~witching process . . . . . . . . . . . . . . . . . 429
8.2.4. M a g n e t o - o p t i c a l d!sk . . . . . . . . . . . . . . . . . . . . . 432
8.2.5. Recorder r e q u i r e m e n t s . . . . . . . . . . . . . . . . . . . . . 432
9. S u m m a r y . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 435
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 436
1. Introduction

Amorphous alloys, sometimes called metallic glasses or glassy metals, have received
much attention in the last decade due to their interesting properties and their
applicability for new devices. These metallic alloys are characterized by a structural
disorder where each atom constitutes a structural unit. This leads to different behavior
with respect to their electronic and magnetic properties as compared to their crystal-
line counterparts where the lattice periodicity and the crystal symmetry are the
dominant elements controlling the basic features. In particular, the magnetic proper-
ties of amorphous alloys are strongly affected by the bond and chemical disorder
causing a distribution in the magnetic moments and exchange interactions. Also, the
random electrostatic fields create local anisotropies via spin-orbit coupling, giving
rise to a varying orientation of the magnetic moments. The structural disorder
introduces new magnetic structures such as speromagnetism, asperomagnetism and
sperimagnetism which posesses a noncollinear arrangement of the magnetic moments.
The existence of competing positive and negative exchange coupling leads to spin-
glass behavior in many alloys. Also, the transport properties differ considerably from
those of the corresponding crystalline materials, which becomes clear from the large
extraordinary Hall effect. The new effects and phenomena associated with amorphous
alloys have led to new theoretical concepts and a tremendous amount of experimental
work in this field which gave new insights in the understanding of disordered systems
with respect to both structure and magnetism. Further, the investigation and develop-
ment of the preparation techniques have led to very homogeneous and ultra-thin
films and have opened a new field in research on interface physics and multilayer
effects.
The interesting properties of the amorphous alloys have been tailored for various
technical applications. Alloys with low coercivity can be used for cores in distribution
transformers or alloys with a high permeability for magnetic shielding. Alloys for
magnetic power controllers, force transducer and magnetic sensors have been investi-
gated. Amorphous rare-earth-transition-metal alloys show suitable magnetic and
magneto-optical properties and have been developed as a medium for erasable,
optical data storage and recording. The class of amorphous alloys include a large
variety of different materials ranging from nonmagnetic alloys via spin glasses and
transition-metal-metalloid glasses to rare-earth-transition-metal alloys. Therefore,
it appears to be almost impossible to present a complete survey of the tremendous
amount of data, effects and ideas reported in the past decade. In this chapter, we

292
MAGNETIC AMORPHOUSALLOYS 293

will focus on the magnetic amorphous alloys with special attention of the rare-earth-
transition-metal alloys which have been recently investigated in more detail. It will
be attempted to present a selection of data and theoretical concepts that are relevant
with respect to the influence of the structural disorder on the magnetic properties of
these alloys. Further extensive information can be found in the reviews on the
structure and the magnetic properties (Cahn 1980, Chen 1980a, Handrich and Kobe
1980, Luborsky 1980, Cargill III 1981a,b, Mizoguchi 1981, Buschow 1984a, Egami
1984, Moorjani and Coey 1984, O'Handley 1987a). Data of amorphous alloys are
compiled by different authors (Ferchmin and Kobe 1983, Buschow 1984a, Hansen
1988a).
This contribution on amorphous alloys is organized as follows: The section
following this introduction deals with a brief description of the most common
preparation methods involving the liquid-quenching and vapor-quenching methods.
In the next section, some basic aspects of the structure of amorphous alloys are
treated in terms of the radial distribution function leading to information on the
local atomic arrangement. In section 4, some models describing the stability of the
amorphous alloys are introduced. The activation energies associated with a thermally
activated rate process are related to the crystallization temperatures. Also, the struc-
tural degradation and effects of oxidation are discussed. A survey on the magnetic
properties is given in section 5. Various experimental data on the composition and
temperature dependence of the magnetic moment, the anisotropy or coercivity are
presented, which demonstrate the strong influence of the structural disorder. The
experimental data are compared with theoretical results inferred from calculations
based o n the mean-field theory. The magneto-optical properties are presented in
section 6. The spectral dependences are interpreted in terms of the relevant band
transitions, and typical dependences of the Faraday rotation on composition and
temperature are discussed. The Kerr rotation plays a special role in the case of the
rare-earth-transition-metal alloys used for optical data storage and recording.
Section 7 deals with the transport properties of amorphous alloys, which are com-
pletely different as compared with their crystalline counterparts. The high resistivity
and the large extraordinary Hall effect reflect the high scattering probability of the
conduction electrons. Some technical applications of amorphous alloys are consid-
ered in section 8. The strong interest in the field of optical recording was taken into
account by presenting a more extended discussion of the particular material proper-
ties, the thermomagnetic switching process and typical disk configurations with a
high signal-to-noise ratio. A summary of the most relevant properties and results in
the field of amorphous alloys is given in section 9.

2. Preparation methods

2.1. Liquid quenching

A common method to produce amorphous alloys is based on the fast quenching of


a liquid. This can be reached with a moving chill block that extracts heat from the
liquid and thus solidifies it. A high thermal conductivity of the chill block is necessary
294 P. HANSEN

to obtain high cooling rates in order to prevent any crystallization process. The heat
transfer from the melt layer to the metal determines the quenching rate and is
important for the maximum thickness of the amorphous layer that can be achieved
by this process. Various techniques (Pond et al. 1974, Chen 1980a) were developed.
The liquid-quenching method cannot be applied to elements that are immiscible in
a liquid as, e.g., Fe and Ag or Fe and Ge. The splat cooling technique is one of the
easiest possibilities to prepare amorphous samples by liquid quenching. In this case,
a drop of liquid is caught against a fixed anvil by a rapidly moving piston. This
leads to disc-shaped samples of typically 1 cm diameter and 20 to 60 gm thickness.
The melt-spinning (Bedell 1975) is a widely used technique to produce amorphous
alloys. It permits a much better control of the preparation conditions as compared
with the splat cooling and provides a large-scale production. The melt-spinning
apparatus is sketched in fig. la. The rf-melted alloy is contained in a quartz or
alumina vessel. The melt is sprayed through a narrow orifice onto the surface of a

'";'}er

di.'

bon

cr

\
(a) rf heating
pressure
•//Aargon
o~rf heating

w h e e l ~

(b)
Fig. 1. Schematic representation of (a) the apparatus for melt-spinning and (b) crucible melt extraction.
MAGNETIC AMORPHOUS ALLOYS 295

spinning wheel that is preferentially made out of copper to provide a high heat
conductivity. Additionally, watercooling of the wheel can be used. The ejection
velocity of the melt is controlled by the pressure of an inert gas like argon. This
method permits the production of thin ribbons of 10-40 ~tm thickness at velocities
of 10-50 m/s. The quality of the ribbon is determined by various parameters (Lieber-
mann and Graham Jr. 1976, Ray 1978, Kavesh 1978, Liebermann 1983). In particular,
the melt jet velocity and the speed of the rotating wheel primarily control the
dimensions of the amorphous ribbon. Alloys such as rare-earth-transition-metal
alloys need process conditions preventing any oxidation. Another modification of
the chill-block method is the crucible melt extraction technique. In this case, the
heat-extracting rotating disc is in contact with the surface of the melt (Maringer and
Mobley 1974, 1978), as sketched in fig. lb. As in the case of melt-spinning, a con-
tinuous ribbon can be obtained. Debris on the disc are removed by a wiper. The
shape of the ribbon depends on the profile of the rotating disc. Their thickness is
less than in the case of melt-spinning due to the substantial low cooling rates caused
by the short time of contact between the liquid and the rotating disc. The roller
quenching is a further technique that uses two counter-rotating rollers (Chert and
Miller 1970, Babi6 et al. 1970, Lewis et al. 1977). In this case, the melt jet is quenched
between two rollers leading to a better contact than in the case of other methods.
However, the contact time is small and, thus, the total heat transfer cannot be
significantly improved. The dimensions of the amorphous ribbons are determined
by the roller pressure or the gap between the rollers. However, this requires an
extremely precise control of the experimental arrangement.

2.2. Vapor quenching

The vapor-quenching method bypasses the liquid state. Thus, many amorphous
alloys that cannot be made by liquid quenching can be well prepared by this method.
Also, a much wider range of alloy concentrations can be realized. The vapor deposi-
tion has to be performed in a protective atmosphere to prevent reactions with oxygen
or nitrogen. The vapor-quenching methods are mostly restricted to the preparation
of amorphous films. The thermal evaporation is a well-established method to produce
multicomponent alloys. In this process, the metal source is heated by resistance,
electron beam or by laser heating. From the melted material, the atoms are released
into the vacuum where the vapor pressure has to be sufficiently high to obtain a
reasonable deposition rate on the substrate. Most alloys do not evaporate congru-
ently. Therefore, it is advantageous to use a multiple source system that permits to
heat each component of the alloy separately and to control the partial vapor pressures
independently. This can be realized by electron-beam multiple source evaporation
(Glang 1970) as sketched in fig. 2 for the case of three sources. An electron beam
generated by an e-gun melts the metal in the crucible. The path of the electron beam
is controlled by means of an electric and magnetic field. It is scanned over the metal
surface to avoid any hole burning. The vapor pressure of each constituent is con-
trolled by a quartz oscillator via a feedback to the e-gun. To avoid lateral inhomo-
geneities of the amorphous film, rotation of the substrate is required. The deposition
296 P. HANSEN

C~

-substrate

-quartz
oscillator

e-beam
e-gun

Fig. 2. Schematic arrangement of a multiple e-gun evaporation system.

rates are of the order of 10-50nm/s. High film quality can be reached for film
thicknesses below 0.3 Ixm. Thicker films tend to vary in composition with thickness
because it is difficult to keep the deposition conditions constant over a longer time
period. The sputtering technique (Glang 1970, Vossen and Kern 1978) is also a well-
known deposition technique where the target is bombarded with ions of a rare gas.
Different sputtering arrangements such as DC, rf and magnetron sputtering have
been investigated. In all cases, a plasma is generated by a potential difference which
also accelerates the rare-gas ions. They transfer their kinetic energy to the surface of
the target releasing atoms of the different constituents which form an amorphous
film on the substrate. A schematic representation of the magnetron sputtering system
is shown in fig. 3. Typically, argon is used as sputter gas and a voltage of some
hundred volts is applied to generate the plasma. A magnetic field induced by a
magnet fixed below the target confines the plasma, leading to higher sputter rates
(Huggins and Guvitch 1983) as compared to DC and rf sputtering where the substrate
serves as anode and, thus, the plasma is spread over the entire volume between the
target and substrate. This anode-cathode configuration for DC and rf sputtering is

vacuum

Fig. 3. Principle of a magnetron sputtering system.


MAGNETIC AMORPHOUS ALLOYS 297

associated with a high electron bombardment of the substrate requiring a substrate


cooling which is not the case for magnetron sputtering. Also, very dense and homo-
geneous films at low argon pressures can be produced by magnetron sputtering at
high deposition rates. Thereby, the product of the sputter gas pressure and the
distance between target and substrate is an essential parameter controlling the film
quality (Thornton 1974, Hoffmann and Thornton 1980, Somekh 1984). These advan-
tages make the magnetron sputtering process very attractive for the production of
thin amorphous films such as rare-earth-transition-metal films for optical recording.
Disadvantages of the sputter techniques are the presence of the sputter gas which is
partly incorporated into the film, very expensive alloy targets and a low variability
concerning compositional variations. The latter can be avoided using a multiple-
target co-sputtering system.

3. Structure of amorphous alloys

3.1. Composition and homogeneity

The interpretation of the electric and magnetic phenomena of amorphous alloys


requires a reliable knowledge about their chemical composition. In particular, ferri-
magnetic alloys exhibit a very sensitive variation of some magnetic properties on
composition. Different nondestructive methods such as X-ray fluorescence or the
outcoming characteristic radiation induced by ion or electron bombardment can be
used. The latter method involves a complex technique and a complicated analysis of
the measured data. However, it was demonstrated that the electron probe microanaly-
sis (EPMA) is a powerful method (Willich and Obertop 1982) that can be applied
universally to thin and thick films of complex composition. This technique also
permits to determine the film thickness. The penetration of oxygen into films depos-
ited by evaporation or sputtering can be well analyzed by secondary-ion mass
spectroscopy (SIMS) which provides essential information concerning the film
quality.
The homogeneity of amorphous alloys depends on the preparation method, the
control of the deposition parameters or structural fluctuations on either atomic or
larger scale. In the latter case, small-angle X-ray scattering can be applied to resolve
small atomic clusters of the order of 1-2 nm (Nold et al. 1980, Osamura et al. 1981),
but also columnar structures of the order of 5-25 nm (Leamy and Dirks 1977) which
have been observed in various rare-earth-transition-metal alloys prepared by evapo-
ration and sputtering (Leamy et al. 1980, Yasugi et al. 1981).
The variation of the deposition parameters during the formation of amorphous
films causes significant changes in composition. Variations in depth can be investi-
gated by SIMS and Auger depth profiles. The depth profile of a B1-xFex amorphous
film is shown in fig. 4 (Moorjani and Coey 1984). Variations of the iron content by
more than 1 at.% occur. In ferrimagnetic alloys, this can be sufficient to cause drastic
changes in the position of the magnetic compensation temperature and, thus, in the
coercive field. Also, the approach to magnetic saturation (Kronmfiller 1981c, Malo-
zemoff 1983, Garoche and Malozemoff 1984, Chudnovsky 1989), dynamic magnetic
298 P. HANSEN

0.53

Bl_xFex

0.52

0.51

0.50 I I i 1 I ~ t i
0.4 0.6 0,8 1.0 1.2 1.4 1.6 1.8
Depth from surface [/zm)

Fig. 4. Variation of the iron content with depth in a sputtered amorphous B-Fe film (Moorjani and
Coey 1984).

processes (magnetic resonance, domain-wall motion) or transport properties can be


strongly affected by structural irregularities.

3.2. Density

Amorphous alloys of composition Mt -xTx with T = Fe, Co or Ni, and M = B or P


can be prepared with high densities. These transition-metal-metalloid alloys have
densities within 2% of the corresponding crystalline counterparts (Luborsky 1980).
The density p decreases with increasing metalloid content, as shown in fig. 5. Calcu-
lated densities based on a dense random-packed structure agree for many composi-
tions within 5% with the measured densities. Some measured densities and calculated
packing fractions q = ~rcrapo are compiled in table 1 where r and Po are the atomic

9.0
8.8q Co(fcc)
8.6
8.4 ~"" " " & 4,,,,,£,,q,, C Ol-x B ×
8.2
E
8.0
--& 7.8
7.6
Fel_ x B x
7.4 °°°°o%

7.2 o
o
o
7.0 I I I
0.10 0.20 0.30

Fig. 5. The mass density versus boron content for amorphous B-Fe and B-Co alloys. The solid circles
represent partially crystalline samples (Hasegawa and Ray 1979).
MAGNETIC AMORPHOUS ALLOYS 299

TABLE 1
Measured densities and calculated packing fraction q=~nr3po for some amorphous transition-
metal-metalloid alloys (Luborsky 1980). r represents either the Goldschmidt radius for transition-metal
atoms or the tetrahedral covalent radius for the metalloid atoms. P0 denotes the average density.
Alloy Composition p Po q Ref.*
x (g/cm3) (at./nm3)
PxNil-~ 0.186 8.00 _ 0.04 90.0 0.678 [1]
0.211 7.93 90.4 0.676
0.228 7.80 89.7 0.667
0.240 7.79 90.2 0.668
0.262 7.73 90.5 0.665
PxCo1_~ 0.150 7.9 86.9 0.677 [2, 3]
0.190 7.97 _ 0.04 89.5 0.668
0.203 7.94 89.8 0.687
0.220 7.89 90.0 0.685
0.236 7.90 90.9 0.688
PxFel -x 0.145 7.252 + 0.04 83.6 0.681 [4]
0.160 7.205 83.7 0.678
0.184 7.128 83.7 0.672
0.216 7.025 83.8 0.665
(FexNil -~)o.soBo.2o 0.25 7.94 98.5 0.697 [5]
0.375 7.83 97.7 0.697
0.50 7.72 96.9 0.697
0.625 7.65 96.6 0.701
0.75 7.53 95.7 0.700
0.875 7.46 95.3 0.703
1.0 7.39 95.0 0.707
(FexCo1_x)o.soBo.2o 0 8.22 104 0.714 [5]
0.125 8.06 99.1 0.709
0.25 7.93 98.1 0.706
0.375 7.84 97.6 0.706
0.50 7.70 96.5 0.702
0.625 7.59 95.7 0.700
0.75 7.52 95.4 0.702
* References:
[1] Cargill III (1970). [4] Logan (1975).
[2] Davis (1976). [5] O'Handley et al. (1976a).
[3] Cargill III and Cochrane (1974).

radius a n d the average density, respectively. The calculations are based o n a twelve-
fold c o o r d i n a t i o n a n d use the G o l d s c h m i d t atomic radii for the metal a t o m s a n d the
tetrahedral covalent radii for the metalloid a t o m s (Cargill III 1975a, Gaskell 1983).
O t h e r alloys prepared by sputtering a n d e v a p o r a t i o n exhibit lower densities due to
the presence of a less dense m i c r o s t r u c t u r e or due to the i n c o r p o r a t i o n of oxygen,
n i t r o g e n or the sputter gas. Then, 5 - 1 5 % lower densities are observed as c o m p a r e d
to the c o r r e s p o n d i n g crystalline alloys. F o r r a r e - e a r t h - t r a n s i t i o n - m e t a l alloys, the
difference in density depends also o n the rare earth as s h o w n in fig. 6 ( F u k a m i c h i et
al. 1987b). Some m e a s u r e d densities a n d calculated p a c k i n g fractions are compiled
300 P. HANSEN

12
Fe2R
10

Y La CePrNd Sm Gd TbDyHoEr Lu
tl i t t f t tt ~ tt It t rt t it i
.0 4.5 5.0 5.5 6.0 6.7 7,0 7.5 8.0 8,5 9.0 9.5 10.0

Density of Rare Earth Hetals (g/cm3)


Fig. 6. Room-temperature density of amorphous and crystalline alloys of composition RFe 2 versus the
density of rare-earth metals (Fukamichi et al. 1987b).

in table 2. The influence of the preparation conditions is demonstrated in fig. 7,


showing the density versus argon pressure for sputtered GdTb-Fe films (Klahn et
al. 1990a). Low argon pressure leads to films with a high density, while those prepared
at high argon pressures or by evaporation show a much lower density. This behavior
can be attributed to the different energy of the atoms forming the film. In the first
case, high-energetic atoms are present giving rise to local rearrangements of the
atoms at the film surface which results in dense films. In the second case, the atoms
are thermalized by the argon gas or by the evaporation process. Their energy is too
low to induce a sufficient atomic mobility at the film surface to produce an energeti-
• cally favorable atomic arrangement. This leads to a less-compact microstructure and
thus to less-dense films. These different conditions for the film growth are determined
by the substrate temperature (Movchan and Demchishin 1969), the mobility of the
adatoms, the initial energy of the sputtered atoms, the sputter gas atoms and the
product of the sputter-gas pressure and the distance between target and substrate
(Thornton 1974, Somekh 1984). Dense films, e.g., can be obtained by magnetron
sputtering using a pressure-distance product of less than 60 Pa mm.

TABLE 2
Measured densities and calculated packing fraction r/= 4nr3po for some rare-earth-transition-metal alloys.
r and Po represent the Goldschmidt atomic radius (twelve-fold coordination) and the average density,
respectively.
Alloy Composition p Po ~/ Reference
x (g/cm 3) (at./nm 3)
Tbl _~Fex 0.67 8.3 55.6 0.75 Rhyne et al. (1974a,b)
Gdl _xCO~ 0.85 8.8 71.1 0.76 Tao et al. (1974)
0.79 8.7 65.5 0.76
0.67 8.5 56.0 0.76
0.54 8.4 48.5 0.76
MAGNETIC AMORPHOUS ALLOYS 301

8.5 -~'-'/Oer y stattine GdTb-Fe

t~
E
o
~8.0

o 1 ctt%Tb N

7.5 ~"Pevaporoted
I I I I
0 0.5 1 1.5 2 2.5
PAr (Pal
Fig. 7. Room-temperature density of amorphous GdTb-Fe alloys prepared by magnetron sputtering
versus argon pressure (Klahn et al. 1990a).

3.3. Microstructure

Amorphous alloys are expected to behave isotropic with respect to their structure
and properties. However, thin films prepared by evaporation and sputtering show a
columnar structure (Dirks and Leamy 1977, Leamy and Dirks 1978, Leamy et al.
1980, Yasugi et al. 1981). This structure was investigated by various techniques such
as microfractography, transmission electron microscopy or small-angle electron and
X-ray scattering. The structure is composed of rod-like columns of 5-25 nm diameter.
They are surrounded by less dense material of 1-25nm thickness, reducing the
average material density. A typical columnar structure of a G d - C o film is shown in
fig. 8a, obtained by microphotography (Leamy and Dirks 1978). The microstructure
of an evaporated GdTb-Fe film observed by transmission electron microscopy
(Klahn et al. 1990a) is presented in fig. 8c.
The column orientation in vapor-deposited films is related to the vapor incidence
(Nieuwenhuizen and Haanstra 1966, Fiedler and Schirmer 1988) and can be explained
in terms of self-shadowing. A computer simulation based on hard spheres (Henderson
et al. 1974) maximizing tetrahedral packing yields an amorphous structure with
columnar structure. This microstructure originates from a low atomic mobility during
film formation, in agreement with results obtained from the occurrance of microstruc-
ture as a function of the preparation conditions (Klahn et al. 1990a, Klahn 1990).
Computer simulation based on hard disks (atoms) and different mobilities of the
adatoms (Brett 1989) leads to columnar structures as shown in figs. 8d and e for low
and high surface mobilities, respectively. The calculations yield a substantial density
increase, a decrease in dislocations and voids inside the columns and an increase in
average coordination number with increasing adatom mobility in agreement with
experimental results for films prepared with decreasing pressure-distance product
(Klahn 1990). Resputtering effects (Gambino and Cuomo 1978, Okamine et al. 1985)
and the incorporation of the sputter gas into the film (Heitmann et al. 1987a,b, Klahn
302 P. HANSEN

Fig. 8. Columnar structure: Transmission electron micrograph of (a) a carbon replica of the fractured
edge of an amorphous Gd0.22Coo.78 film prepared by evaporation and (b) the phase contrast of the
microstructure of a 50nm thick sputtered Gdo.zsCoo.76 film, left: underfocussed, right: overfocussed.
The low-angle region of the diffraction pattern of the film in (b) is given in the inset (Leamy and Dirks
1978). (c).Transmission electron micrograph of an amorphous GdTb-Fe film prepared by evaporation
(Klahn et al. 1990a). (d) and (e) computer simulation of columnar structures for low and high adatom
surface mobility, respectively (Brett 1989).

1990) also gives rise to a strong influence on the microstructure. F u r t h e r , c o n c u r r e n t


ion b o m b a r d m e n t leads to significant m o d i f i c a t i o n s of the m i c r o s t r u c t u r e of v a p o r -
d e p o s i t e d films (M/iller 1987, Rossnagel a n d C u o m o 1989).
MAGNETIC AMORPHOUS ALLOYS 303

3.4. Atomic structure

The atomic structure in single-crystalline materials is defined by the crystal symmetry.


The knowledge of a few atomic positions in the unit cell is sufficient to describe the
crystal. Further, a three-dimensional close-packed structure like the fcc lattice can
be built only from tetrahedra and octahedra while the bcc lattice is made of octahedra
alone. In amorphous alloys, the symmetry is absent and principally each atom
constitutes a structural unit. However, their high density and the magnitude of
magnetic moments and other properties being close to those of crystalline compounds
suggest that the short-range order and thus also local structures are very similar to
those in crystalline materials with high coordination numbers. This is due to the
presence of a short-range order which extends over a range of some neighbor shells
of atoms up to a distance of about 1 nm. There is no long-range order like in
microcrystalline materials. However, there is a continuous transition from amorphous
via 'nearly crystalline' (O'Handley 1987a) involving nanocrystals (Li and Smith 1989)
to the microcrystalline structure depending on composition, preparation process and
the conditions for local atomic ordering. Computer simulation in terms of the dense
random-packed hard-sphere (DRPHS) model using different polyhedra (Bernal 1960)
indeed lead to an amorphous structure that involves bond and topological disorder.
Both play probably an essential role in the formation of an amorphous structure
and are illustrated in fig. 9 for a binary alloy by a two-dimensional network (Moorjani
and Coey 1984). The bond disorder (fig. 9b,e) is caused by a nonsymmetric arrange-
ment of the atoms while topological disorder (fig. 9c, f) involves a different number
of bonds and/or a different number of ions on a closed path. In addition, amorphous
alloys exhibit chemical disorder which also can be present in crystalline alloys as,
e.g., in the fcc PtCo, where the Pt and Co atoms are distributed at random as
indicated in fig. 8d. Various further structure models were built (Finney 1970, Bennett
1972, Henderson et al. 1974, Ichikawa 1975), but still result in a packing density of
less than 0.66 which is significantly lower than the measured data. These structures
involve a predominance of tetrahedral interstices. A more satisfactory solution was
obtained with a soft-sphere model using the Lennard-Jones type potentials and a

bond order bond disorder topologica[ order

chemical
order

ct b

chemicctt
disorder

d e f
Fig. 9. Types of disorder on two-dimensional binary lattices (Moorjani and Coey 1984).
304 P. HANSEN

relaxed DRPHS structure (Weaire et al. 1971, Barker et al. 1975, von Heimendahl
1975). This model reaches a packing fraction of 0.74 which is almost equal to that
of an fcc lattice and is due to a much lower void content.
Experimentally, the amorphous structure can be studied by means of X-ray,
electron and neutron diffraction as well as by extended X-ray absorption fine structure
(EXAFS) measurements (Cargill III 1975b). An X-ray diffraction diagram for an
Zro.66Coo.34 alloy before and after crystallization is presented in fig. 10. The amor-
phous structures do not show any sharp diffraction peaks due to the absence of long-
range periodicity. The crystallized sample shows the typical X-ray pattern of the
CuA12 structure (Buschow 1984a).
As any symmetry is absent in amorphous alloys, only statistical information and
averaged data can be determined. The average distribution of atoms can be described
by the radial-distribution function (RDF) 4nr2p(r) and represents the number of
atoms,

N = 4~
f: p(r)r 2 dr,
1
(1)

between rl and r 2. It is convenient to use the reduced radial-distribution function,


G(r) = 4nr[p(r) - P0 ], (2)
because it is directly related to the Fourier transformation of the interference function
S(q) which represents the diffraction intensity divided by the square of the scattering
factor. P0 denotes the average density and q is the defraction vector. For monoatomic
solids, S(q) is related to G(r) by

G(r) = ½n fooq[S(q) - 1] sin(qr)dq. (3)

o
C O4 ~O'~
I C~ I
[ I I |j[ Q4 |

. . . . . . . . . .

4.2 40 38 36 3/* 32 30 28 26 2/* 22 20 18 16 1/*


(3 (deg)

Fig. 10. X-ray diffraction pattern of melt-spun Zr0.66Co0.34before crystallization (upper curve) and after
crystallization (bottom curve). The indexing of the bottom curve corresponds to the tetragonal CuA12
structure (Busctlow 1984a).
MAGNETICAMORPHOUSALLOYS 305

For polyatomic solids, p(r) is composed of the partial distribution functions (Faber
and Ziman 1965, Wagner 1972, 1978, Cargill III 1975a, Waseda 1980). The Fourier
transform for a binary alloy of composition A1 -xBx takes the form

. fB(q) o , ,
SA(q) = (1 -- X)2(f(q))fA(q) SAA(q)-t- X2(f(q)) ~AB~q)" (4)

SAA and SAB are the compositionally resolved structure factors for A-A and A-B
pairs, respectively, fg(q) is the scattering factor of atom A and fB(q) of atom B. ( f ( q ) )
denotes to compositionally averaged factor. SB(q) can be expressed accordingly. The
total radial distribution function thus is composed of the averaged partial distribution
functions of the different types of atoms. In the case of binary alloys, at least three
independent sets of data are necessary to determine the three partial distribution
functions. This can be achieved using different diffraction methods (X-ray, electron,
neutron diffraction), different isotopes with different neutron-scattering factors (Mizo-
guchi et al. 1978, Kudo et al. 1978) or by the substitution of elements of similar size
and chemical affinity but different scattering power (Chipman et al. 1978, Williams
1982). Also, different kinds of radiation were applied to determine the partial distribu-
tion functions (Waseda and Tamaki 1976).
A comparison of experimentally determined and calculated reduced radial distribu-
tion functions was made for various transition-metal-metalloid alloys (Cargill III
1975b). For many alloys, a good agreement between experimental and calculated
data was achieved. This is demonstrated in fig. 11a for amorphous
Po.14Bo.o6Feo.4oNio.4o. The curve represents G(r) deduced from dispersive X-ray
diffraction and the histogram is the reduced G(r) of amorphous Fe calculated from
the soft-sphere model (Egami 1978a). The good agreement of the G(r) of amorphous
iron with the experimental data in respect to both the entire radial dependence and
the individual peak heights reveals the dominance of the transition-metal ion or the
much lower X-ray scattering power of the metalloid. Therefore, pronounced changes
in G(r) only occur for alloys with metalloid atoms differing significantly in their
atomic radius. This is demonstrated in fig. llb for amorphous Bo.25Feo.v5 and
SioA5Bo.loFeo.75, the latter representing the case that a larger metalloid is involved
(Aur et al. 1982). The second peak is shifted towards larger distance by 0.025nm,
reflecting the difference in the atomic radii of about 0.02 nm between Si (0.11 nm)
and B (0.091 nm). Further, the study of the G(r) reveals a strong short-range order
which means that there is a strong preference of the metalloid atoms to have
dissimilar atoms as nearest neighbors. Also, the apparent radius of the metalloid
atom in the alloys is significantly smaller than the radius of the free atom.
The rare-earth (R)-transition-metal (T) alloys play a special role within the amor-
phous materials with respect to the substantial difference in their atomic radii and
scattering power. The structure can be well accounted for by the binary dense
random-packing model (Cochrane et al. 1974, 1978a,b, Cargill III and Kirkpatrick
1976, Williams 1981, Cargill III 1975a,b, 1983) achieving packing densities of 0.76
(see table 2). The difference in their atomic radii leads to a split of the first peak of
G(r) into subpeaks according to the three contributions from the partial distribution
306 P. HANSEN

6 Po.lz.B o.o6F eo.z.o N io.~o

/,

2 't

0 0.5 1.0 1.5


r(nm)

..... Si0.15Bo.lo Feo.7s


/..0 . . . . . . Bo.25Feo.75

2.0
"C
(D

0.0 ,,v,y

-2.0
(b)
F I ] I I I I ]
0,16 0.32 0,/.8 0.6/. 0.80
r(nrn)
Fig. 11. Reduced radial-distribution functions for amorphous PB-FeNi determined by the energy disper-
sive X-ray diffraction. The histogram represents G(r) (a) for amorphous Fe calculated from the soft-sphere
model (Egami 1978a) and (b) for amorphous B Fe and BSi-Fe ribbons. The shift of the second peak
reflects the difference in the radius between Si (0.11 nm) and B (0.091 nm) (Aur et al. 1982).

functions PRR(r), PRT(r) and PTT(r). This also applies to Lal_xFex alloys where the
three correlations for Fe-Fe, Fe-La and La-La are well distinguishable (Matsuura
et al. 1988), as shown in fig. 12. The arrows indicate the positions of the peaks
correlated to the different interatomic distances. The extracted nearest-neighbor
distances and coordination numbers are listed in table 3. The split of the first peak
of G(r) for amorphous Gdo.36 Feo.64 is shown in fig. 13. The broken lines represent
Gaussian fits of the three possible neighbor-pair peaks (Cargill III 1975a,b). Therefore,
in this case, one diffraction experiment is sufficient to deduce the nearest neighbor
distances rij and coordination numbers N u based on eqs. (1)-(4). The radial depen-
dence of G(r) for amorphous Tbo.33Feo.67 is shown in fig. 14. Again, the first peak
is split and the maxima of the partial distribution functions are indicated by the
arrows. The structure of the second peak originates from the indicated atomic clusters
MAGNETIC AMORPHOUS ALLOYS 307

/ Lo.1-xFe×
[ Im~'~ OFe

"L" =~_

0 . . . . .

xj
-2 I I i t
0.2 0,/~ 0.6 0,8 1.0 1.2
r(nm}
Fig. 12. Reduced radial distribution function for amorphous La-Fe alloys (Matsuura et al. 1988). The
arrows indicate the positions of the interatomic distances corresponding to the sketched types of atomic
arrangements.

(Rhyne 1974). Average distances of nearest-neighbor atoms and coordination


numbers are compiled in table 3.
The compositional short-range order (CSRO) is less pronounced in the rare-earth-
transition-metal alloys as compared to the transition-metal-metalloid alloys. The
CSRO was discussed in terms of the heat of mixing (Buschow and van Engen 1981a,
Buschow and van der Kraan 1981). A negative heat of mixing implies the attraction
of dissimilar atoms and a positive value the opposite behavior. Therefore, the heat
of mixing can be used as an additional parameter to improve the dense random
packing model.

4. Stability of amorphous alloys

4.1. Crystallization temperatures

Amorphous alloys are in a metastable state and tend to transform into stable
crystalline phases. At temperatures below the crystallization temperature, structural
relaxation effects take place and are caused by atomic rearrangements. The crystalli-
zation is associated with nucleation and growth processes (Avrami 1939, 1940, 1941).
The transition to the crystalline state is accompanied by an exothermic heat effect
giving rise to a sharp peak in the temperature dependence of the exothermic heat.
Therefore, differential scanning calorimetry is a widely used technique to study
thermally induced transformations in amorphous alloys and to determine the crystal-
lization temperature, T~. The magnitude of T~ is very different for amorphous
materials and depends strongly on composition. The activation energy ranges typi-
cally between 2 and 6 eV. Amorphous metals (Fe, Co, Ni) with a very low impurity
content exhibit crystallization temperatures below 70 K (Felsch 1969, 1970a,b, Wright
308 P. H A N S E N

TABLE 3
Average distances (ru) of nearest-neighbor atoms and coordination numbers (Nu) for amorphous and
crystalline rare-earth-transition-metal alloys.

Partials Amorphous Crystalline Ref. t


rq(nm) Nu ru(nm) Nu

Gdo.36 Feo.64 GdFez [i]


Gd-Gd 0,347 ±_ 0.005 6.0 ± 1 0.320 4
Gd-Fe 0.304 4- 0,005 6.5 4- 0.6 0.306 12
Fe-Fe 0,254 _ 0.005 6.5 + 0.5 0.261 6
Gdo.18Coo.82 GdCos [1]
Gd-Gd 0.34 3 4- 1
Gd-Co 0,297 ___0.0005 12 + 1 0.287 6
0.318 12
Co-Co 0,247 ±_ 0.005 7.2 ± 0.7 0.245 4,8
0.249 2.4
Tbo.33 Feo.67 TbFe2
Tb-Tb 0.35 6.4 0.318 - [1,2]
Fe-Tb 0.30 - 0,304 -
Fe-Fe 0.26 6.3 0.254 -
Hoo.3oCoo.7o HoC%
Ho-Ho 0.352 - - 4 [3]
Ho-Co 0.298 - - 12 [4]
Co-Ho - - 6
Co-Co 0.251 - - 6
Lao.35 Feo.6s LaFe2 *
La-La 0.360 5.3 - - [5]
La-Fe 0.315 3.3 - -
Fe-Fe 0.252 6.1 - -
* The intermetallic c o m p o u n d has not yet been synthesized.
t References:
[1] Cargill III (1975a,b). [4] Cochrane et al. (1978a).
[2] Rhyne et al. (1974a). [5] M a t s u u r a et al. (1988).
[3] Nagy ct al. (1977).

Gd°'36F%'6~~
I I I

10
-E
£D

./K/ i\ \
"\ ',
0.2 0.3
r(nrn) 0.4
Fig. 13. The first peak in the reduced radial distribution function for a m o r p h o u s G d - F e . The broken lines
represent the three possible neighbor peaks (Cargill III 1975a,b).
MAGNETIC AMORPHOUS ALLOYS 309

2~

v I/!'!"..
oo,
DO0
L,/ o°o •

orb

0 0.2 0.4 0.6 0.8 1.0


r(nm)

Fig. 14. The reduced radial distribution function for amorphous T h - F e at T = 433 K . The arrows indicate
the contributions of respective atomic clusters (Rhyne 1974).

1976), which rapidly increase if the impurity content is raised. Thus, alloys of
composition M1-xT~ with T = Fe, Co or Ni need a very low M concentration of
typically x < 0.15 to establish the amorphous state with T~ above room temperature
(Fujimori et al. 1984, Fukamichi and Hiroyoshi 1985). For example, the crystallization
temperatures for amorphous Zrl_~Fex with 0.88 ~<x ~<0.93 range between 882 and
806K (Ryan et al. 1987b). The compositional dependence of T~ for amorphous
transition-metal-metalloid alloys (Naka et al. 1976, Luborsky and Walter 1977a) is
shown in fig. 15. Although these alloys exhibit a pronounced dependence of T~ on

i I
750 P°'13C°°'°7 Fe°'8°-x M x
i'4=Cr
0 Feo.80-xNix

700

650 ~ P°'lZ'B°'°6 F e°'e°-x Ni '

..,x
60c X/
PO.lZ,BO.O6Feo.so-xNix
55 i I I
0 0,20 0.40 0.60 0.80
x
Fig. 15. Crystallization temperatures for amorphous P-CoFeM alloys (open symbols) measured with a
heating rate of 5 K/min (Naka et al. 1976) and BP-FeNi, B-FeNi (solid symbols) measured with a heating
rate of 40 K/rain (full line) and two hours anneals (broken line) (Luborsky and Walter 1977a).
310 P. HANSEN

composition, the overall T~ values remain high. This applies to most of the alloys
containing glass formers, leading to a high stability of these alloys in the room-
temperature range, which is of importance for technical applications. At the M-rich
side of M - T alloys, T~ again decreases as shown for amorphous Snl _~T~ alloys
(Geny et al. 1982).
The T~ values for many rare-earth-transition-metal alloys range between 400 and
800K as obvious from table 4. The strong dependence of T~ on composition is
demonstrated in fig. 16a for R-T alloys with R = Gd or Tb and M = Fe or Co
(Klahn 1990) and in fig. 16b for amorphous N d - F e (Suzuki 1985) and Tb-Co
(Buschow 1980a,c). T~ increases with increasing atomic number of the rare-earth-
element (Buschow 1980b, Buschow and Beekmans 1980). In the case of Tb-Co, T~
interferes with the Curie temperature which restricts the investigation of the magnetic
properties in the amorphous state. This limitation occurs for all rare-earth-cobalt

TABLE 4
Crystallization temperatures for amorphous rare-earth-transition-metal (R-T) alloys. Preparation
methods used are liquid-quenching (lq), and vapour-quenching (vq). Further T~ data are compiled in
previous reviews (Luborsky 1980, Buschow 1984a, Hansen 1988a).
R1 -xT~ x T~ (K) Preparation method Ref.*
Smi _xFex 0.40 618 lq, vq [1, 2]
Gd 1 _xFe x 0.40 410 lq
Tbl _xF% 0.40 590 lq
Dyl - x F% 0.31 603 lq [-3]
Hol_xFe x 0.31 608 lq [1,2]
Erl_~F% 0.25 433 lq
0.31 631 lq
0.40 628 lq
Tm 1_xF% 0.31 648 lq [4, 5]
Ndl _~Cox 0.31 460 lq, vq
Sml_~Co~ 0.31 516 lq
Gd l_~Co~ 0.31 550 lq
Tb 1 - ~Cox 0.25 528 lq
0.31 570 lq
0.40 573 lq
0.60 528 lq
Hol _~Co~ 0.31 408 lq
Er 1 _xCo x 0.31 633 lq
Pr~ _~Nix 0.31 449 lq [6]
Nd 1 - xNix 0.31 489 lq
Gd 1 _~Ni~ 0.31 553 vq
0.40 538 lq
Tbl -~Nix 0.31 587 lq
Dy i _xNi~ 0.31 591 lq
Er i - x Ni~ 0.31 620 lq
* References:
[1] Buschow (1981a). [4] Buschow and Beekmans (1980).
[2] Buschow and van der Kraan (1981). [5] Buschow (1980a).
[3] Buschow (1984a). 1-6] Buschow (1980b).
MAGNETIC AMORPHOUS ALLOYS 311

1000

800 R 1-x Fex

600

400

200

(o)
I I I I I I I I I
0.50 1.00
X

800
oe Ndl_x Fex
700 , , ~ om Tb 1 x Cox

600 ----2(\

500

400

3OO
200 [b)l ITc~ i i I "~'4
0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
X D,

Fig. 16. (a) Crystallization temperature versus composition for amorphous R-T alloys (Ktahn 1990):
(©) Buschow (1980a, 1985), (A) Campbell (1972), (E]) Felsch (1970b), (O) Felsch (1969), (A) Buschow
(1985). (b) Crystallization temperature and Curie temperature versus composition for amorphous Nd-Fe
(Suzuki 1985) andTb-Co (Buschow 1980a,c, Hansen et al. 1989) alloys.

alloys and is even more pronounced for amorphous Fe, Co and Ni with their very
low T~.
The dependence of T~ on the heating rate s = dT/dt can be used to determine the
activation energy of crystallization (Luborsky and Liebermann 1978). Considering
the fraction x of amorphous material transformed into the crystalline state in time
t and at temperature T, one obtains for the first-order rate process (Kissinger 1957,
Boswell 1980),

ex) =/,(1 - x). (5)


312 P. HANSEN

For a thermally activated process, the rate constant k obeys an Arrhenius type of
equation

where ko is a constant and AE is the activation energy. Combining eqs. (5) and (6)
and using dx = (~x/&)r dt + (~?x/~T)t d r with (Ox/Or) dt ~ O, one obtains

dXdt k o ( 1 - x ) e x p - ~ - ~ . (7)

At the peak of the exothermic heat, the change of the reaction rate dZx/dt 2 is equal
to zero, yielding with T = T~,

ko exp - = s. (8)

From the measured data of s and T~, the activation energy can be deduced from
the slope of a plot of ln(T)/s) versus TZ 1. This is shown in fig. 17a for amorphous
Yo.33Feo.67 and Yo.2sFe0.75, leading respectively to activation energies of 2.5 and
3.2 eV (Croat 1982) and in fig. 17b for an amorphous GdTb-Fe film with an activation
energy of 2.6 eV (Klahn et al. 1987). A similar method based on eq. (5) is the isothermal
annealing, leading to almost the same AE values.
The stability of amorphous materials was studied in terms of the DRPHS models
(Polk 1972). The significance of the chemical bonds was investigated (Chen 1980b)
and a stability criterion based on a nearly free electron approach was proposed
(Nagel and Tauc 1975). However, problems arise for these models concerning the
interpretation of the various alloy systems. The kinetic approach to thermal stability
based on a diffusion-controlled rate of transformation (Uhlmann 1972, Davies 1976,
Takayama 1976) leads to the relation
T~ = eE, (9)
where e is a constant and E is a measure of the potential-energy barrier for coopera-
tive atomic transitions. Assuming E to be proportional to the formation enthalpy,
AHh, of a hole corresponding to the size of the smaller type of atom in a binary
alloy A l_xBx (Buschow and Beekmans 1979a, Buschow 1981b), a relation of the
form
T~ = flAHh (10)
was proposed where fl is again a constant. Unfortunately, the values of AHh cannot
be determined experimentally. However, these hole enthalpies can be calculated using
a semiempirical approach to describe energy effects in metals (Miedema et al. 1980)
by expressing AHh in terms of the formation enthalpies AHA and AHR of a monova-
cancy in a pure A and B metal by the relation
AH~ = cAN B + (1 - - c)(VB/VA)5/6~HA, (11)
MAGNETIC AMORPHOUS ALLOYS 313

/*'01 yo.25Feo,75
/ A E=3.2eV

~ 2.0 3F %.67
1.0[ / / AE=2,5eV

1,0 1.1 1.2 1.3


Txl(10-3K-1)

-13

E" -14
.ic
"7 ~ 7 0 0 800 900
' N ~ T(K)
¢_" -15
-, ~ & E=2.6eV
Gdo.'r75Tbo.ossF e°.78 ~ , ~
-16

-17 (b) I ~ I I I I I I
1.00 1.04 1.08 1.12 1.16
T~{10-3Kq)
Fig.17.Plot of log (T~/s) v e r s u s Tx- 1 for (a) amorphous Y - F e alloys (Croat 1982), and (b) amorphous
GdTb-Fe(Klahn et al. 1987) alloys. Tx and s represent the crystallization temperature and the heating
rate, respectively.

where c is the effective concentration,


xV2/3
c = xV2/3 + (1 -- x ) V2/3" (12)
VA and lib represent the molar volumes. Using the values listed for most metallic
elements (Miedema 1979), Tx can be calculated from eqs. (10) to (12). The result is
presented in fig. 18 and shows an excellent agreement between eq. (10) and the
measured T~ values for a large number of alloys (Buschow 1982a).

4.2. Structural relaxation

The structure of amorphous alloys is temperature dependent due to its metastable


state. Therefore, temperature treatments tend to change the atomic arrangement.
314 P. HANSEN

1500 _n Sn~_x Co x
o Sn1_ x F e x
• Sn1_ x Nix .,,Te%,sslro.z,s
• Til_x COx /.., .
0 U- T ~/~Do ssir0 zs
1000 -4' Tl~"lX_x~e x .,¢.,~TCto.ssRh o.ls
* Wl x Fex d - + ' N~o__.Rh_ ._
+T.o.:_x Tx ' ~ * < > ~;os7 Vo~ ....
:_x *Tl,.xNlx , ~ . / ~ . •
500 ,~oo~ -

//
0 Pi I I I I I I I I I I I I I I i I I I ] I

100 200
AHh (kJ/rnol)

Fig. 18. Experimental crystallization temperatures versus calculated hole formation enthalpies (Buschow
1982a,b, 1984a). T represents transition metals.

Annealing at temperatures below the glass or crystallization temperature is thus


associated with a change of various properties which depends on that particular
annealing time and temperature (Egami 1978b, 1981, 1983a,b, 1984). Reversible
relaxation phenomena are associated with different effects and properties such as the
field-induced anisotropy, isochronal changes in the Curie temperature, electrical
resistivity, elastic constant, etc. The reversible relaxation effects occur even at very
low temperatures, while the irreversible processes take place at higher temperatures
as compared to the glass or crystallization temperature. Generally, three typical
effects are observed: (i) reversibility, (ii) 'crossover' effects involving a crossover of
the changes of a property versus annealing time, ta, for two different annealing
temperatures and (iii)the presence of 'ln t,' kinetics. First approaches to account for
these effects utilize an activation energy being linearly related to the instantaneous
magnitude of the measured property (Egami 1978a) or an discrete spectrum of
activation energies (Woldt and Neuh/iuser 1980). However, difficulties arise using
these models to explain all phenomena simultaneously. This has led to a new
approach based on a continuous spectrum of relaxation processes (Gibbs and Evetts
1982, Gibbs et al. 1983, Fish 1985). Then, the total measured change of a property
can be expressed by (Gibbs et al. 1983, Fish 1985)
('
AP(t) = J~ c(E)q(E, T~, ta) dE, (13a)

where q(E, T., &) in the case of first-order kinetics is of the form

q(E, T~,ta)=qo(E, Ta) {1 - e x p [ - Vtaexp(-- ~T~)]}, (13b)

provided each process is thermally activated, c(E) is the measured property change
if only one such process having an activation energy E is thermally activated per
unit volume of the material, q(E, T~, ta) is the number density of processes of the
MAGNETIC AMORPHOUSALLOYS 315

activation energy E which have contributed to the relaxation after annealing time
ta. c(E)q(E, Ta, ta) thus is the property change related to relaxation processes having
activation energies in the range E to E + dE. The parameter v appearing in eq. (13b)
is the frequency factor. In the case of one isothermal annealing experiment and where
q(E, T,, ta) is approximated by a step function (Primak 1955) with q(E, T,, ta) = 0 for
E < Eo, the property change is given by
AP = co(E)qo(Eo, T~)kTa ln(vta), (14a)
where
Eo = kT, ln(vta) (14b)
representing the typical 'ln ta' kinetics.
Irreversible changes are associated with a reduction in volume (Chen 1978a,b)
where the kinetics of the variation in linear dimensions were investigated (Kursumovic
et al. 1980a,b). These changes in volume involve changes in the radial distribution
function and, thus, can be studies by diffraction methods. Also, diffusivity and
viscosity changes with annealing temperature (Anderson and Lord 1980, Taub and
Schaepen 1980). Any structural relaxation affects the local electron density and
thus gives rise to changes in the magnetic properties. Their variation with thermal
annealing will be discussed in section 5.10.

5. Magnetic properties

5.1. Magnetic structures

The magnetic properties and rules in crystalline materials are governed by the crystal
symmetry. The possible magnetic structures are defined by the magnetic space groups.
In the case of collinear arrangements of magnetic moments, this leads to the familiar
forms of magnetic order like ferromagnetism, antiferromagnetism and ferrimagnetism,
and, in the case of noncollinear structures, to helimagnetism or canted spin structures.
The presence of chemical and structural (bond and topological) disorder in amor-
phous materials produces an inequivalency of sites that leads to a distribution (i) in
magnitude of magnetic moments, (ii) in exchange interactions and (iii) induces large
randomly varying electrostatic fields giving rise to locally varying single-site anisot-
ropy. Although magnetic order depends sensitively on distances and local environ-
ments, the structural disorder does not prevent the existence of collective magnetic
order. When the local anisotropies are negligible as compared to the exchange
interactions, the possible collinear arrangements of magnetic moments in amorphous
materials are sketched in fig. 19 for the common types of magnetic order (Coey 1978).
It has been assumed that each atom carries a time-independent moment but any
local fluctuation of its magnitude is not represented by the length of the arrows. The
ferromagnetic structure shown in fig. 19a is observed for many materials as, e.g.,
Gdl-xM~ alloys with M = Ag, Au, A1 or Cu and transition-metal-metalloid alloys
like Tl-xMx with T = Fe, Co or Ni and M - - B , P or C. The existence of an
amorphous antiferromagnet is still questionable because a consistent subdivision of
316 P. HANSEN

(a) ferromagnet (b) a n t i f e r r o ma g n e t (c) ferrimagnet

magnetization I~1 , 0 ~:0 ~,o


~ = -~= ~ > -~
Fig. 19. Collinear magnetic structures in amorphous materials (Coey 1978).

a random structure into two sublattices is impossible (Simpson 1974). In crystalline


materials, antiferromagnetism is defined by symmetry and requires equivalent atoms
on equivalent sites. Both requirements cannot be fulfilled in an amorphous structure.
The collinear ferrimagnetic structure shown in fig. 19c is observed for Gdl_xRx
alloys with T = Fe, Co or Ni and thus represents a two-subnetwork structure contain-
ing two distinguishable groups of atoms. In this sense, e.g., FeNi or FeCo will not
be discussed under the topic of the two-subnetwork alloys.
In addition to these familiar types of magnetic order, noncollinear structures with
0 < S~Sj < IS~IISjI like speromagnetic, asperomagnetic and sperimagnetic order occur
in amorphous alloys as illustrated in fig. 20. These structures involve a competing
random anisotropy and exchange interaction where the local anisotropy aligns the
magnetic moments along the locally varying crystalline field axis. All these structures
represent arrangements of magnetic moments which are frozen at a sufficiently low
temperature. The speromagnet can be considered as a random antiferromagnet and
the asperomagnet as a random ferromagnet. Both represent one-subnetwork mag-

(a) a s p e r o m a g n e t (b) s p e r o m a g n e t (c) sperimagnet

magnetization ~1 ¢ 0 ~1 = 0 ~1 , 0
Fig. 20. Noncollinear magnetic structures in amorphous materials (Coey 1978).
MAGNETIC AMORPHOUS ALLOYS 317

netic structures. The speromagnetism was observed for Yl-xFex and Fo.75Feo.25
while anisotropy-dominated asperomagnetism was found for D y o . z l N i o . 7 9 . Other
spin structures, where the ferromagnetic alignment is limited to domains of some
nanometers, are discussed for amorphous Zr-Fe alloys (Rhyne and Fish 1985, Ryan
et al. 1987b).
Sperimagnetism occurs in two-subnetwork structures like rare-earth-transition-
metal alloys of composition Rt _xT~ where the large spin-orbit coupling of the non-
S-state rare earths give rise to large local anisotropies. Two classes of alloys can be
distinguished containing heavy and light rare earths. In the first case, the R and T
moments are coupled antiparaUel and in the second case parallel according to Hund's
rule and the exchange coupling between the spins of the R and T atoms involving
negative coupling of the 5d rare-earth electrons and the 3d transition-metal electrons
(Campbell 1972). The first case, representing the antiparallel alignment of the mo-
ments, is illustrated in fig. 20c and fig. 21a for a collinear T sublattice which is
representative for Co-based alloys due to the strong ferromagnetic coupling of the
Co sublattice. For light rare earths with a parallel alignment of R and T moments,
the situation is shown in fig. 2lb. In Fe-based alloys also a distribution of the Fe
moments can occur as shown in fig. 21c,d. The strength of the R - T exchange coupling
depends on the R element (Beloritzky et al. 1987) and is related to the variation of
the 4f-5d interaction which is larger for light rare earths because the spatial extent
of the 4f and 5d electrons is reduced.
The distribution of moments induced by local anisotropies is not uniquely associ-
ated with disordered structures. Rare-earth-iron garnets also show a distribution of
rare-earth moments due to the high single-ion anisotropy at low temperatures (Clark
and Callen 1968, Englich et al. 1985).
Another possibility is the presence of an exchange interaction of both signs which

Sperimognetic structures

a) Dy-Co b) Nd-Co

._---~ - __~

c) D y - F e d) Nd-Fe

r a r e e a r t h moment •
t r a n s i t i o n meta[ moment . . . . . -~

Fig. 21. Sperimagnetic structures in amorphous rare-earth-transition-metal alloys containing heavy rare
earths (Dy, Tb, etc.) and light rare earths (Nd, Pr, etc.) (Taylor et al. 1978). The configurations sketched
in (a) and (b) represent an asperomagnetic R suhlattice and a collinear T sublattice, while in the
configurations (c) and (d) both sublattices exhibit a noncollinear structure.
318 P. HANSEN

is associated with the spin-glass behavior. These materials possess a typical spin-
freezing temperature defined experimentally by a sharp peak in the low-field suscep-
tibility as shown in fig. 22a for amorphous Sbo.s0Feo.5o (Xiao and Chien 1985). The
magnetic phase diagram of Sb-Fe alloys shown in fig. 22b indicates the ranges of
composition and temperature where spin-glass behavior is observed. However, also
speromagnets, exhibiting a random distribution of moments due to the random local
anisotropy, show a spin-freezing temperature associated with the typical cusp in the
low-field susceptibility and thus might be considered as a spin glass. A large variety
of compositions of either crystalline and amorphous materials show spin-glass beha-
vior (Moorjani and Coey 1984). There is a principal difference between amorphous
Fe-based and Co-based alloys with respect to the magnetic structure. The former
tend to spin-glass behavior or speromagnetic structures (Xiao and Chien 1987, Ryan
et al. 1987a,b, 1988, Fukamichi et al. 1988, Wakabayashi et al. 1987, 1990, Kakehashi
1990a,b) while the latter in general exhibit good ferromagnetic order. This is associ-

10

>, 5

(o) i i
20 ~0 60
T(K}
300

Sb1_ x F e x

200 para
/
o

F- 100 fe

spin g t a ~ s / Ie

o I ,l
0,30 0.40 0.50 0,60 0.70
x

Fig. 22. (a) Low-field susceptibility (2.4 kA/m) versus temperature exhibiting the typical cusp at the spin-
freezing temperature Tf for amorphous Sbo.soFeo.so and (b) the magnetic phase diagram for amorphous
Sb-Fe alloys (Xiao and Chien 1985).
MAGNETICAMORPHOUSALLOYS 319

ated with the different bonding character of these alloys as it is shown by X-ray
photoelectron spectroscopy (XPS) valence-band spectra (Amamou 1980, Amamou
and Krill 1979, 1980) and the much higher exchange interaction of the Co alloys.

5.2. Influence of preparation conditions

The structure of amorphous alloys represents no ideal statistical arrangement of


atoms but depends on the particular preparation conditions leading to differences
in the local atomic environments and nearest-neighbor distances, the occurrence of
a typical microstructure or the incorporation of gas atoms. This gives rise to varia-
tions in the magnetic properties. The magnetic moment is less sensitive to structural
differences while the Curie temperature, the coercivity, the uniaxial anisotropy
and the compensation temperature occurring in ferrimagnets are strongly affected
by structural changes and thus depend on the preparation parameters. The com-
positional variation of the specific magnetization and the Curie temperature for
evaporated (Miyazaki et al. 1987a) and liquid quenched (Miyazaki et al. 1986)
amorphous Sml-xFex alloys are shown in fig. 23. The magnetization data are
independent of the preparation technique and agree with those of crystalline com-
pounds indicated by the squares. The corresponding Tc variation reveals a completely
different behavior for the two amorphous alloys, probably due to the higher
amorphicity of the evaporated films. In the case of amorphous films prepared by
sputtering, the deposition parameters are of significant importance (Niihara et al.
1985, Hashimoto et al. 1987, Heitmann et al. 1987a,b) as demonstrated in fig. 24
for the uniaxial anisotropy constant, Ku. The influence of the sputter atmosphere
on Ku for magnetron sputtered Gdo.24Tbo.o4Feo.72 films (Heitmann et al. 1987a)
is shown in fig. 24a, while the influence of the argon gas pressure on K~ for diode
sputtered [-(Gd, Tb)l _xCox] 1_yAry films (Heitmann et al. 1987b) is shown in fig. 24b
for different bias voltages. In the first case, an increasing oxidation of the rare earths
causes a reduction of K~ and in the second case a change of the microstructure and
resputtering effects at the growing film surface lead to the observed Ku variations.
In these cases, also the compensation temperature, the Curie temperature and the
coercivity are strongly affected. A very important parameter controlling the structural
disorder is the effective cooling rate that can be influenced via the substrate temper-
ature of vapor-quenched films. This is demonstrated in fig. 25, showing the variation
of the coercive field, He, versus the substrate temperature, T~, of sputtered Sm~_ ~Cox
films (Munakata et al. 1984). Films prepared at T~~<500K exhibit an amorphous
structure and reveal low coercivities. The films prepared at higher substrate temper-
atures are characterized by an increased amount of crystalline particles causing an
increase of He. For these substrate temperatures, the effective cooling rate is too low
to maintain good amorphicity.
These few examples taken out of a great number of investigations may demonstrate
the relevance of deposition conditions on the magnitude and sign of magnetic
properties. Therefore, a meaningful comparison of amorphous alloys with the same
composition concerning their structurally sensitive properties can only be made when
320 P. HANSEN

200
Srnl. x Fe x

160

120
E °°
<
b
80 oQ

40

0
0 0.20 0./,0 0.60 0.80 1.00
x

80O Sin1. x Fe×

700 Sm Fe 2.
Sm F e a ~

600 Sm2 F e T ~
Sm F e s ~ X ' ~

v 50O

I.-
400

300 .... 2 ' "17

200

(b)
100 I I I I
0.20 0.40 0.80 0.80 1.00
x

Fig. 23. Concentration dependence of (a) the specific magnetization at H = 1.2 x 106A/m and T = 0 K
and (b) the Curie temperature for amorphous Sm-Fe alloys prepared (0) by evaporation (Miyazaki et
al. 1987a)and (©) by liquid quenching (Miyazaki et al. 1986).The squares represent crystalline compounds
(Miyazaki et al. 1987b).

the preparation parameters and the knowledge of their influence on the structure
are available.

5.3. Low-temperature magnetic moments

5.3.1. One-subnetwork alloys


5.3.1.1. Transition-metal-based alloys.
Models of moment variation. The average
magnetization is based on the exchange coupled magnetic m o m e n t s which are
M A G N E T I C A M O R P H O U S ALLOYS 321

15
%. (Gd 0.2t,T bo.oz,F e0.72) l_y Ay
atmosphere
~. a ArlN 2
"~ ~_ • Ar/O 2
~E
o 5

-5
0 0.02 0.0/, 0.06 0.08
YN'Yo (at %)

{Gd,Tb)l_x C°×]l-y Ary


10
O Vb=0V ~.~--
fi V b =-50V
zx V b =-100V ~
o

°,+//I
-5

Fig. 24. Dependence of the uniaxial anisotropy constant at room temperature on (a) the nitrogen A = N
and oxygen A = O content in amorphous Gdo.2+Tbo.o+Feo.72 films prepared by magnetron sputtering
(Heitmann et al. 1987a) and (b) on the argon pressure for amorphous (Gd, Tb)l _~Co~ films prepared by
diode sputtering at different substrate bias voltages (Heitmann et al. 1978b).

correlated to the sum of spin-up and spin-down d electrons outside the last filled
shell. The magnitude of the magnetic moment per atom and its temperature depen-
dence have been discussed in terms of two extreme pictures of fully localized electrons
(Heisenberg model) and fully itinerant electrons (Stoner model). The localized model
cannot account for the fractional values of the observed moments and leads to a
much too low Curie temperature, while the itinerant model gives rise to a Curie
temperature which is an order of magnitude too high and also the moments in the
paramagnetic state are completely destroyed. Thus, various approaches have been
made to improve or combine these theories. An essential element of the disordered
local moment theories (Hubbard 1979a,b, 1981a,b, Hasegawa 1979, 1980a,b, 1983,
1984, Oguchi et al. 1983, Moruzzi et al. 1986, Moruzzi and Marcus 1988) and the
local band theories (Capellmann 1974, Prange and Korenman 1975, 1976, 1979a,b,
Korenman et al. 1977a,b,c, Korenman 1985) is the introduction of a fluctuating mean
322 P. HANSEN

Sml.xCO x
1.4
o d= g4Onm
1.2 x d= 13/.Onrn
[] d= 6gOnm
1.0 A d= 230nm

<z
0.8 o
o

-r 0.6
L

0.4

0.2

0
~ 1 ! i
300 500 700 go0 1100
T~ {K)
Fig. 25. Coercive field versus substate temperature for Sm-Co alloys prepared by diode sputtering
(Munakata et al. 1984). The argon pressure was 0.076Torr for the 940nm thick film and 0.063 Tort for
the other films.

field, which means that a local mean field is introduced, whose direction may vary
in space and time. In the disordered local moment theories the mean fields are
assumed on lattice sites with no correlation among them, while for the local band
theories it is assumed that the excitations maintain sufficient short-range magnetic
order even above Tc. The real difference between these two viewpoints is not the
absolute scale of short-range magnetic order, but in whether or not local ordering
plays a significant role in the energetics of creating local magnetic strength (Korenman
1985).
Although these theories account for various basic features of ferromagnetism of
3d metals (Williams et al. 1982, 1983a,b), there remain still some unsolved problems.
Therefore, a basic theory for amorphous alloys with 3d metals which involves the
additional problem of structural disorder is not yet available.
Different models were developed to interpret the trends in magnetic property
variation with alloy content. The first two models discussed focus on the valence per
atom or the splitting of bands for alloys but take no account of the particular atomic
structure and nearest-neighbor environments. The bond model is based on the local
symmetry or coordination and the chemical bonding and was applied to explain
primarily the magnetic moment variation of transition-metal-metalloid crystalline
and glassy alloys (Corb et al. 1982, 1983, Corb 1985). The environment model
considers only the influence of the local coordination of the disordered structure but
does not account of any band features. First attempts to interpret the moment
variation of Slater-Pauling plots representing the moment versus composition or
electron concentration (fig. 26a) were based on a charge-transfer model (Mizoguchi
et al. 1973a,b). However, the better understanding of compositional effects in terms
of bonding and experimental XPS (X-ray photoelectron spectroscopy) and UPS
(ultra-violet photoelectron spectroscopy) work have formed a more realistic picture
of the influence of valence band structure and bonding on the moment variation of
MAGNETIC AMORPHOUS ALLOYS 323

alloys. In M - T alloys, (sp)-d bonding dominates where the hybridization between


metalloid sp states and metal d states reduces the degree of localization of d electrons
tending to broaden the d band into a more free-electron-like configuration
(O'Handley 1987a). The metal-metalloid (sp)-d hybridization feature is much more
significant for M - F e alloys than for M - C o alloys as shown by XPS measurements
(Amamou and Krill 1979, 1980). This is consistent with the higher degree of d state
localization in Co and Ni compared to Fe. Co and Ni, therefore, are expected to be
less sensitive to their environment than Fe. The magnetic properties are primarily
affected by p - d hybridization while s-d interaction is less significant. The strong p -
d covalent hybridization for C is expected to suppress the Fe moment as observed
for carbon additions in B-Fe glasses (Mitera et al. 1978, Kazama et al. 1978a,b).
This demonstrates the general trend that chemical bonding weakens magnetism
(O'Handley 1987a). For R - T alloys, the situation is slightly different due to the
highly localized nature of the 4f electrons.
Band-gap theory. The moment variation with the average number of electrons
(Slater-Pauling curve) was first discussed in the concept of virtual bond states
(Friedel 1958, Mott 1964). This model has been extended and reinterpreted (Terakura
and Kanamori 1971, Terakura 1976, 1977) and expressed in a more generalized form
referred as band-gap theory (Malozemoff et al. 1983, 1984, Williams et al. 1983a,b).
The atom-averaged magnetic moment in alloys of composition Ml-xTx with T =
Fe, Co or Ni and M the solute can be expressed in terms of the atom-averaged
number of spin-up and spin-down electrons,
fix = #a( NT -- N~)• (15)
The average atomic valence Z is related to N T and N ~ by
Z = N ~+ N 1. (16)
Combining eqs. (15) and (16), one obtains
f,~ = # . ( 2 N ~ - Z). (17)
Thereby, the total number of spin-up electrons consists of d and sp electrons and,
thus,
N ~ = N~ + N~p. (18)
In the alloy, Z has to be considered as an averaged number of the T and M atoms,
Z = (1 - X)ZM + xZx, (19)
and thus/~x can be expressed in the form (Malozemoff et al. 1984)
/ix(x) = #. [(1 - x)(2N~M -- ZM) + x(2N~T -- ZT) + 2N~p], (20)
where 2NTT = 10 for iron and elements to its right and 2N~r = 0 for elements to its
left. For #o = #.(2N~ T _ ZT)+ 2#.N~p and 2NTd. = 0, eq. (20) reduces to the form
(Friedel 1958)
fiT(X) = #o _ (1 -- x)(10 + ZM -- Zx)#.. (21)
324 P. HANSEN

This relation predicts a linear reduction of the average transition-metal moment


with increasing M concentration, 1 - x, where #o is the T moment in the pure metal.
Some typical averaged moments fiT per transition-metal atom at T = 4.2K are
compiled in tables 5 and 10. It is seen that fiT is decreased with the presence of
nonmagnetic elements as expected from eqs. (20) and (21). A typical Slater-Pauling
curve for amorphous alloys is shown in fig. 26a. This plot of/~T versus average atomic
valence reveals a shift of fiT for amorphous alloys towards smaller values of Z as
compared to that of crystalline alloys without metalloids (dotted lines). Figure 26b
shows the Fe moment variation for amorphous B-FeNi and BP-FeNi alloys (Kaul
1981a). A plot of fiT expressed in the form
= Zm + 2/v p, (22)
and expressed versus the average 'magnetic valence' Z m = ( I - - x ) ( 2 N ~ T - Z T ) +
x(2N~M-Z~) and N~p = 0.3 (full line) is presented in fig. 27 for amorphous B-Co
alloys (Malozemoff et al. 1984). The experimental data lie above the calculated line
[eq. (22)], indicating strong magnetism. This feature is confirmed by the continuous
moment and Tc variation across the phase change from crystalline to amorphous
for various Co- and Ni-based alloys (Egami 1984, Stein and Dietz 1989) while for
many Fe-based alloys a discontinuity was observed (Sumiyama et al. 1983, Stein and
Dietz 1989). Various crystalline (Williams et al. 1983a,b) and amorphous (Malozemoff
et al. 1983, 1984) alloys were interpreted in terms of this generalized Slater-Pauling

TABLE 5
Average transition-metal moment at T = 4.2 K for amorphous alloys of com-
position M1 _~T~ with x ~ 0.8.
Composition #(#B) Reference
Bo.2oFeo.8o 1.99 Hasegawa et al. (1976)
Po.2oFeo.so 2.1 O'Handley et al. (1977)
Sio.2oFeo.8o 2.23 Felsch (1969, 1970a)
Geo.2oFeo.8o 2.2 Kazama et al. (1978b).
Zro.2oFeo.8o 1.65 Buschow and Smit (1981)
Hfo.2Feo.8o 1.45 Buschow (1984b)
Yo.zoFeo.so 1.89 Coey et al. (1981)
Lao.24Feo.76 1.53 Heiman and Lee (1976)

Bo.2oC0o.8o 1.28 O'Handley et al. (1977)


Po.22 C0o.78 1.18 Cargill III and Cochrane (1974)
Mgo.2oCoo.8o 1.23 Buschow (1984b)
Zro.2oCoo.so 0.83
MOo.2oCOo.8o 0.48
Hfo.2oCOo.so 1.75 Buschow (1982b)
Nbo.2o COo.so 1.35
Yo.2oCoo.8o 1.65 Buschow et al. (1977)
Lao.2oCoo.8o 1.3 Heiman and Kazama (1978a)

Yo.lvNio.sa 0.04 McGuire and Gambino (1978)


Lao.2oNio.so 0.04 Buschow (1984b)
MAGNETIC AMORPHOUS ALLOYS 325

i i i I
.../Fe-Co
(o) ;ff:Z...~.,~ g e - Ni HI.xT x

¢o
x i
\, "..
; ,\
+
+
a6\\x
I I I [] I %'~ I
7(Mn) 8(Fe) 9(Co) 10(Ni) 11(Cu)
Avero.ge number of 3d electrons

2.75
*~'O'c~ BO.2o(F e× Nil-x )o.8o
"w, %..~
2.50
,--g
w... ",,,~
:&
v °%,Ji ]k.
,,= 2,25 "_'+, +,~
"q o'%
2.00
Po lt.Bo 06(Fex Nil-x )0.80 .%
• • • 'El,,
(b)
I I I I
1.75
0 0.25 0.50 0.75 1.00
X

Fig. 26. (a) Slater-Pauling plot for amorphous M 1-xT~ alloys (e) Bo.2o(Fe, Co)o.so, (&) Bo.2o(Fe, Nio.so),
(m) Bo.2o(Co, Ni)o.so (O'Handley et al. 1976a,b); (©) Po.2o(Fe, Co)o.so, (~) Po.2o(Fe, Ni)o.so, (D) Po.2o(Co,
Ni)o.so (O'Handley et al. 1977); (+) Bo.ls(Fe, Cr)o.ss (Dey et al. 1980); (x) Bo.2o(Fe, Mo)o.so (Sostarich et
al. 1982). (b) Compositional variation of the Fe moment for amorphous FeNi-B and BP-FeNi alloys
[B-FeNi: (©) Kaul (1981a), (e) Becket et al. (1977), (A) O'Handley et al. (1976a); PB-FeNi: (D) Kaul
(1981a), (m) Beeker et al. (1977)]. The broken lines represent the linear relations ~ro = 2.00 + 0.80(1 - x )
and #F, = 1.85 + 0.80(1 - x ) for B-FeNi and PB-FeNi, respectively (Kaul 1981a).

model. It was pointed out that Slater-Pauling-like behavior can be attributed to a


gap or minimum in the conduction-band density of states. Such a gap tends to
conserve the number of conduction electrons in an alloy series leading to a straight
line of #a- versus Zm- Although the systematics and rationalizing in the moment
variation was improved, various alloys show significant deviations from the predicted
linear dependence. This can be attributed to the limitations of the model that takes
insufficient account of the atomic structure, nearest-neighbor environments and
chemical disorder. The deviation from linearity is also obvious from the comparison
of the concentration dependence for amorphous alloys of composition M l - x C o x
with M = Mg, Zr or Mo (Buschow 1984b) and M = La, Y, HI', Zr, Ti, Ta, Nb, W or
Mo (Shiba et al. 1986) as shown in fig. 28a,b. The full lines in fig. 28a represent the
326 P. HANSEN

I S
2.5! B1-x Cox s'/
SS J
2.0 SJ S
SSJS
1.5

1.0
S~ x

SST
L~
0.5

I I I I I I
°_2 -1 0 1 2
Zm
Fig. 27. Generalized Slater-Pauling plot for amorphous Bl-=Cox alloys (Malozemoff et al. 1984).
(x, ©) Hasegawa and Ray (1979), (@) McGuire et al. (1980a), (A) Watanabe et al. (1978).

compositional variation predicted by eq. (21). Good agreement between the calculated
and measured data is achieved for Z r - C o alloys indicating the presence of a strong
ferromagnetism, while for M g - C o alloys a strong deviation from a linear concentra-
tion dependence occurs. Almost all Co-based alloys are good ferromagnets for x > xc
where xc represents the critical concentration for the appearance of collective mag-
netic order. Some amorphous alloys still exhibit ferromagnetism where their crystal-
line counterparts are already Pauli paramagnets, which applies, e.g., for Sn-Co
(Marchal et al. 1980), Ce-Co (Malterre et al. 1988) or M - C o alloys with M = Hf,
Nb or Ta (Buschow 1982b). This was attributed to the number of neighboring Co
atoms which are necessary for a Co atom to develop a moment.
Various Fe-based alloys are less strong magnets as compared to the corresponding
Co alloys which can be ascribed to the much higher sensitivity of the F e - F e exchange
interaction on structural and chemical disorder due to a higher degree of d state
delocalization. For bond lengths below 0.25nm, even antiferromagnetic coupling
occurs (Masumoto et al. 1980, Coey et al. 1981, Maeda et al. 1981, Yamauchi et al.
1984, Fukamichi et al. 1989b). The behavior of Fe alloys is demonstrated in fig. 29
for compositions M1-xFex with M = B (Fukamichi et al. 1978, Mitera et al. 1978,
Hasegawa and Ray 1978, Chien et al. 1979, Buschow and van Engen 1981a, Chien
and Unruh 1982, Stobiecki and Stobiecki 1983), M = P (Durand and Yung 1977,
Mitera et al. 1978, Fukamichi et al. 1978), M = Zr (Masumoto et al. 1980, Buschow
and Smit 1981, Fujimori et al. 1982, Shirakawa et al. 1983, Coey et al. 1984, Ryan
et al. 1987b), M = Y (Chappert et al. 1981, Coey et al. 1981, Buschow 1982b, Gignoux
et al. 1982) and M = Sc (Fukamichi et al. 1986a). In the case of B-Fe or P - F e shown
in fig. 29a, the extrapolation to x ~ 1 leads to #Fe ~ 2.3#B which is in good agreement
with the iron moment of crystalline iron. The moment variation in fig. 29b reveals a
critical concentration, x~, where the average Fe moment disappears. The magnitude
of x~ reflects the structural and chemical environment of the transition-metal atom.
In the range x > xc, B-Fe, P - F e and various other Fe-based alloys are ferromagnets
in contrast to Z r - F e alloys where /TFe passes through a maximum followed by a
sharp decline and approaching #Fe = 0 at a second critical concentration x~ = 0.95
MAGNETIC AMORPHOUS ALLOYS 327

~1_xCOx

M= g

o I

o •
0 M=

(al
0
0.2 0.4 0.6 0.8 1.0

• Lo Ml-xC°x Zr
1500 [] y 11
• Hf Nb Hf
o Zr To--'7 ] ~ "
® Ti tOo~o A
• To Ti~ n
1000 A Nb
E • W ° • . D4
~•
<
_~
vMo °a•
o ~
A
O
I" A
[] vA
500 o•®
mOA
A @

[] V

0"--0 I @I I i
0.50 0.60 0.70 0.80 0.90 1.00
X

Fig. 28. Concentration dependence of the average Co moment at T = 4.2K (Buschow 1984b) and (b) the
room-temperature magnetization (Shiba et al. 1986) for different amorphous alloys. The full line in (a)
represents Friedel's model [eq. (21)]. The arrows in (b) indicate the metal concentration necessary to
obtain amorphous films (Fujimori et al. 1984).

(Read et al. 1989). Compositions with x > 0.94 even reveal asperomagnetic order
(Ryan et al. 1987b). This is confirmed by amorphous Ce-Fe (Fukamichi et al. 1988),
La-Fe (Wakabayashi et al. 1987, 1989) and Hf-Fe (Hiroyoshi et al. 1985, Ryan et
al. 1987a) alloys. The magnetic phase diagram for Cel-xFex alloys is displayed in
fig. 30, demonstrating the limited compositional range for the existence of ferromag-
netism. At high x and low temperatures, a speromagnetic or spin-glass-like state is
present. In this figure, the phase boundaries are determined by the concentration
dependence of the Curie temperature at high temperature and by the spin-freezing
temperature at the low-temperature side with two tricritical points (Tc = Tf) at the
Fe-rich and the M-rich side. The tricritical point compositions for amorphous Fe-
328 P. HANSEN

I Ml-x Fex
M '

A 2.0

'~" 1,5
a) F I I I
1.00 0.95 0.90 0.85 0.80 0.75
X

2.5

M 1-xF ex
2.0

._~ 1.5

1.0

0.5

(b)
0
0.20 0.40 0,60 0.80 1,00
X

Fig. 29. Concentration dependence of the average iron moment for amorphous (a) B-Fe and P-Fe alloys
(Durand and Yung 1977) and (b) B-Fe (Chien and Unruh 1982), Y-Fe (Chappert et al. 1981, Coey et al.
1981) and Zr-Fe (Masumoto et al. 1980, Buschow and Smit 1981, Fujimori et al. 1982, Coey et al. 1984)
alloys.

rich M - F e alloys with M = Zr, Hf, Ce or La were found to be 0.065, 0.07, 0.1 and
0.1, respectively. The corresponding Tc values are 150, ,-450, 120 and l l 0 K . The
noncollinear structures associated with these low Curie temperatures for many Fe-
based alloys are due to antiferromagnetic exchange requiring high magnetic fields to
align the iron moments completely. Large single-ion anisotropy as present in R-T
alloys can be ruled out as an origin for the noncollinear structure because the spin-
orbit coupling is significantly lower for the T elements than for rare-earth elements.
In the case of Zr-Fe, M6ssbauer experiments indicate that below Tf only the
transverse spin components freeze (Ryan et al. 1987b). A less dramatic moment
reduction for x ~ 1 was also observed for Sc-Fe alloys (Felsch 1970b) or for rare-
earth-iron alloys discussed in section 5.3.2. An intermediate situation occurs for
amorphous Y-Fe alloys which also reveal a noncollinear magnetic structure
(Chappert et al. 1981). This can be well studied by M6ssbauer spectroscopy because
the influence of the local environment on the iron moment is directly reflected in the
hyperfine field distribution P(Hw)(Rodmarcq et al. 1980, Chappert et al. 1981, Chien
and Unruh 1982). The M6ssbauer spectrum of amorphous Yl-xFex alloys at T =
1.6K (Chappert et al. 1981) is shown in fig. 31. In this case, the moment variation
can be directly derived from the hyperfine-fleld distribution because the proportional-
MAGNETIC AMORPHOUS ALLOYS 329

250
Cel. x Fex

200

para

150

i
i ferro
100

50
, J
spin glass

0 I I I I
0.50 0,60 0.70 0.80 0.90 1.00
X
Fig. 30. Magnetic phase diagram for amorphous Ce-Fe alloys (Fukamichi et al. 1988).

>
t~
>

o~

E "1-
c

k I I I
-6 -3 0 6 0 1 2 3
v(mm/s) Hhf (107A/m)

Fig. 31. Mrssbauer spectra of amorphous Y-Fe alloys at T = 1.6K. The corresponding hyperfine-field
distributions are shown on the right-hand side of the figure, indicating a relatively increasing fraction of
nonmagnetic iron (shaded area) with a decreasing iron content (Chappert et al. 1981).
330 P. H A N S E N

ity between hyperfine field and magnetic moment was established from crystalline
Y-Fe alloys (Gubbens et al. 1974). The hyperfine-field distribution is shown on the
right-hand side of fig. 31 and the shaded area indicates the relatively increasing
fraction of nonmagnetic iron with decreasing x. The analysis of the spectra reveal
an asperomagnetic order in the entire range of concentrations x > xc as demonstrated
in fig. 32. This random ferromagnetism again can be attributed to a broad distribution
of Fe-Fe exchange interactions including a significant fraction of antiferromagnetic
bonds.
The moment variation for amorphous La-Ni and Hf-Ni alloys (Buschow 1984b)
is presented in fig. 33. Very similar results have been found for amorphous Y-Ni
and Ce-Ni alloys (Fr6my et al. 1984). These alloys are ferromagnetic, but the onset
of magnetic order takes place at much higher concentrations as compared to the
corresponding Fe or Co alloys.
Split bands. Valence bands of Ml-xTx alloys split into two resolvable parts at
different energies when the difference in atomic number is greater or equal to two
(Beebey 1964, Velicky et al. 1968). The split-band character is a consequence of
polar d-d bonding and was experimentally confirmed for various amorphous alloys
(Oelhafen et al. 1979, 1980). The concept of split bands, originally applied to crystal-

Yl-xFex

x=0.32 0.57 0.71 0.82

Fig. 32. Schematic representation of speromagnetic structures in amorphous Y-Fe alloys (Chappert et al.
1981).

0.6 Ml_xNix /

"-b O.Z,
,:&
0.2 j ~ M=Hf

0 0.8 0.9 1.0


0.7
x

Fig. 33. Concentration dependence of the average Ni moment in amorphous Hf-Ni and La-Ni alloys at
T = 4.2 K (Buschow 1984b).
MAGNETIC AMORPHOUSALLOYS 331

line alloys (Berger 1977, Berger and Bergmann 1980), was extended to amorphous
alloys (O'Handley and Berger 1978). The interpretation of magnetostriction
(O'Handley 1978a, O'Handley and Berger 1978) and Hall effect (O'Handley 1978b)
measurements for amorphous Bo.2o(FeCoNi)o.8 o alloys indicate that Fe and Ni
indeed contribute to the valence-band density of states at different energies and thus
a minimum in the density of states is present. This model was shown to account well
for the compositional variation of the magnetostriction and spontaneous Hall effect
in these alloys.
Coordination bond model. This model was derived from valence-bond theory to
account for the magnetic moment variation of crystalline and glassy transition-
metal-metalloid (M-T) alloys (Corb et al. 1982, 1983, Corb 1985). It assumes that
the average T moment is suppressed according to the number of M neighbors and
the strength of the M - T bonds. The model is based on a strong p - d bonding, causing
a T moment reduction of one fifth of its moment for each 3d electron participating
in the nonmagnetic covalent T - M bond. This leads to an average T moment
suppressing for M1 -xTx alloys of the form

fiT(X) = #° [1 -- (1 + ~--~-~)(1 -- 1)1, (23)

representing a linear decrease of fiT with the metalloid concentration 1 - x. Z~ is the


number of T atoms surrounding an M atom. This model accounts for the moment
variation for various M - T alloys (Corb et al. 1982, 1983, Corb 1985, O'Handley
1987a, Stein and Dietz 1989), however, the band-gap model appears to be able to
explain more general trends in a wider class of alloys (O'Handley 1987a). Although
these models are contradictory in their basic origin, both lead to a linear decrease
of 17 according to the relation
fix(X) = #°1-1 - p(1 - x)],
where the slope p differs according to the assumptions made in these models.
Environment model. The third picture to interpret the moment variation is based
on the structural disorder and accounts for the different environments of a transition
metal (Jaccarino and Walker 1965). The probability of a T atom to have j nearest T
neighbors out of a maximum number n is given by
n!
P(x, k,n) - k!(n - k)!" (24)

Assuming a minimum number j of nearest T neighbors to be necessary for a T


atom to carry a moment #1 and for less than j T neighbors a moment #2, then fi(x)
can be expressed in the form
fi(x) = #1Pj(x) + #2 [1 - P~(x)], (25)
where

Pj(x) = ~ P(x, k, n).


k=j
332 P. HANSEN

In the case #1,/~2 ~ 0, no critical concentration is expected. For #2 = 0, a critical


concentration can be calculated from the tangent of the turning point of/i(x) at
x = xt by xc = xo/P}(xt) where xt is determined by Pj'(xt)= 0.
The comparison between experimental data for amorphous Y - F e (Chappert et al.
1981, Coey et al. 1981), Y - C o (Buschow et al. 1977) and S n - C o (Teirlinck 1981) and
the environment model is shown in fig. 34. The full lines represent the calculated
result with j = 7 for Fe, j = 8 for Co in Y and j = 9 for Co in Sn assuming #2 = 0.
This means that at least seven nearest Fe neighbors and eight or nine nearest Co

2.5
Y1-x mx

2.0

1.5
m

1.0

0.5

0
0 0.2 0.4 0.6 0.8 1.0
X

150 Snl_ x Co x •

100
<
E

50

Ib)
0
0.2 0.4 0.6 0.8 1.0
X

Fig. 34. Concentration dependence of (a) the average transition-metal moment for amorphous Y-Fe,
(El) Chappert et al. (1981), ( I ) Coey et al. (1981); Y-Co, (O) Buschow et al. (1977) and (b) the specific
magnetization for amorphous S n - C o alloys (Teirlinck 1981) at T = 4.2K. The full lines were calculated
from the environment model I-eq. (24)] with #2 = 0 and j = 7 for Fe, j = 8 for Co in Y and j = 9 for Co in
Sn. In the latter case, the density variation was not taken into account.
MAGNETIC AMORPHOUSALLOYS 333

neighbors are necessary to carry a magnetic moment for a Fe or Co atom, respectively.


Nuclear magnetic resonance experiments for amorphous Y-Co alloys reveal already
a disappearance of the Co moment for j = 6 (Alameda et al. 1985). Ni atoms require
an even higher number of nearest neighbors to develop a moment, which is obvious
from their higher xc values because xc increases with increasing ].
The models so far discussed can account for the moment variation for various
alloys, but there are also various counter examples where/~(x) cannot be explained
in terms of one of these models. This is associated with the extreme assumptions
that are made in the case of the band-gap model disregarding certain influences of
structure and local environment and in the case of the coordination bond or environ-
ment model where particular band features are ignored.
The appearance of magnetism at xc reflects the influence of chemical environment
and structural disorder. A significant fraction of transition-metal atoms in amorphous
alloys experience a larger number of nearest-neighbor T atoms as compared to the
corresponding crystalline alloys leading to higher xc values. Also, compositional
short-range order plays a prominent role and from the heat of formation it can be
concluded that its influence increases from Mg to Zr, leading to an increase of M
neighbors that tends to increase x¢. Various xo values for M-rich amorphous alloys
are compiled in table 6. They are deduced from the compositional variation of the
saturation magnetization, the Curie temperatures or hyperfine fields. Many alloys
show spin-glass behavior for M-rich compositions close to xo as, e.g., for M = Sn
(Piecuch et al. 1983) or M = Sb (Xiao and Chien 1985). In that case, the smaller
value in the brackets in table 6 indicates the onset of spin-glass behavior and the
larger value in the brackets represents the tricritical point (Tc = Tf) where long-range
magnetic order appears. The x~ values calculated from the band-gap theory are given
in the last column. These values roughly reproduce the observed trends of an
increasing x~ from Fe- to Ni-based alloys and for alloys containing elements from
group II to group V. This confirms the importance of the chemical environment and
disorder for the onset of collective magnetic order and the magnitude of the transition-
metal moment as it is already well-known for crystalline alloys as, e.g., for PtCo
(Sanchez et al. 1989).
The models and experimental results suggest that magnetic order in one-subnet-
work T-based alloys is favored by (O'Handley 1987a,b) (i) a strongly positive heat
of formation for like-atom clustering, (ii) minimum p - d hybridization, (iii) more
negative average alloy valence and hence more positive magnetic valence and moment
and (iv) a small interatomic spacing for Co-rich alloys but a larger spacing for Fe-
rich alloys (negative exchange coupling at small Fe-Fe distances).
Many other amorphous alloys not discussed in this section have been studied
intensively and the data were compiled in different reviews (Buschow 1984a, Moorjani
and Coey 1984, Coey and Ryan 1984, Egami 1984, Hansen 1988a).

5.3.1.2. Rare-earth-based alloys. The alloys of composition Rl-xMx, where R rep-


resents the rare earths carrying a magnetic moment, form a second class of magnetic
one-subnetwork alloys. The magnetism is controlled by the localized 4f electrons and
the 5d conduction electrons. However, in contrast to the transition-metal-based
334 P. HANSEN

TABLE 6
Critical concentrations for amorphous alloys of composition M1 -xTx with T = Fe, Co or Ni. The values
in the last column have been calculated from the band-gap theory (Malozemoff et al. 1983, 1984). The
values in the brackets represent tricritical points (Tc = Tf) where the Curie temperature equals the spin-
freezing temperature.
Group M Fe Co Ni xc (Theory)
xc Ref.* xc Ref.* xc Ref.* Fe Co Ni
II Mg - <0.2 [2] 0.7 [6] 0.35 0.47 0.70

III B 0.3,0.4 [1] ~0.6 [18] - 0.48 0.60 0.80


Y 0.4 [5,6] 0.5 [19] 0.8 [22]
La - 0.45 [13] 0.8 [13]

IV Si 0.4 [4] 0.5 [3] - 0.57 0.68 0.85


Ge 0.3,0.4 [1,4] - -
Sn 0.4(0.25) [1,7,8] 0.5 [4] -
Ti 0.45,0.5 [9,10,25] 0.7 [20] -
Zr 0.4,0.95 [11,24] 0.65 [13] 0.8 [22]
Mo 0.5 [12] 0.65 [13] -
Hf 0.4 [13] 0.5 [6] 0.85 [13]
Th 0.4 [14] - -

V P 0.45 [151 0.7 [211 0.8 [231 0.63 0.73 0.88


Sb 0.5(0.4) [16] - -
Nb 0.65 [17] 0.6 [6] -
Ta 0.6 [12] 0.7 [6] -
* References:
[1] Buschow and van Engen (1981a). [14] Buschow et al. (1978).
[2] Buschow and van Engen (1981b). [15] Coey and Ryan (1984).
[3] Buschow and van Engen (1981c). [16] Xiao and Chien (1985)
[4] Teirlinck (1981). [17] Chien et al. (1983),
[5] Chappert et al. (1981). Chien and Unruh (1982).
[6] Buschow (1982b). [18] Hasegawa and Ray (1979).
[7] Teirlinck et al. (1981). [19] Heiman and Kazama (1978a).
[8] Piecuch et al. (1983). [20] Aboaf and Klokholm (1981).
[9] Sumiyama et al. (1983). [21] Pan and Yurnbull (1974).
[10] Liou and Chien (1984). [22] Moorjani and Coey (1984).
[11] Heiman and Kazama (1979). [23] Berrada et al. (1977).
[12] Xiao and Chien (1987). [24] Read et al. (1989).
[13] Buschow (1984b). [25] Sumiyama et al. (1990).

alloys, t h e t y p e o f m a g n e t i c o r d e r is d e t e r m i n e d b y the e x c h a n g e i n t e r a c t i o n a n d the


s t r o n g s i n g l e - i o n a n i s o t r o p y . T h e l a t t e r d o m i n a t e s for m o s t r a r e earths, e x c e p t for
S-state G d w i t h a s m a l l a n i s o t r o p y . T h e q u a n t u m n u m b e r s , i o n i c r a d i i a n d the de
G e n n e s f a c t o r for the r a r e e a r t h s in t h e i r m o s t c o m m o n c o n f i g u r a t i o n s are g i v e n in
t a b l e 7.
The Hamiltonian describing the exchange interaction between two atoms with
spins Si a n d Sj g e n e r a l l y is o f the f o r m

= - ~ J~S~Sj, (26)
i>j
MAGNETIC AMORPHOUS ALLOYS 335

TABLE 7
Quantum numbers, ionic radii and de Gennes factors for the rare-earth elements in their most common
configurations (Moorjani and Coey 1984).
Valence 4f" Radius (nm) L S J g gJ (g -- 1)2J(J + 1)
ionic metallic
Y 3+ 0.092 0.180
La 3+ 0 0.114 0.183 0 0 0
Ce 4+ 0 0.094 0.171 0 0 0
Pr 3+ 2 0.106 0.183 5 1 4 4 3.20 0.80
Nd 3+ 3 0.104 0.182 6 ~ 9 Tr8 3.27 1.84
Sm 3+ 4 0.100 0.180 5 ~ ~ z7 0.71 4.46
Eu 2+ 7 0.109 0.204 0 ~} ~ 2 7.00 15.75
Gd 3+ 7 0.094 0.180 0 7 7 2 7.00 15.75
Tb 3+ 8 0.093 0.178 3 3 6 3_
2 9.00 10.50
Dy 3+ 9 0.092 0.177 5 ~ ~-15 _43 10.00 7.08
Ho 3+ 10 0.091 0.177 6 2 8 ~ 10.00 4.50
Er 3+ l1 0.089 0.176 6 ~ ~-15 _6s 9.00 2.55
Tm 3+ 12 0.087 0.175 5 1 6 7 7.00 1.17
Yb 3+ 13 0.086 0.173 3 ½ ~ -~ 4.00 0.32
Lu 3+ 14 0.085 0.173 0 0 0

and can proceed either t h r o u g h a positive intra-atomic 4 f - 5 d exchange combined


with a direct positive interatomic 5 d - 5 d exchange (Campbell 1972) leading to ferro-
magnetic order or indirectly via spin-polarized conduction electrons where b o t h
ferromagnetic and antiferromagnetic order can occur. The latter interaction is k n o w n
as R K K Y ( R u d e r m a n n - K i t t e l - K a s u y a - Y o s i d a ) interaction ( R u d e r m a n n and Kittel
1954, K a s u y a 1956, Yosida 1957). The exchange coupling constant for disordered
alloys can be expressed by (de Gennes 1962a,b)
Jij ~ j2fF(¢u) e-'i~/z,

where

¢ij cos ~i~ - sin ~ij


=

Jsf is the rare-earth conduction-electron interaction, 2 is the electron m e a n free


path and ~ij = 2 k F r i j where kF is the Fermi m o m e n t u m and r~j = tri - rjl. The struc-
tural disorder leads to a distribution of rij and thus to a distribution of exchange
coupling constants. These can be positive or negative because F(~ij) is an oscillating
function. A p r e d o m i n a n t l y positive distribution results in a ferromagnetic coupling.
W h e n the distribution is almost zero, then spin-glass behavior is expected involving
b o t h positive and negative exchange constants. The first case is realized for Gd-rich
alloys while the second was observed for m a n y diluted systems (Moorjani and C o e y
1984).
The locally varying electrostatic fields and the strong s p i n - o r b i t coupling of the
non-S-state rare earths give rise to a strong locally changing anisotropy inducing a
varying orientation of magnetic moments. The electrostatic field H a m i l t o n i a n can
336 P. HANSEN

be expressed generally in terms of tensor operators O~,,


oetoe= ~ B.O.,
k k (27)
k,n

which are directly related to the angular momentum operators (Stevens 1952). The
sum over n is restricted to even terms and to n ~<4 for d electrons and n <~6 for f
electrons. The general case of structurally disordered alloys cannot be solved. There-
fore, it has been assumed that the atomic site retains a high degree of local symmetry
but the local axes of the crystal field will vary at random from one site to another.
For an axial symmetry only even order terms occur. Choosing the local axis at site
i along the direction n~ of the lowest energy of the moment, then ~ ) reduces to the
simple form
~f(i) = __ Di(n i . ji)2, (28)
where D = - 3 B ° and D > 0. The distribution of second-order electrostatic fields
correspondingly leads to a distribution of energy level splitting controlling the local
anisotropy (Cochrane et al. 1974, Cochrane et al. 1978a, Fert and Campbell 1978,
Czjzek et al. 1981).
The influence of exchange interaction and a random axial anisotropy on the
average magnetic moment can now be studied using the Hamiltonian (Harris et al.
1973),
= - Z J~jJi" J~ - ~ Di(n~" g~) - g#B ~, H . J~, (29)
i>j " i

where the third term on the right-hand side represents the field energy due to an
externally applied field. In the classical limit, the energy at T -- 0 K can be expressed
by
Ei = - # o ( H + Hex~h)COSOi -- D cos2(~bi - 0~), (30)
where the exchange interaction was treated in the molecular field approximation. It
was assumed that D~ has the same value at each site. 0i and ~b~represent the angles
between the direction of the magnetic field and the magnetic moment and the local
easy axis, respectively. From the condition dEi/dO~ = 0 and the spherical average
nl2

#= I
do
cos Oi sin q~id~bi,

the reduced magnetic moment m---#/#o = ( J z ) / J can be calculated (Callen et al.


1977). In the absence of any magnetic field, a speromagnetic order is expected. The
presence of a magnetic field tends to align all magnetic moments, forming a random
hemisphere around the positive field direction giving rise to an asperomagnetic order
where the cone angle decreases with increasing field. Experimental studies show,
however, that significant barriers are present to rotate the moments in the field
direction and in certain cases large fields are necessary to reach the asperomagnetic
state.
Various features of this random axial anisotropy model have been studied (Callen
M A G N E T I C A M O R P H O U S ALLOYS 337

et al. 1977, Chi and Alben 1977, Patterson et al. 1978, Chi and Egami 1979, Harris
1980, Fibich 1990).
The field dependence of the reduced magnetization resulting from this simplified
model (Coey 1978) is shown in fig. 35. The curve for J = ~ corresponds to the
classical limit. With increasing field, the fanning angle of the asperomagnetic structure
reduces and saturation can only be reached when the applied field exceeds the local
anisotropy field significantly. However, it turns out that for most amorphous
alloys of composition Rl-xMx, this limit cannot be reached even in fields up to
2.5 × 10 7 A/m as demonstrated in fig. 36 showing the reduced magnetization at T =
4.2K for M = Ag and x = 0.5 (Boucher 1976, 1977, Pappa 1979). The broken lines
represent the random anisotropy model reproducing the experimental data quite

1.00 J=8

0.80

E 0.60

I
0.40

0.20

I I I
0 1 2 3
g/.ZBH/DJ

Fig. 35. Relative magnetization curves for an array of magnetic ions with random easy axis anisotropy at
T = OK (Coey 1978).

1.0 f ~ _': ":'_ " ~. -- " R= Gd


//"

oot.?:/./,---
E trf ¢ R0,0Ag0,0

H(10/A/m)
Fig. 36. Reduced magnetization for amorphous Ro.soAgo.so alloys at T = 4 . 2 K versus magnetic field
(Boucher 1976, 1977, Pappa 1979). The broken lines are fits based on the random anisotropy model.
338 P. HANSEN

well in fields greater than 8 x 10 6 A/m which is the field needed to establish the
asperomagnetic state. Magnetic saturation can be achieved only for Gd-based alloys.
In this case, the local anisotropy is small and the next-nearest-neighbor exchange
coupling leads to ferromagnetic order. The compositional variation of the saturation
magnetization for amorphous Gdl -x Gex alloys at T = 4.2 K (Gambino and McGuire
1983) is shown in fig. 37. A critical composition for the disappearance of ferromagne-
tism can be estimated from the fall-off of Ms yielding xc --- 0.55. Similar results were
observedfor Gd-A1 (Mizoguchi et al. 1977a,b, McGuire et al. 1978), Gd-Ag (Hauser
1975, Boucher 1976, Pappa 1979), Gd-Au (Poon and Durand 1977a, Gambino et
al. 1981), Gd-Cu (Heiman and Lee 1976, Heiman and Kazama 1978b, McGuire et
al. 1978), and Gd-Si (Simonnin et al. 1986), revealing xc values between 0.2 and 0.4.
For x < x~, the nearest-neighbor coupling is no longer dominant and an intermediate
range occur where a significant portion of antiferromagnetic interaction is present.
With increasing dilution, the cluster glass region and the dilute spin-glass region
(x < 0.01) appears.
The magnetic properties of alloys containing non-S-state rare earths are controlled
by the large random anisotropy and negative exchange interactions. Different systems
have been investigated in more detail such as R-Au (Poon and Durand 1977a,b,
1978, McGuire and Gambino 1979, Berrada et al. 1979, Friedt et al. 1980, Hadjipa-
nayis et al. 1981), R-Ag (Boucher 1977, Pappa 1979), R-Si (Hauser 1986, Simonnin
et al. 1986), R-Cu (Heiman and Kazama 1978b, McGuire and Gambino 1979).
Various alloys have been discussed in terms of the random anisotropy model (Ferrer
et al. 1978) and good agreement with the experimental data was found for the high-
field region while disagreement occurs in the low-field region which can be attributed
either to hysteresis and magnetic after-effects. The predictions of this theory are
inadequate to account for the magnetic properties of the light rare earths. Low-
temperature magnetic moments are compiled in table 8 for some R-M compositions.

2 Gdl_xGex

.<

0 I
0 0.25 0,50 1 0,75 1.00
x Xc
Fig. 37. Compositional variation of the saturation magnetization for amorphous Gd-Ge alloys at
T = 4.2K (Gambino and McGuire 1983).
MAGNETIC AMORPHOUS ALLOYS 339

TABLE 8
Curie temperature and low-temperature magnetic moment for amorphous R - M alloys with nonmagnetic
M elements. The numbers in the brackets represent Curie or N~el temperatures of the corresponding
crystalline compounds.
R~M~_~ Tc(K) ~R(#B) Re~rences
Euo.ToMgo.3o 131 6.85 Maurer and Friedt (1983)
Euo.75Zno.25 134 6.8
Euo.7oZno.3o 133 6.85
Euo.81 Ago.19 70 3.8 Buschow and van der Hoogenhof (1979)
Euo.75 Ago.25 37 7.39
Euo.Ts Cdo.z5 134 6.8 Maurer and Friedt (1983)
Gdo.60Mno.40 230 5.3 Buschow et al. (1980)
Gdo.76Cu0.24 142 6.4
Gdo.voCuo.ao 144 6.6
Gdo.ss Gao.ls 148 5.8
Gdo.s5 Ruo.15 78 6.1
Gdo.7o Ruo.3o 77 7.1
Gdo.6o Ruo.4o 57 6.0
Gdo.sz Rho.18 111 5.6
Gdo.54Ago.46 122(138) 6.7 Boucher (1976)
Gdo.83 Pto.17 150 6.5 Buschow et al. (1980)
Gdo.aoAuo.2o 149 7.0 Poon and Durand (1977b)
Tbo. 50Ago,50 64(106) 5.0 Boucher (1977)
Dyo.soAgo.5o 18(59) 4.8
Hoo.~oAgo.5o 11(33) 5.7
Ero. 50Ago.50 6(20) 4.9

5.3.2. Two-subnetwork alloys


In rare-earth-transition-metal alloys R 1 xTx, both the R and T atoms contribute
to the magnetic properties. Their spins are coupled antiparallel via a strong exchange
coupling between the 3d electrons of the transition metal and the 5d electrons of the
rare earth where the latter are polarized by the 4f shell (Campbell 1972). This leads
to a parallel alignment of the T and R moments for light rare earths (J = L - S) and
to an antiparallel alignment for the heavy rare earths (J = L + S). Thus, for Gd-based
alloys a collinear order is present leading to ferrimagnetism. However, as in the case
of one-subnetwork alloys the presence of non-S-state rare earths involves local
random electrostatic fields giving rise to large local anisotropies competing with the
exchange interaction. This leads to spero-, aspero- or sperimagnetic order as
illustrated in fig. 21. Some typical data for the rare earth ions are compiled in table 7.
The average magnetic moment of an R1 _xTx alloy is given by
]~av = 1(1 -- X)IAR ~ XfiTI, (31)
where + and - refer to light and heavy rare earths, respectively. The average rare-
earth moment,/~R, is controlled by the localized 4f electrons. The average transition-
metal moment,/~a-, is determined by the structural and chemical environment and
by hybridization and mixing effects. Data for/Tav of some amorphous and crystalline
alloys are listed in table 9. The kTavfor many amorphous alloys are lower than for
340 P. HANSEN

TABLE 9
Averaged magnetic moments/7~ = (1 -X)~R- x/Tr at T = 4.2 K for amorphous and crystalline R-Fe and
R-Co alloys (Cochrane et al. 1978a).
Amorphous x /~av( # B ) Comparable #~ (/~B) Ref.*
alloy crystalline
compound
Gdl _~Co~ 0.67 1.4 GdCo2 1.67 [1]
0.75 0.7 GdCoa 0.55
0.78 0.47 Gd 2Co7 0.27
0.79 0.27 [2]
0.83 0 GdCo5 0.2 [1]
Tbl _xFe~ 0.67 1.57 TbFe z 1.57 1-3]
Tbl _xCo~ 0.67 1.43 TbCoz 2.23 [1]
0.75 0.63 TbCo 3 0.85
0.83 0.4 TbCos 0.08
Dya _xFe~ 0.79 0.4 [4]
0.67 1.5 DyFez 1.83 [3]
Dyl _xCo~ 0.77 0.53 1-2]
Ho 1_xFex 0.67 0.67 HoFe 2 1.83 [1]
0.75 0.70 HoFe3 1.15
Hol _~Co~ 0.67 1.6 HoCo2 2.6
0.75 1.08 HoCo3 1.4
0.77 0.79 [2]
Era _~Cox 0.74 0.48
* References:
1-1] Lee and Heiman (1975). I-3] Rhyne et al. (1974b).
1-2] Jouve et al. (1976). [4] Rebouillat et al. (1977).

the crystalline counterparts. The compositional variation of the saturation magne-


tization

Ms = IM, - Mx[, (32)


and the transition-metal spin value ST = ~T/gx#B at T = 4 . 2 K (Hansen et al. 1989)
are shown in fig. 38. Ms is simply related to/Tav by the n u m b e r of atoms N per unit
volume V (Ms = N(qv/V expressed in e m u / c m 3 corresponds to 1 k A / m in SI units).
gT denotes the g-factor of the T atoms and Ma- and MR are the T and R sublattice
magnetizations, respectively. The ferrimagnetic order of these alloys leads to a
compositional compensation at X¢omp where Ms = 0 and g-rSr=gRSR. Xcomp is a
sensitive function of composition. The MR sublattice dominates for x < Xcompand the
MT sublattice for x > Xcomp. The triangles in fig. 38a represent single crystalline Ms
data reduced by 5% to a c c o u n t for the lower density of the evaporated films. The
data for Sa- displayed in fig. 38b are calculated from eq. (31) assuming SGd = 7 and
gGd = 2. The full lines in fig. 38b are calculated from the environment model (Jaccarino
and Walker 1965) w i t h j = 5 a n d j = 7 for the G d - F e and G d - C o alloys, respectively.
Thus, less nearest T neighbors are necessary for the onset of magnetism as c o m p a r e d
to the one-subnetwork alloys like Y - F e and Y - C o . The dotted lines indicate the
critical composition for the appearance of a magnetic m o m e n t at the T a t o m yielding
xo ~ 0.4 for T = Fe and T = Co, respectively. These slightly lower x~ and j values
M A G N E T I C A M O R P H O U S ALLOYS 341

2000

1500 /Gdl.xCox I
Gdl.xFex/

100(

500

(a)
l i l t I

0.2 0.4 0.6 0.8 1.0


X

1.00 o°O°°
Gdl_xFex, ~ ~ -
,0.75 / ~

,.n0.50

0.25

00 0.2 0.4 0.6 0.8 1.0


X

Fig. 38. Compositional variation of (a) the saturation magnetization and (b) the spin value for amorphous
G d - F e and G d - C o alloys at T = 4.2 K (Hansen et al. 1989). The full squares in (a) were taken from the
work of Taylor (1976). The full lines in (b) were calculated from the environment model.

with respect to the one-subnetwork alloys (see table 6) can be attributed to the
additional R T exchange coupling but also reflects the dominance of the strong T-
T exchange coupling. The variation of Ms and the average transition-metal moment
at T = 4 . 2 K for amorphous Gdl_~T x alloys with T = F e , Co or Ni and
342 P. HANSEN

Gdl-x(FeyCot-r)=, Gdl-=(F%Nil-y)~ and Gdl_=(CorNil_y)x alloys with


0.17 ~<x ~<0.22 and y ranging from 0.19 to 0.94 (Taylor and Gangulee 1980) are
presented in fig. 39. The variation of the average transition-metal moment with the
averaged number of 3d electrons corresponds to the Slater-Pauling curve of the one-
subnetwork alloys (see fig. 26), but the moments are slightly shifted towards a higher
number of 3d electrons except for the Gd-FeNi alloys. This shift can be attributed
to the additional R - T exchange interaction. From the binary alloys, the moments
for Fe, Co and Ni can be calculated yielding fiFe ~-" 2.09#B,/iCo = 1.45#B and ]~Ni ~-
0.31#B.
Amorphous Gdl_~Mn~ alloys were also found to exhibit ferrimagnetic order

800
O Gd-T
600 n Gd-FeCo o° /
o Gd-CoNi o/
A
A Gd-FeNi /
400
<

:~ 200
i & [] A

-200
(a)
I

6{Fe) 7(Co) 8{Ni)


Averag e number of 3d electrons

2.5

a, Gd-T
~o e • Gd-FeCo
2.0 o Gd-CoNi
• x Gd-FeNi

A 1.5
A

8
,:t x
o
1.0 o
o
o
o
o
o
0.5 ×

(b)

0 I I I
6(Fe} 7[Co) 8(Ni)
Average number of 3d electrons
Fig. 39. (a) Difference of sublattice magnetization and (b) transition-metal moment versus average number
of 3d electrons for amorphous G d - T alloys with T = Fe, Co or Ni and G d - F e C o , G d - F e N i and
G d - C o N i alloys at T = 4.2K (Taylor and Gangulee 1980).
M A G N E T I C A M O R P H O U S ALLOYS 343

(McGuire and Taylor 1979) as opposed to the ferromagnetic order in crystalline


alloys (Kirchmayr and Steiner 1971). The Mn moment depends on composition
going from 2#B to 0.5#B per Mn for 0.8 <~ x ~< 0.2. The magnetic moment behavior
of Mn in G d - M n is similar to that found in crystalline alloys where a well-developed
moment occurs only when the transition-metal atoms are isolated from one another.
The moment variation for amorphous iron alloys of light rare-earth elements (Dai
et al. 1985, 1986, Miyazaki et al. 1987a, 1988) and for cobalt alloys (Dai et al. 1985,
Takahashi et al. 1988a,b) is shown in fig. 40. The moments were deduced from
magnetization measurements in fields ~<1.4 x 106 A/m and were extrapolated to T =
0 K from temperature-dependent measurements between 77 and 450 K, yielding the
/i T data /iPr = 0.35#B, ~Nd=0.26#B, fiSm= 0. These values were determined from
eq. (31), assuming compositionally independent moments for the R atoms. The broken
lines in fig. 40a,b were inferred from fig. 38. The much higher/iFe for N d - F e alloys
can be attributed to the very low Nd moment used to evaluate these data. A higher
fiNa (Taylor et al. 1978, Dai et al. 1985) reduces the #tEe data to those for the Pr-Fe

3.0

Rl-x Fe× o
2.0
-%

t~

1.0

0.20 0.40 0.60 0.80 1.00


X

2.0

RI xCox R=Pr
1.5
-% Nd s"

•3 1.0
I::&

0.5

b)
I I I
0.20 O.Z,O 0.60 0.80 1.00
X

Fig. 40. Compositional variation of (a) #F~ (Miyazaki et al. 1987a, 1988) and (b) #co (Takahashi et al.
1988a,b) for amorphous R-Fe and R-Co alloys with R = Nd, Pr or Sm at T = 0 K.
344 P. HANSEN

200
Prl_ x F e x

160 x:O.90

~ 120

~ 80

40

I I I
0.2 0.4 0.6 0.8
H{107 A/m)
Fig. 41. Field dependence of the specific magnetization for amorphous Pr-Fe alloys at T = 20 K (Croat
1981a).

alloys. A fit of low-field magnetization data for N d - F e alloys leads to a good


agreement between experiment and calculated data using eq. (23) with/~vel = 2. I#B,
#ve2 = 1.2/~a, find = 1.13#B andj = 6 (Dai et al. 1985). Generally, however, the compo-
sitional variation of fiFe differs from that of the G d - T alloy in particular in the
rare-earth-rich regime. This can be associated with the difficulty to determine accurate
moments per atom for non-S-state rare-earth alloys due to noncollinear subnetworks.
Very high fields are necessary to align the R moments as demonstrated for the
specific magnetization o- (a = Ms/p expressed in emu/g corresponds to Am2/kg in SI
units, where p is the density) in fig. 41 for amorphous Prl-xFex alloys (Croat 1981a).
Therefore, at low fields, the contribution of the almost speromagnetic Pr (Croat
1981a, Wan et al. 1990) and Nd (Croat 1981b, Taylor et al. 1978, Siratori et al. 1990)
subnetwork to the net magnetization at low temperatures is quite small. This is not
in agreement with the random anisotropy model predicting a remanence of the
relative magnetization of about 0.5 at T = 0 K and H = 0. An additional complication
arises in R-Fe alloys from the Fe subnetwork for compositions approaching the
critical concentration xe. In this range of compositions, an increasing portion of the
Fe subnetwork exhibits an asperomagnetic behavior inducing an additional effective
dependence on composition when transition-metal moments are evaluated from low-
field magnetization data.

5.4. Temperature dependence of the magnetization

5.4.1. Models
5.4.1.1. Mean-field theory. The lack of a general analytic expression to predict the
temperature dependence of M s had led to the use of simple approximations allowing
at least a qualitative interpretation of the experimental data. The mean-field theory
is widely used to describe the concentration and temperature dependence of Ms and
MAGNETIC AMORPHOUSALLOYS 345

to determine the subnetwork magnetizations in ferrimagnetic alloys (Hasegawa 1975,


Gangulee and Kobliska 1978b, Hansen and Urner-Wille 1979, Mansuripur and
Ruane 1986, Hansen et al. 1989, Hajjar and Mansuripur 1989). The mean-field theory
is based on the Heisenberg model assigning an i atom a magnetic moment,
#i(T) = #i(0)(m~), (33)
where pi(0) = gi#BJi and (mi) = (Ji)/Ji. gi and Ji are, respectively, the gyromagnetic
factor and the total angular momentum of an i atom where the latter is composed
of the spin and orbital angular momenta. (m~) represents the time-averaged pro-
jection of the total angular momentum along the quantization axis and is defined
by
+Ji
miexp(-E,/kBT)
(mi) = m,=+-S,
Ji
(34)
exp(-E,/k.r)
mi = -- J i

where
E~ = - #~(0)(H + H~'~)ch)m,. (35)
ka denotes the Boltzmann constant. In the molecular-field approximation the
exchange field can be written in the form

~o - EJ gigj#B
Hexch(T)- ~ #j(T), (36)

where the Jii represent the exchange constants and the summation in eq. (36) can be
performed leading to the Brillouin function

where

#i(O)" H u~O (38)


Zi = k-~ + **exch).

Assuming all moments of the same atomic species to be equal, the sublattice
magnetization Mi(T) is defined by
mi(T) = Nxi#i(T). (39)
N is the total number of magnetic atoms per unit volume and xi is the fraction of
the ith species in the alloy. The saturation magnetization is given by

Ms(T) = ,=~ M,(T) , (40)

where r is the number of subnetworks.


346 P. HANSEN

Structural disorder causes fluctuations (Handrich 1969, J/iger 1977) of the exchange
interaction. Restricting to nearest-neighbor exchange, eq. (36) can be rewritten as

q) _ 2 ~ n~J(J~J) (1 + Aij)Mj(T). (41)


Hexch(T)- Ngi#a j=l gjPB

i and j now denote the subnetworks and n o are the number of nearest neighbors.
Assuming the exchange fluctuation A~j to be small (A~j~ 1), the Brillouin function
can be expanded in a Taylor series yielding in the absence of an external field for a
one-subnetwork alloy (Handrich 1969, Kaneyoshi 1986),

Ms(T) = ½Ms(O){Bs[z(1 + A)3 + Bs[z(1 - A)]}, (42)


and, for a two-subnetwork (J/iger 1977),

M~ (T) = ¼M~(0){Bj, [2~ M~ (T) + 272M2(T)] + Bs~E21-~M~ (T) + 272M2(T)]


+ Bj~ [2~~M~ (T) + 2i-2M2(T)] + Bj, [2~-~M d T ) + 2~-2M2(T)]},
(43)
where

2i~ - N~g#f3
g ' j (dij)(1 + Aii ). (44)

A corresponding equation applies to M2(T). The average values 3ij are obtained
from (A21A22) =AxlA22.
-2 -2
The mixed terms ( A l l A 2 2 ) vanish because the fluctuations of Jll and Jz2 are
independent. For 3~ = 0, these equations correspond to the molecular-field equations
applied to crystalline ferrimagnets. In this case (J~j--0), the Curie temperature can
be expressed in terms of the 2~j by

Tc = ½{Tll + T22 + x / ( T l l - T22) 2 + 47"12Tz~ }, (45a)


where the T~j are related to the 2o by

Tit = N(Ji + 1)#i(O)2xi


3kB 2ij. (45b)

The relations given in eqs. (42) and (43) were used to calculate the sublattice magne-
tizations regarding the exchange constants as adjustable parameters.
The calculations reveal an increasing flattening of the temperature dependence
of M s with increasing exchange fluctuation parameter 3 (Handrich 1969) and a
pronounced shift of the compensation temperature towards lower temperatures
(Kaneyoshi 1986, 1987). The mean-field theory does not account for spin-wave
excitations at low temperatures and does not account for the exchange fluctuation
in the range of the Curie temperature. The comparison of mean-field calculations
with experimental results will be discussed in section 5.4.2.
M A G N E T I C A M O R P H O U S ALLOYS 347

5.4.1.2. d-band model with random axial anisotropy. The mean-field theory based
on the localized-electron picture as outlined in the preceeding section omits the
itinerant character of the d electrons. A model that accounts for the d-band features
was developed for crystalline compounds (Bloch and Lemaire 1970, Bloch et al. 1975)
and was applied to different amorphous Rl-xTx alloys (Jouve et al. 1976,
Bhattacharjee et al. 1977a,b). This model treats the T atoms in terms of itinerant
d-electron magnetism and the R atoms with localized 4f electrons interacting via the
polarized d band by a local exchange interaction. The corresponding Hamiltonian
can be expressed by

= .~ hjc+cj, + I ~ n]ni ~- Jexcn ~ J , ' a j - D ~, ~j. (46)


t,j i j j

The first two terms on the right-hand side describe the kinetic energy and correlation
in the d band (Hubbard 1963), ci,+ and cj~ are the creation and annihilation operators
for a d-band electron of spin a at site i and j, respectively, ni ~ and ni ~ are, respectively,
the particle number operators for spin-up and spin-down electrons, tij is the d-band
transfer matrix and I is the strength of the on-site Coulomb interaction. These two
quantities represent the properties of the average d band. The third term describes
the exchange interaction between the R and T atoms where Jexch is the exchange
coupling constant and tri is the Pauli spin matrix for the d band. The last term
represents the random anisotropy. Using the molecular-field approximation and the
z-direction parallel to the molecular field, the Hamiltonian of eq. (46) can be written
in the form

= ~, ('~k-- ltTA)e~reka -- 2 (JexchtrzJzJ + D~j), (47)


k j

where

A= I+ ~.

ek is the Fourier transform of t~j, x is the transition metal concentration and A


represents the molecular field of the d electrons. ~z = n T - n * is the polarization of
the d-band electrons that is related to the Fermi function f(e) and the total number
of d electrons N = n T+ n * inside the d band. ¢7~ and N are correlated by the self-
consistent equations,

#~ = ~ [f(~k -- ½A) --f(ek + ½A)], (48a)


k

N = ~ [f(ek -- ½A) +f(ek + ½A)]. (48b)


k
From these relations, the net magnetization
M~ (T) = #B [ x ~ + (1 - x)gJ z ] / V (49)
can be determined assuming the d-band to correspond to a free electron band. This
requires the knowledge of the electron density of states N F at the Fermi level. The
348 P. HANSEN

numerical calculation then can be performed treating I, J0xCh,D and NF as adjustable


parameters.

5.4.1.3. Other concepts. The treatment of a spin cluster in an effective field was
analyzed in terms of the BPW (Bethe-Peierls-Weiss) approximation (Bethe 1935,
Peierls 1936, Weiss 1948, Moorjani and Ghatak 1977). The relevant Hamiltonian is
of the form
N
~'=-2 ~" J o i S o ' S i - g # a H ' S o - g # , H 1 ~., Si. (50)
/=1 /=1

So represents the spin operator interacting with the nearest and next-nearest
neighbors limiting the exchange coupling constants to Jol and Jo2. N is the number
of the atoms in the cluster and H is the applied magnetic field.//1 denotes the sum
of H and an internal field representing the interaction of the spins in the cluster and
with the rest of the medium. This approach applies to the high-temperature range
where the effect of clustering is important for the transition from the magnetically
ordered state to the paramagnetic regime. The Curie temperature is determined by
the condition that the internal field vanishes, leading to the equation
2
[Ll(zl) + L2(z2)] - x(n - 1~' (51)

where L;=cothz~-z; -1 is the Langevin function and zi=2S(S+ 1)Joi/kBT. The


critical concentration for the onset of ferromagnetism also can be calculated from
eq. (51) yielding (Handrich 1972) xc = ( n - 1) -1 (Ising model) and xc = 3 ( n - 1) -1
(Heisenberg model) where n denotes the number of nearest neighbors.
The interaction between two neighboring atoms represents another molecular-
field approach and was discussed in the pair approximation (Smart 1966, Kobe and
Handrich 1972).
The coherent potential approximation (CPA) offers a further approach to calculate
the concentration and temperature dependence of magnetic properties in amorphous
alloys (Elliott et al. 1974, Tahir-Kheli 1976, Foo and Wu 1972). The application of
this concept to disordered alloys involves the introduction of a coherent exchange
interaction. In the weak-coupling limit the CPA results agree with those derived from
the mean-field theory while in the strong-coupling limit the CPA goes beyond the
mean-field theory and is capable to predict critical concentrations. The magnetization
and the Curie temperature usually are calculated using the Green function method.
This leads to the equations
Gij(2 ) = Gij Jv Z Jek [Gij(Z) -- Gkj(Z)], (52)
k
and

1 -mm 1 f +_f [
~N exp (mk__~T)]-1-- 1 Trim(z) dz, (53)

where z = E/m and m = (Sz)/S. From the second equation, the relative magnetization
MAGNETIC AMORPHOUSALLOYS 349

m is determined self-consistently. The coherent exchange interaction .l~k is related via


T-matrices to the exchange coupling constants Jll, Jlz and Jzz as reported for binary
alloys (Foo et al. 1971). The Curie temperature can be determined from a cubic
equation and the critical concentration in CPA is xc = 2/n (Tahir-Kheli 1972a,b,c,
Jones and Yates 1975).
The mean-field theories do not account for local magnetic excitations and thus
cannot provide an accurate description of the low-temperature behavior of the
magnetic properties. This problem can be solved by the spin-wave theory that is
based on the exchange Hamiltonian given in eq. (25). In the quasicrystalline approxi-
mation and the long-wavelength limit, the spin-wave energy can be expressed by
(Keffer 1966)
E k = E o -t- Dk 2 + Fk 4 -I- "", (54)
where k is the wave vector of the spin wave and D and F are the spin-wave stiffness
constants. The presence of spin waves gives rise to a reduction of the average
magnetization, leading to a temperature dependence of the form
Ms(T) = Ms(0)[1 -- B T 3/2 -t- C T 5/2 + -..]. (55)
The coefficients B and C are related to the spin-wave stiffness constant D by

3 g#B
,

and

C _ 3 x l r 2 \ r t 5 ~ ( g#R "~ ( kB y/2

((2a-) = 2.612 and ( ( I ) = 1.341 are the zeta functions and (r 2) represents the average
mean-square range of the exchange interaction. D is directly proportional to the
exchange constants. With increasing exchange strength, the slope of Ms(T) versus
T 3/2 decreases as expected from experimental results of alloys with increasing Tc.

5.4.2. One-subnetwork alloys


5.4.2.1. Transition-metal-based alloys. The temperature dependence and the Curie
temperature reflect the strength of the exchange coupling. The structural disorder in
amorphous alloys induces an exchange fluctuation that causes a pronounced flatten-
ing of the Ms(T) curves. This is demonstrated in fig. 42 showing the reduced hyperfine
fields versus reduced temperature for amorphous and crystalline FeF2 alloys (Litterst
1975). The temperature curve for the amorphous alloy lies substantially below that
for the crystalline compound. This reduction of Ms was explained in terms of the
mean,field theory [eq. (42)] using an exchange fluctuation parameter

A= AJij .~. Jij, (56a)


l,J

which is defined according to the Aij introduced in eq. (41) and adapts values in the
350 P. HANSEN

1.0
ysta|[ine
0.8

0.6 FeF2 e ~

0.4

0.2

I t t I
0.2 0.4 0.6 0.8 1.0
T/Tc
Fig. 42. Normalized hyperfine fields versus reduced temperature for crystalline and amorphous F0.67Fe0.33
(Litterst 1975).

range 0.4 ~<A ~<0.6. A comparison of experimental data with mean-field results is
shown in fig. 43 for Mo.z4Fe0.76 alloys with M = Zr or La (Heiman and Kazama
1979) and Feo.40Nio.4oP0.14B0.o6 (Kaul 1981b) and in fig. 44 for an Po.246Coo.754
alloy (Pan and Turnbull 1974). The experimental data lie typically below the theoreti-
cal curve at low temperatures and above it at high temperatures when the calculated
curve represents an overall fit. This inadequate description of the experimental data
corresponds to the situation for crystalline alloys and has to be attributed to the
mean-field approximations and the temperature-independent treatment of A. In
general, A is a function of temperature (Kaneyoshi and Tamura 1984). Using the
empirical relation

A=Ao 1- , (56b)

the discrepancy between theory and experiment could be removed as it was shown
for amorphous B-FeCo (Prasad et al. 1980) and BSi-CoFe (Bhatnagar et al. 1982,
1984) alloys. The low-temperature data can be described in terms of the spin-wave
theory [eq. (55)] as reported for P - C o (Cargill III and Cochrane 1974, Hfiller and
Dietz 1985) and PB-FeNi (Chien and Hasegawa 1977, Kaul 1981b) and other Fe-
and Co-based alloys (Fernandez-Baca et al. 1987, Liniers et al. 1989). Some results
are reproduced in fig. 45. Deviations from the calculated line occur for T ~>0.2Tc to
0.4Tc due to the neglect of critical fluctuations or the temperature dependence of the
spin-wave stiffness constant. In the corresponding crystalline compounds, deviations
from the T 3/2 law occur already for T >~0.15Tc. However, not all amorphous ferro-
magnets obey the T 3/2 dependence. In the case of Ni-based alloys, the weak itinerant
ferromagnetism leads to a temperature dependence of Ms given by
M2(T) = M2(O)[1 -- (T/Tc)"]. (57)
MAGNETIC AMORPHOUS ALLOYS 351

1.0
&=O
0.5
0.6
0.8 - 0.7
0.8

0.6
E
0./, o Zro.2~ Feo.Ts ~,k~

0.2

I I I I
0.2 0.4 0.6 0.8 1.0
T/Tc

1.0

A=0
0.8
000 0

= 0.6
E

O.A Po.~l,B o.oeFeo,4oNio.~.o ~'l

0.2
(b)
0 i i i i i i i i i
0.2 0.4 0.6 0.8 1.0
T/Tc
Fig. 43. Relativesaturation magnetization m, = M,(T)/Ms(O) versus reduced temperature for (a) amorphous
Zr-Fe and La-Fe alloys (Heiman and Kazama 1979)and (b) an amorphous PB-FeNi alloy (Kaul 198lb).
The full lines represent mean-field calculations [eq. (42)] for different exchange fluctuation parameters.

In the case that spin-wave excitations are neglected, the band theory yields t / = 2.
If collective spin fluctuations are taken into account the theory predicts values for t/
smaller than 2 (Murata and Doniach 1972, Moriya and K a w a b a t a 1973a,b). Amor-
phous Y t - x N i x alloys obey relation (57) with t / = 2 for 0.93 <~x <~ 0.97 while for
x < 0.93 q decreases with decreasing x reaching ~/= 1 for x = 0.833 (Li6nard and
Rebouillat 1978). A corresponding behavior was found for amorphous Cet_xNix
alloys (Fr6my et al. 1984) revealing 1.4 ~<q ~< 1.0 for 0.91 ~<x ~<0.85.
The Curie temperature depends on the transition-metal spin value and the ex-
change constant. The disordered structure causes variations in the nearest-neighbor
distances and the number of nearest neighbors producing local fluctuations of the
magnetic m o m e n t and the exchange coupling. Random fluctuations of the exchange
352 P. HANSEN

JLa=/~=0

0.5
0.8

0.6
Y
0.4 i

0.2

I I I I I I I I I I ~'
O0 0.2 0.4 0.6 0.8 1.0
T/Tc
Fig. 44. Relative saturation magnetization ms = a,(T)/as(O) versus reduced temperature for an amorphous
P-Co alloy (Pan and Turnbull 1974). The full lines represent mean-field calculations [eq. (42)] for different
exchange fluctuation parameters.

coupling constant leads to an increase of Tc provided the ferromagnetic coupling is


maintained. This applies to Co-based alloys which are strong ferromagnets and
exhibit larger Tc values as compared to the corresponding crystalline counterparts
as demonstrated in fig. 46a for amorphous and crystalline Y-Co alloys (Fukamichi
et al. 1987a). The Tc values for x > 0.70 represent estimated data because Tc interferes
with the crystallization temperatures. The opposite situation occurs for Fe-based
alloys where both the moment and the average exchange coupling constant either
are equal or smaller than in the crystalline compounds and thus a corresponding
behavior is observed for the Curie temperatures as shown in fig. 46b for amorphous
and crystalline B-Fe alloys (Stobiecki and Stobiecki 1983). This can be attributed
to the sensitive dependence of the Fe-Fe exchange on the
atomic distances that can even change sign for distances below 0.254nm. The
competing positive and negative exchange constants are also the reason for
the turndown of Tc in the limit of Fe-rich alloys as shown in figs. 46b and 47a. In
the case of amorphous M - F e alloys with M = Zr, Hf, Ce, La or Lu, the magnetic
phase diagrams (see fig. 30 for Ce-Fe) indicate a transition to an asperomagnetic or
spin-glass-like state (Fukamichi et al. 1989a, Kakehashi 1990a, Krey et al. 1990).
However, many other Fe-based alloys behave similar to the Co-based alloys and
exhibit a monotonous rapid fall of Tc with increasing solute concentration as, e.g.,
observed for amorphous Ti-Fe alloys (Liou and Chien 1984, Xiao and Chien 1987).
The variation of Tc for amorphous FeNi alloys (Kaul 1981a) is presented in fig. 48.
The full lines were calculated in terms of the coherent potential approximation.
Variations in the metalloid composition produces changes in Tc from 500 to 700 K
for Fe-rich alloys (Luborsky 1978, 1980). Amorphous alloys containing Co exhibit
significantly higher Tc values due to the stronger Fe-Co exchange as compared to
MAGNETIC AMORPHOUS ALLOYS 353

1.000

~ x=0.81
0.975 / x=0.78

x=0'76/ /~

(o) \
0.950 i i t i
0 4.00 800 1200 1600 1800
T~(K~)

1.00

0.95 ~ o a 4 B o . o s Feo.l,oNio.~.o

o.9o
0.85

0.80
(b)
0.75 = i i i
0.0 0.1 0.2 0.3 0.4 0,5
(T/Tc )3'2
Fig. 45. The change of relative saturation magnetization ms = M,(T)/M,(O) versus reduced temperature
(a) for amorphous P1 _=Cox alloys (Cargill III and Cochrane 1974) and (b) for an amorphous P B - F e N i
alloy (Kaul 1981b). The full lines were calculated from the spin-wave theory [eq. (55)].

the Fe-Fe exchange. Curie temperatures and low-temperature magnetic moments


for some transition-metal one-subnetwork alloys are compiled in table 10. The critical
concentration Xc decreases in 3d based alloys in the sequence Ni, Co, Fe from roughly
0.8 to 0.4, in agreement with magnetic moment and hyperfine-field measurements.
The xc values reflect the different number of nearest TM neighbors necessary to
establish magnetic order according to the environment model. Critical concentrations
are collected in table 6 for various amorphous alloys. It should be noted that the
354 P. HANSEN

1500
j
o! /
1000 /'/

amorpho

500- crysta[tine

0.50 0.60 0.70 0.80 0.90 1.00

-Fel
BFe2 ~,=
1000

,,~BFe 3
800

+ 0 x

~ 600 BFe~" ~x
o +x
~+
0 O 0°
400
+ Bl-x Fex

200
(b) +
0
0 1 ++1 I I
0.20 0.40 0.60 0.80 1.00
X

Fig. 46. Concentration dependence of the Curie temperature for (a) amorphous (Fukamichi et al. 1987a)
and crystalline (Buschow 1980c) Y-Co alloys and (b) for amorphous B-Fe films [(O) magnetization
(Stobiecki and Stobiecki 1983), (O) anomalous Hall effect (Stobiecki 1982), (4-) M6ssbauer (Chien and
Unruh 1981), (x) ribbons (Hasegawa and Ray 1978), (11) crystalline B-Fe compounds (Vincze et al.
1979)].

pronounced sensitivity of the exchange interaction on the structural disorder gives


rise to a dependence of Tc on the preparation conditions. In particular for liquid-
and vapor-quenched amorphous alloys large Tc differences were observed as shown
in fig. 23b.

5.4.2.2. Rare-earth-based alloys. The temperature dependence of the magnetization


of amorphous alloys with the rare earths as the only magnetic component are
MAGNETIC AMORPHOUS ALLOYS 355

350
MI_x Fex
300
M=Sc
25£

200

150

100

5O
(a)
0
0.40 0.60 0.80 1,00
X

30O

Til-x Fe x
200

10[

{b)
I I
0.20 0./.,0 0.60 0.80 1.00
X

Fig. 47. Concentration dependence of the Curie temperature for (a) amorphous M1 -xFex alloys with M =
Sc, Hf or Zr (Fukamichi et al. 1986a) and (b) for amorphous [(0) Liou and Chien (1984)] and crystalline
[([]) Fukamichi et al. (1982)] Ti-Fe alloys.

controlled by the RKKY exchange interaction and the random anisotropy. In Gd-
based alloys, the latter is small and the temperature behavior is governed by the
relatively weak RKKY interaction. The structural disorder gives rise to the character-
istic flattening of the Ms curves as already discussed for the TM-based alloys. A
typical example is shown in fig. 49a for amorphous Gd-Au (Durand and Poon 1977).
The full lines represent the mean-field theory for different values of the exchange
fluctuation parameter [eq. (56a)]. The best fit was obtained for A = 0.4 and a spin
value of S = 7 for the Gd atom. Thus, the magnitude of A corresponds to those found
for alloys containing TM elements. The low-temperature magnetization obeys the
T 3/2 law derived from the spin-wave theory I-eq. (55)] as shown in fig. 49b (Durand
356 P. HANSEN

800

700

600

500

400

3OO

200

100

0
0
//V•
.Zt T °"
0.25
Bo.20(Fey Nil_y )o.eo
Bo.19 Si o.01(Fey Ni 1-y )0.80

0.50 0.75 1,00


Y
Fig. 48. Concentration dependence of the Curie temperature for amorphous B-FeNi [(~7) Kaul (1981a),
(!?) Becker et al. (1977)], BSi-FeNi [(A) Kaul (1981a)] and BP-FeNi [((3) Kaul (1981a), (@) Becker
et al. (1977), (D) Chien et al. (1977), (111)Krause et al. (1980)] alloys. The full lines were calculated from
the coherent potential approximation (CPA).

and Poon 1977). However, the agreement of the exchange values evaluated from the
spin-wave theory and the Curie temperature is not satisfactory.
The Curie temperature of Gdl _xMx alloys is shown in fig. 50, displaying a strong
increase of Tc with increasing x (Heiman and Kazama 1978b). The full line was
calculated from a statistical nearest-neighbor model yielding (Oguchi 1971)
[ - I / -' ] X~ n - 1

where n is the coordination number. If n is taken to be 12, the critical concentration


xc = 4/n is found to be ½, in good agreement with the experimental results.
The RKKY interaction is expected to be sensitive to fluctuations in nearest-
neighbor distances. This tends to reduce Tc in amorphous alloys as compared to the
crystalline counterparts as observed for various compositions. A few examples are
listed in table 8. The numbers in the brackets are the Curie temperature of the
corresponding crystalline compounds. Tc values and low-temperature magnetic
moments for some R - M alloys are also listed in table 8. It should be noted that the
random anisotropy for non-S-state rare earths is rather high and thus very high
magnetic fields are required for magnetic saturation. Discrepancies in magnetic
moments, therefore, in many cases can be ascribed to insufficient saturation.

5.4.3. Two-subnetwork alloys


The temperature dependence and the Curie temperature in amorphous R1 -xTx alloys
is determined by the spin and angular momentum of the R and T atoms and their
MAGNETIC AMORPHOUS ALLOYS 357

TABLE 10
Transition-metal moment at T= 4.2K and Curie temperature for some amorphous
one-subnetwork alloys.
MI-~Tx kit (#B) Tc(K) References
Bo.zoFeo.8o 1.55 651 Becker et al. (1977)
Po.125Co.o75Feo.so 1.68 586 Tsuei et al. (1968)
Sco.25Feo.75 1.13 277 Fukamichi et al. (1986a)
Tio.zsFeo.75 0.65 238
Zro.a5Feo.T5 0.95 283
Yo.52Feo.48 0.86 18 Coey et al. (1981)
Y0.32Feo.68 1.69 70
Yo.2oFeo.so 1.89 108
Lao.24Feo.76 1.53 330 Kazama et al. (1980)
Ceo.2oFeo.8o 0.80 190 Buschow and van Engen (1980)
Geo.53Feo.47 0.90 230 Suran et al. (1976)
Bo.2oCoo.so 1.02 ~765 O'Handley et al. (1976a)
Po.24COo.76 0.88 512 Pan and Turnbull (1974)
Yo.sTCoo.43 0.20 450 Heiman and Lee (1975)
Yo.4oCoo.6o 0.60 195 Fukamichi et al. (1986b)
Yo.2oCoo.8o 1.65 > 600 Buschow et al. (1977)
Yo.2~Nio.75 0.03 20 Li6nard and Rebouillat (1978)
Yo.13Nio.s7 0.12 118
Yo.o~Nio.95 0.36 327
Yo.oaNio.97 0.41 390
Co.ogNio.91 0.21 164 Fr6my et al. (1984)
Co.lsNio.85 0.034 11
Bo.2oFeo.4oNio.4o 1.03 662 O'Handley et al. (1976a)
Bo.2oFeo.40Coo.4o 1.43 > 800

exchange coupling parameters JR--R, JR--T and JT-T. The direct exchange between
the 3d electrons of the T atoms leads to a ferromagnetic coupling. The much weaker
R K K Y interaction between the rare-earth atoms also gives rise to a parallel alignment
of their moments, but JR-R ~ JT--T" The exchange between the 3d transition-metal
electrons and the 5d rare-earth electrons induce a negative JR-T producing a parallel
alignment of T and R moments for light rare earths and an antiparaUel alignment
for heavy rare earths. Collinear structures are expected at low fields only for alloys
containing S-state rare earths and for x values sufficiently above the critical concen-
tration. Alloys exhibiting speromagnetic or sperimagnetic order (see figs. 20 and 21)
require very high magnetic fields to reach magnetic saturation due to the random
anisotropy discussed in section 5.3.1.2. Both the transition-metal spin value and the
exchange coupling constants are sensitive functions on composition.
Most attention has been focussed on evaporated and sputtered (Gd, Tb)l _~Fex
and T b l _~(Fe, Co)~ alloys with 0.6 < x < 0.8. They are suitable candidates for mag-
neto-optical recording (see section 8.1). The temperature dependence of amorphous
G d t _~F% alloys (Hansen et al. 1989) is shown in fig. 51. The strong variation of the
low-temperature magnetization with x is associated with ferrimagnetic order (see
also fig. 38a) and leads to a variety of different magnetization curves. The appearance
of magnetic compensation (Ms = 0) is limited to a very narrow range of compositions.
358 P. HANSEN

~,,
A=O
0.8 0.4
0.6

0.6

0.4 • Gdo.eoAuo.2o
0.2 • crystalLine ~

n la) I t t I
0.2 0.& 0.6 0.8 1.0

[ 20K
215 ~ o Gdo.6s Nio.32
195 ~ • Gdo.8o Auo.2o

~£~ 155

135

115
{b} •
g5 i t t
500 1000 1500
T3/2 (K 3/2)

Fig. 49. (a) Relative saturation magnetization ms = Ms(T)/Ms(4.2 K) versus reduced temperature for amor-
phous Gd-Au and (b) temperature dependence of Ms versus temperature for amorphous Gd-Au and
Gd-Ni alloys (Durand and Pooh 1977). The arrows indicate deviations from the T 3/2 law.

The full lines were calculated from the mean-field theory using eq. (43) for A = 0. The
good agreement between experimental and calculated results was confirmed for
various other Gd-Fe-based and R-Fe-based alloys (Heiman et al. 1976b, Taylor and
Gangulee 1976, Gangulee and Taylor 1978, Mimura et al. 1978, Hansen and Urner-
Wille 1979, Hartmann et al. 1984b, Hansen and Hartmann 1986, Mansuripur and
Ruane 1986, Hansen et al. 1989), but it should be noticed that now three exchange
parameters were used to adjust the theory to the experimental data when compared
to one-subnetwork alloys. The better fit thus obtained is no indication for a higher
accuracy of the extracted values for the exchange constants. They should only be
regarded as empirical parameters. However, they are well suited to calculate the
sublattice magnetizations and to model the temperature dependence of other mag-
netic properties which can be expressed in terms of the sublattice magnetizations.
MAGNETIC AMORPHOUS ALLOYS 359

300
crystalline Gd

200
Gdx tl_x /

H=A~ ;rNi

Cu ~ AI
100

0
Nio/
0 0.20 0.40 0.60 0.80 1.00
X

Fig. 50. Concentration dependence of the Curie temperature for amorphous Gd-M alloys with M = A1,
Au, Cu or Ni. (O) Heiman and Kazama (1978b); (O) Lee and Heiman (1975), Boucher (1977), Durand
and Poon (1977), Mizoguchi et al. (1977a). The full line represents a statistical nearest-neighbor theory
[eq. (57)].

150C
. ~ Gdl-xFex
--calculated
" ~x=0.37
. ~ ~ 0 . 4 7
I00C - •

0.80

100 200 300 400 500


T(K)
Fig. 51. Temperature dependence of the saturation magnetization for amorphous Gd-Fe alloys prepared
by evaporation (Hansen et al. 1989). The full lines were calculated from the mean-field theory I-eq.(43),
with Z = 0].
360 P. HANSEN

Expressions for the compositional dependence of the exchange constants Jik extracted
from the mean-field analysis are compiled in table 11 for some R - T alloys. Their
validity is restricted to a limited concentration range. A plot of the sublattice
magnetizations versus temperature is given in fig. 52 for a composition exhibiting a
compensation temperature T~ompat 295 K. These curves are based on experimental
data and correspond to one of the curves shown in fig. 51.
The temperature variation of the spontaneous moment per formula unit for crystal-
line and amorphous TbFe2 (Rhyne et al. 1974b) is presented in fig. 53. It demonstrates
the strong reduction in moment and Curie temperature in the amorphous state that
applies to all R-Fe alloys. Part of the moment reduction can be attributed to the
asperomagnetic order of both sublattices as indicated by the arrows. This problem
arises for all low-field magnetization data of R-Fe alloys containing non-S-state rare
earths (Taylor et al. 1978, Croat 1981a,b, Mansuripur and Ruane 1986, Hansen and
Witter 1988).
Amorphous R-Co alloys are characterized by significantly higher exchange inter-
actions inducing a collinear Co sublattice and producing much higher Tc values for
Co-rich alloys. The temperature dependence of Ms is shown in figs. 54, 55 and 56
for some R-Co alloys (Jouve et al. 1976, Honda and Yoshiyama 1988a, Hansen et
al. 1989). The large Tc for alloys with x > 0.7 interferes with the crystallization
temperatures which prevent Ms and Tc measurements for T > T~. The full lines in
figs. 54 and 55 were obtained from mean-field calculations based on eq. (43) with
A = 0. Choosing A to be of the order of 0.5 as found for the one-subnetwork alloys,
the exchange coupling constants used to fit the experimental data have to be slightly
modified to account for the shift in Teompinduced by A. The dependence of Jik o n x
used for the theoretical lines shown in fig. 55 are given in table 11. The mean field
analysis reveals a strong rise of JCo-Cofor x > xo in contrast to JFe-Fe passing through
a maximum around x ~ 0.5 due to the increasing portion of negative exchange for
Fe-rich R - F e alloys. It should be noticed that in the case of N d - T alloys (fig. 55c)
no T~ompoccurs due to the parallel alignment of the sublattice magnetizations which
holds for all alloys containing light rare earths (Dai et al. 1986, Takahashi et al. 1987,
Yang and Miyazaki 1988, Yang et al. 1988). A model that accounts for the itinerant
d-electron ferromagnetism of the Co sublattice was described in section 5.4.1.2. It
was used to calculate the temperature variation o f Ms for different R - C o alloys
(Jouve et al. 1976, Bhattacharjee et al. !977b). The comparison of experimental and
calculated results is shown in fig. 56. In the range of the broken lines the formation
of crystalline phases takes place. The theory was treated in terms of the molecular-
field approximation. Thus, both mean-field models based on the localized and d-band
approach account for the measured temperature variation of Ms. A comparison of
the extracted parameters with data obtained from independent measurements will
decide which of the two models represent the better approximation. Generally, the
comparison between theoretical and experimental results suffer from the lack of data
in the high-temperature regime for Co-rich alloys. Various other mean-field results
were reported for G d - C o based and R - C o alloys (Hasegawa 1975, Hasegawa et al.
1975a,b, Taylor and Gangulee 1976, Roberts et al. 1977, Gangulee and Kobliska
1978a,b, Honda and Yoshiyama 1988a,b).
MAGNETIC AMORPHOUS ALLOYS 361

~ .... E E

A~

c~
C~

r__, ~
i.~ r~

~"O
"O

r"-i r-'l

] I
I I I I

t ~~ ~ooo
.

~V
--~ VWW~
~
v~

I
"O t_.~ t . . ~ L . . I
362 P. H A N S E N

1000

800
~
..... ~.~
6d

600
~-~ Gdo.25s Fe 0.7t.s %
400
Z

x 200

oF , . IT'°: ° , . 3
0 100 200 300 400 500
T(K)
Fig. 52. Temperature dependence of the saturation magnetization and the sublattice magnetization for an
amorphous Gd-Fe alloy exhibiting a compensation temperature. The Ms curve corresponds to that
shown in fig. 51 for x = 0.745.

i ~ bFe2

staHine

0 i I I b i i i~
0 100 200 300 400 500 600 700 800
T(K)
Fig. 53. Temperature dependence of the spontaneous moment per formula unit for crystalline and
amorphous TbFe2 (Rhyne et al. 1974b).

The room-temperature concentration dependence of Ms is shown in fig. 57 for


some amorphous R-Fe and R-Co alloys prepared by evaporation (Orehotsky and
Schr6der 1972, Roberts et al. 1977, Hansen et al. 1989). The full lines in fig. 57b
represent mean-field theory results. Ms = 0 at low transition-metal concentrations
refer to alloys with Tc equal to room temperature. The corresponding x value is
higher for Co-based alloys because their x c value is higher as compared to the Fe-
based alloys. The magnitude of Ms at the maximum is determined by the sublattice
moments and the exchange coupling constants and reflects the position of xo and
MAGNETIC AMORPHOUS ALLOYS 363

1500 ~q.
Gdl-x Cox
a~=0.436 - - catcutctted

1000 0.538
~k

500 \ ._.

o.839~ ( p ~ 4 ~ ~-

O~- --w''''-"''~
100 200
' ' ~ 3'00 400
' UlO
50
T{K)
Fig. 54. Temperature dependence of the saturation magnetization for amorphous Od-Co alloys prepared
by evaporation (Hansen et al. 1989). The full lines were calculated from the mean-field theory [eq. (43),
with zT= 0].

Tc. The magnetic compensation is defined by


MR(x¢omp, T~omp)-- Mx(x¢omp, T~omp)= 0, (59)
where X¢omp and T~omp are the compensation composition and the compensation
temperature, respectively. For amorphous Gd-Fe, e.g., Xcomp= 0.77, Tcomp = 4.2 K
and Xcomp= 0.745, T~omo= 295 K were found. The room temperature T~omp values for
most alloys prepared under the same conditions appear in range 0.7 < X~omp< 0.8
(fig. 57a,b) except for Er-T alloys exhibiting Toompvalues below room temperature
(Dirks et al. 1977). The variation of T~ompwith composition is presented in fig. 58 for
some binary and ternary amorphous alloys (Hansen et al. 1989). The plots yield a
Tcomp shift of 40K/at.% for T b - T alloys and 100K/at.% for G d - F e alloys which
demonstrates the high sensitivity of TcompOn composition. The compositional varia-
tion of T~ompfor Dy- and Ho-based alloys yields smaller shifts of the order of 20 K/
at.% (Hansen et al. 1991). Therefore, any changes on the R or T moment and the
exchange interaction by small additions of nonmagnetic atoms, the presence of
impurities, oxidation effects, thermal treatments or structural changes lead to drastic
variations in Tcomp(Katayama et al. 1977, Mfiller et al. 1977, Biesterbos et al. 1979,
Tsunashima et al. 1980, Schelleng et al. 1984, Heitmann et al. 1987a,b). Compensation
temperatures for some amorphous alloys are listed in table 12.
The reduced Fe moment and the sensitivity of the iron exchange on the structural
disorder produces lower Tc in amorphous R-Fe alloys than in the corresponding
crystalline phases. This is obvious from the Tc data given in table 12 and those for
crystalline (Buschow 1977) and amorphous (Heiman et al. 1976b) RFe2 alloys which
are presented in fig. 59a showing the trends across the series. The downwards trend
364 P. HANSEN

1400

1200 Tbl-xCox
--calculated
I000
x=O.g2
< 800

6OO
0,85
400

200

100 300 500 700


T(K)
1400]

?oOoOoE Dyl-xCox

~
--calculated

800 =0.92
.<
600
E
z~ 0.82
400

200

} I ~t*fr l I~l'~'r- 7 I "% I I


0
100 300 500 700 900
T(K)
1400
Ndl_× Cox
1200 ~ - - . ~ o - - catculctted

1000
x=O.g2

-~ 800
<
~--- 600 ~ 0.78
400
~ 0 . 6 4
200
(c)

( 100 300 500 700 go0

T(K)
Fig. 55. Temperature dependence of the saturation magnetization for amorphous (a) Tb-Co, (b) Dy-Co
and (c) Nd-Co alloys prepared by diode sputtering (Honda and Yoshiyama 1988a). The full lines were
calculated from the mean-field theory [eq. (43) with z/= 0].
MAGNETIC AMORPHOUS ALLOYS 365

E r 0.225 C00,77s

3
.765

:&
v 2
:&
//-Dyo,,, Coo,,, \\\

I
2O0 &O0 600 800 1000
T(K)

Fig. 56. Temperature dependence of the average spontaneous magnetic moment per formula unit for
amorphous R-Co alloys (Jouve et al. 1976). The full lines were calculated from the d-band model [eqs. (48)
and (49)]. The Curie temperature was estimated from partially crystallized samples.

of Tc towards La and Lu is associated with the increasing portion of negative Fe-


Fe interaction and a decrease in average Fe moment (Buschow and van der Kraan
1981). The crosses represent the Tc values of the pure crystalline rare-earth elements
revealing almost the same turndown of Tc which indicates the importance of the R-
R exchange for low and medium Fe concentrations. In this case, Tc is expected to
vary according to the de Gennes factor (g - 1)2J(J + 1) listed in table 6. This variation
indeed was verified for these alloys (Rhyne 1976) and for amorphous R1 _~Co~ alloys
with x=0.31 (Buschow 1980b), x=0.40, 0.50 (Yang et al. 1990) and RI_~Ni~
(Buschow 1980a) alloys with x = 0.31. In these cases, the T concentration is below
xc and thus the Co or Ni atoms carry no magnetic moment. Therefore, the Co and
Ni atoms just dilute the R matrix and Tc is determined only by the R-R exchange
interaction that can be expressed in terms of the de Gennes factor. This variation is
shown in fig. 59b for amorphous Ro.69Co0.31 alloys (Buschow 1980b).
The amorphous R~ _~Co~ alloys with x > xo are characterized by a strong cobalt
subnetwork exchange that causes a collinear Co subnetwork and gives rise to large
Curie temperatures. These are larger than the Tc values of the crystalline counterparts
(see table 12) and for x > 0.7 even exceed the crystallization temperatures in contrast
to the R-Fe alloys. This difference in Fe- and Co-based alloys is demonstrated by
the plot of Tc versus x shown in fig. 60 for amorphous R-Fe and R-Co alloys with
R = Gd, Tb, Dy or Ho (Buschow and van der Kraan 1981, Hansen et al. 1989, 1991)
and in fig. 61 for amorphous R-Co alloys (Takahashi et al. 1988a,b). All R-Fe alloys
exhibit a maximum of Tc around x --- 0.7 followed by a strong turndown. However,
the compositional range 0.9 ~<x ~< 1.00 has not been investigated in detail for R-Fe
alloys and the existence of a tricritical point has not yet been documented. The initial
decrease of Tc for Gd-rich Gd-Co-based alloys with increasing x in fig. 60 reflects
the nonmagnetic behavior of the Co atoms and, therefore, the Tc variation agrees
366 P. HANSEN

1000

< Gdl-× Fex Gdl-×Cox|11


500
z

{a)

0 0.4 0.6 0.8
X

500
Rl-xC°x / /
,oo

Dy
~300
<

~2oo

100

0
0,55 0.60 0.65 0.70 0.75 0.80 0,85 0.90
X

Fig. 57. Compositional variation of the room-temperature saturation magnetization for amorphous
R1 -=Tx films prepared by evaporation: (a) R = Gd or Tb, T = Fe or Co (Hansen et al. 1989) and (b) R =
Ho or Dy, T = Co (Roberts et al. 1977). The full symbols in (a) refer to Ms data taken from different
investigators [Gd-Fe: Taylor (1976), Mimura et al. (1978), Lee et al. (1986); Gd-Co: Taylor and Gangulee
(1976); Tb-Co (dotted line): Choe et al. (1987)].

with t h a t of diluted o n e - s u b n e t w o r k r a r e - e a r t h alloys as s h o w n b y the chain line


representing a m o r p h o u s G d - C u alloys ( H e i m a n a n d K a z a m a , 1978b). F o r x n e a r
xc, the increasing R - C o exchange leads to a n increase of Tc a n d for x > x , the s t r o n g
C o - C o exchange causes the d r a m a t i c rise of Tc. F i g u r e 61 shows the c o m p o s i t i o n a l
v a r i a t i o n of Tc for different R - C o alloys revealing Tc to be a l m o s t i n d e p e n d e n t of
M A G N E T I C A M O R P H O U S ALLOYS 367

6OO

500 -- -- I : - - ~ -- ~ ..Zc ........~.

400 ~Gd1-xFex~dl-xC°x

300
I'-° .\ --.Tcomp/ \
200 ~ i m

tO0
. \
0 ~
0.70 0,75 X 0.80 0.85
500

......-~. . . . . ~ . ~ . . , ~ T b1-×Fex
400 Tc \o\" -"~"
co.

300
Tb1-x(Fe 1-
&
×y,o.21, \ ~...Tb1_xMX
200 o,. Foe

100 o ~

[b) I I
(~6 0.7 0.8 0.9
X
Fig. 58. Compositional variation of the compensation temperature for amorphous R 1 -xTx alloys prepared
by evaporation (Hansen et al. 1989). (a) R = Gd; T = Fe or Co and (b) R = Tb or GdTb, T = Fe, Co or
FeCo. The full symbols in (a) and (b) refer to T~omp data taken from different investigators [Gd-Fe:
Heiman et al. (1976b), Taylor (1976), Mimura et al. (1978), Biesterbos et al. (1979); G d - C o : Tao et al.
(1974), Taylor and Gangulee (1976), Biesterbos et al. (1979); Tb-Fe: Mimura et al. (1976a,b, 1978),
Takayama et al. (1987); Tb-Co: Biesterbos et al. (1979), Heitmann et al. (1985), Choe et al. (1987)]. The
dotted line in (b) represents sputtered G d T b - C o films (Kryder et al. t987).
368 P. HANSEN

TABLE 12
Curie temperatures Tc and compensation temperatures Teompof crystalline and
amorphous R1 _~T~ alloys with T = Fe, Co or Ni. + denotes no compensation, the
magnetization is dominated by the indicated element (Lee and Heiman 1975, Arrese-
Boggiano et al. 1976, Alperin et al. 1976, Heiman et al. 1976a,b, Rhyne 1976).
R~ _~T~ Tc (K) T=omp(K)
Crystalline Amorphous Crystalline Amorphous
Gdo.s7 Feo.43 350 Gd +
Gdo.4oFeo.6o > 500 Gd +
GdFe z 785 490 Gd + 450
GdFe 3 728 460 150
Gd6 Fe23 659 420 ~ 100
GdCo 2 409 550 Gd + 510
GdCo 3 612 750 Gd + 400
GdzCo 7 775 > 500 410 300
GdCo 5 1008 > 500 80
GdNi z 85 38 Gd + Gd +
TbFe2 711 390 Tb + Tb +
TbFe3 648 405 Tb + Tb +
Tb6 Fe23 547 387 Tb + Tb +
Tb2 Felt 409 365
TbCoz 256 > 600 Tb + 500
TbCo3 506 > 600 Tb + 250
TbCos 980 > 600 100 Co +
DyFe2 638 287 Dy + Dy +
DyFe3 600 333, 350 Dy + Dy +
Dy 6 Fe23 524 351
DyCo 3 450 > 900 230
DyNi3 69 47
Hoo.4o Feo.60 250 180
HoFe z 612 260 Ho + 120
HoFe 3 567 290 400 50
Ho 6Fe23 501 300 40 Fe +
Hoo.45 Coo.ss 375 Ho +
Hoo .40Coo .60 600 325
HoCo/ 85 > 600 Ho + 270
HoCo3 418 > 600 350 150
HoCo5 1000 > 600 80 Co +
Ho2Ni17 162 > 400
HoNi2 22 15
HoNi5 10 400
Ho6Mn23 434 Not magn.
ErFe2 575 105 Er + Er +
TmFe2 565 < 50 Tm+ Tm +

the rare earth except for Er-Co alloys. This can be attributed to the dominance of
the Co-Co exchange for medium and high Co content. It should be noticed that the
magnitude of Tc strongly depends on the preparation conditions affecting the struc-
tural disorder and thus the exchange coupling constants. In particular strong differ-
ences in Tc are observed for vapor-quenched and liquid-quenched alloys (Miyazaki
et al. 1987a, Takahashi et al. 1988a,b), shown in fig. 23.
MAGNETIC AMORPHOUS ALLOYS 369

1000

RFe 2
800

600

400 . ~ - orphous

200

a)
0 i ~ P t t t t i t ~ ~ ~ t
La Ce Pr Nd P m S m Eu Gd Tb Dy Ho Er Tm Yb Lu
R

200
Ro.69 C 00.31

150

100

50 0 0

De-
/ a Ce Pr N d P m S m Eu Gd Tb Dy Ho Er Tm Yb l u
R

Fig. 59. Curie temperature for (a) crystalline(Buschow 1977)and amorphous (Heimanet al. 1976b)RFe2
alloys and (b) amorphous Ro.69Co0.31 alloys (Buschow 1980b).

5.5. Critical exponents

Magnetically ordered systems undergo a second-order phase transition when passing


the critical temperature. The magnetization represents the order parameter that is
controlled by the reduced temperature e = (T/Tc) - 1 and the external magnetic field.
The critical phenomena are discussed in terms of the static hypothesis (Domb and
Hunter 1965, Widom 1965, Wilson 1974a,b), which can be expressed in the general
form

I~lp +,
where the signs + and - refer to T < Tc and T > To respectively. # and 6 are the
critical exponents which refer respectively to the temperature dependence of M below
370 P. HANSEN

~0 0.2 0.4 0.6 0.8 1.0


x
Fig. 60. Compositional variation of the Curie temperature for amorphous R-T alloys prepared by evapora-
tion (Hansen et al. 1989). The full symbols refer to Tc data taken from different investigators [Tb-Fe:
Alperin et al. (1976), Heiman et al. (1976b), Busehow and van der Kraan (1981), Takayama et al. (1987);
Tb-Co: Lee and Heiman (1975), Buschow (1980a), Busehow et al. (1980); Gd-Cu: Heiman and Kazama
(1978b)]. The full lines for Dy-Fe and Ho-Fe were taken from Hansen et al. (1991).

1000

Rl-x COx /m
oR=Sm Id
• p~ % ' f

• Er
500 []z, Nd
D y e /

i i i i
O0 0.20 0.40 0.60 0.80 1.00
x
Fig. 61. Compositional variation of the Curie temperature for amorphous R-Co alloys prepared by
evaporation (Takahishi et al. 1988a,b).
MAGNETIC AMORPHOUS ALLOYS 371

Tc and the field variation of M at Tc according to the relations

M ~ ( - e ) p ( H = 0 , T < Tc), (61a)


M~H 1/~ ( T = To, H--*0). (61b)
The critical exponent 7 associated with the susceptibility above Tc is defined by the
expression

X-1 ~ (T>~ Tc). (61c)


The critical exponent ~ is associated with the specific heat and is defined by

Cn ~ e-" (T ~> rc). (62)


The static hypothesis, eq. (60), leads to the equalities

7 =/~(6 - 1), (63a)


= 2(1 -/~) - 7. 63b)

A comparison of the micromagnetic theory of phase transitions (Kronmfiller and


F~ihnle 1980, F~ihnle and Kronmiiller 1980, F~hnle 1980, Herzer et al. 1980, 1981,
Meyer and Kronmfiller 1982) with experimental data reveal that exchange fluctua-
tions are the primary origin of the differences of the critical exponents between
amorphous and crystalline materials.
The examination of the critical behavior requires materials which possess values
of Tc well below their crystallization temperatures. The critical exponents were
determined from different representations such as Kouvel-Fisher plots, modified
Arrott plots or scaling plots (Reisser et al. 1988), leading to slightly different results.
Typical values of the critical exponents for some crystalline metals (Fe, Ni, Co, Gd)
and amorphous alloys are compiled in table 13. The critical exponents predicted by
the three-dimensional Heisenberg model are also listed in table 13. They are in
reasonable agreement with the experimental data of many amorphous alloys. Also,
the scaling laws are satisfied for many alloys. This can be ascribed to the long-
wavelength dependence of the critical fluctuations near Tc and, thus, the critical
behavior is independent of the local atomic structure. This is confirmed by theoretical
investigations (Harris 1974) predicting no influence of the structural disorder on the
critical exponents. However, outside the critical region strong differences between
amorphous and crystalline materials occur. This is obvious from the temperature
dependence of the susceptibility and the effective exponent y(T) defined by (Kouvel
and Fisher 1964)

dlnx -~
(64)
7(T)- dine "

For amorphous alloys 7(T) typically shows a maximum in the temperature depen-
dence as shown in fig. 62 for amorphous Zr-Fe (Reisser et al. 1988) in contrast to
crystalline compounds where 7(T) decreases monotonically with temperature (F/ihnle
372 P. HANSEN

izl
_,.- ~ o
"-r. ~
o
"-r. -,=,
~ ~.~
0 r~
"a 8

,~ ~ ~ ~ .~ o
o ~ o

~f
o o ~ o o o o o o o o o o o
I I + I + I I I I I I I I I

o
+1 +1 +1 ~ +1 +1 +1 +1 +1
,,-,1

S
o'-'a
~t
4-1 +1 4-1 +1 +1 r.-- 4-1 +1 4-1 ,,~

~ ~1 ~ ~ 0 ~ ~1 ~ ~1
I~ +1 +1+1~ +1+1 +1 ~

,-fi
°

8
~1=o "I. c~ d "I, "I, ,-~
MAGNETICAMORPHOUSALLOYS 373

1.70

0 0
0 0

1.60 0 0

0 0
o o
o
1,50 0 o

/ 0
0
0
I.&0

Z ro.lO F eo.@o

1.30 i i i
200 250 300 350 &00
T(K]
Fig. 62. Temperature dependenceof the effectiveexponent7(T) for amorphous Fe-Zr with Tc= 207.5K
(Reisser et al. 1988).

et al. 1983). A similar behavior for 7(T) was reported for disordered Po.25Feo.75 and
Pdo.75Feo.zs alloys satisfying also the scaling laws (Seeger and Kronmiiller 1989).

5.6. Uniaxial magnetic anisotropy

Amorphous magnetic alloys are expected to behave magnetically isotropic or to


exhibit a low anisotropy in the case of thin films where the deposition on a substrate
induces intrinsic stresses (d'Heude and Harper 1989) which may cause a stress-
induced anisotropy. However, for most alloys a uniaxial anisotropy energy,
E, = Ku sin 2 0, (65)
was found with values of the uniaxial anisotropy constant, Ku, ranging between
some hundred J m -a for field-induced anisotropies in metal-metalloid alloys and
some hundred kJ m-3 for rare-earth-transition-metal films. 0 represents the angle
between the preferred axis and the direction of magnetization. Significant larger K~
values occur for amorphous transition-metal-based alloys when the alloy is formed
of two different magnetic atoms or for Co-based alloys (Ounadjela et al. 1989). The
strong uniaxial anisotropies occurring in R-T films are due to the deposition process
(evaporation, sputtering) causing locally an anisotropic atomic arrangement that
leads to a preferred axis parallel to the film normal (K, > 0) or to an easy plane
(K, < 0) of magnetization. The magnitude of K, depends on the degree of short-
range order and the magnitude of the magnetic anisotropy per atom. The former is
primarily controlled by the energy of the atoms at the growing surface and the latter
by spin-orbit coupling leading to high anisotropies for R-T alloys with non-S-state
rare earths like Tb or Dy. Different origins were discussed to account for the observed
anisotropies such as structural inhomogeneities (Graczyk 1978, Herd 1977, 1978,
1979, Katayama et al. 1977, Leamy and Dirks 1979, Mizoguchi and Cargill III 1979,
374 P. HANSEN

Yasugi et al. 1981, Kusuda et al. 1982a,b), incorporation of oxygen (Brunsch and
Schneider 1978, Dirks and Leamy 1978, Leamy and Dirks 1979, Biesterbos et al.
1979, Tsunashima et al. 1980, Hoshi et al. 1982, van Dover et al. 1986, Heitmann et
al. 1987a,b, Klahn et al. 1988), columnar microstructures (Suzuki 1983), stress-induced
anisotropies (Tsunashima et al. 1978, Leamy and Dirks 1979, Takagi et al. 1979,
Togami 1981, Labrune et al. 1982), dipolar interactions (Chaudhari and Cronemeyer
1975, Mizoguchi and Cargill III 1979, Wang and Leng 1990), pair ordering (Gambino
et al. 1974, Taylor and Gangulee 1976, 1977), anisotropic exchange (Meiklejohn et
al. 1987), bond orientation and anelastic deformation (Egami et al. 1987, Hirscher et
al. 1990, Y. Suzuki et al. 1987).
In one-subnetwork 3d-based alloys, usually low anisotropies are found originating
primarily from the magnetostriction coupling the magnetization to the internal
stresses (Egami et al. 1975, O'Handley 1975) or pair ordering (Luborsky 1977,
Miyazaki and Takahashi 1978). The anisotropy constant of amorphous SiB-FeCo
alloys in the as-quenched and field-annealed state (Miyazaki and Takahashi 1978) is
shown in fig. 63. The observed anisotropies were associated with pair ordering.
Typically Ku values up to 6 x 10 2 J m -3 were reached by field annealing (Luborsky
1978, Fujimori et al. 1984, Maehata et al. 1986) which is shown in fig. 64 for
amorphous B-FeNi alloys. These anisotropies are associated with short-range pair
ordering and interstitial or monoatomic ordering of the metalloids. However, also
large anisotropies have been found in rare-earth-based one-subnetwork alloys pre-
pared by liquid quenching as reported for Gd-Cu alloys (Algra et al. 1980). The
uniaxial anisotropy in vapor-deposited rare-earth-transition-metal films strongly
depends on the composition, the R and T component and the deposition parameters.
The presence of oxygen leads to a selective oxidation of the R component and reduces

°'4/(a)
'~ i~ Si°'lO B°'13 (Fel-× Cox)o.77
0.2 4 • • • .
~ ~, • ~---¢"-

0/0 ,
0.2 0.4 0.6 0.8 1.0
X

O.lO

0.05
:z
I I
0 0,2 0,4 0,6 0.8 1.0
X

Fig. 63. Compositional dependence of the room-temperature uniaxial anisotropy constant for amorphous
(a) as-quenched and (b) annealed (T~ = 573K and H = 184kA/m) SiB-FeCo alloys prepared by melt
quenching (Miyazaki and Takahashi 1978).
MAGNETIC AMORPHOUS ALLOYS 375

6 Bo.2o(Vex
Nil->,)o.8o / ~
7o:===oo~ ' / ~ . . ~ - ' k ' k " , /
]
= j .~_ . _ , _ L Z/'06.. ..... _'x.~\~, ... /

.,
7° , 32s~...", ~ ~ . k . \ ~ . - ' t
# 1i...:j ..... ...., . . . . . ".M
/ lillS'O.IOBo.12(FexC01-x)0.78.-#, : " ~ ~ ; ~
15.oi.4" i.." { ~'. -~
/Po.,<Bo.os(FexNil-xlo.so11:2..216~0 ~325 %~
0i b: ./ /300 / "
0,2 O.A 0.6 0,8 1.0
x
Fig. 64. Compositional dependence of the magnetically induced uniaxial anisotropy (Luborsky 1978) in
B-FeNi (open symbols)(Luborsky and Walter 1977a), PB-FeNi (solid symbols) (Luborsky and Walter
1977b), and SiB-FeCo (crossed symbols)(Fujimori et al. 1977).

Ku (Shen et al. 1981, van Dover et al. 1986, Klahn et al. 1988). Columnar structure
appears to be less significant and also stress-induced effects via magnetostriction are
not dominant as concluded from films before and after removal of the substrate
(Hoshi et al. 1982, Y. Suzuki et al. 1987).
Pair ordering and dipolar interaction lead to a relation between Ku and the
sublattice magnetizations that can be expressed in the form
K u = ~ CikMiMk. (66)
i g:k

Using the sublattice magnetizations inferred from the mean-field analysis, the
dependences of Ku on composition and temperature can be calculated. The com-
parison with experimental data revealed a good agreement for various alloy com-
positions (Taylor and Gangulee 1976, 1977, Taylor et al. 1978, Hansen and Hartmann
1986, Mansuripur and Ruane 1986, Hansen and Witter 1988, Hansen et al. 1989,
1991). The compositional variation of K~ at room temperature is shown in fig. 65
for amorphous G d - F e (Hansen et al. 1989) and G d - C o (Taylor and Gangulee 1976)
alloys prepared by evaporation. The full lines were calculated from eq. (66). Although
a good fit of the experimental data was achieved, the reliability of the evaluated
coefficients appears to be limited due to the high sensitivity of Ku with respect to
small structural variations or to additions of other elements like Si, Ge, Sn or Au,
giving rise to even negative K, (Hartmann et al. 1985b). These dependencies demon-
strate the strong influence of composition on sign and magnitude of Ku which also
applies to sputtered films where the sputter gas, pressure-distance product, bias
voltage or substrate temperature are relevant parameters affecting Ku. The influence
of the substrate temperature on Ku is demonstrated in fig. 66a for magnetron sput-
tered T b - F e alloys (van Dover et al. 1985), reflecting the reduction in mobility of
the adatoms on the film surface which decreasing substrate temperature T~. The
increase of Ku with T~ was confirmed for N d - F e (T. Suzuki et al. 1987) and T b - F e
376 P. HANSEN

0.4

0.2

"o 0 -~--~ , , , , l, ~0
2

-0.2

[a) o -u

-0.4

0.4

0.2
OO

Gdl_xCo x
-- calculated
E

o
-0.
2

-0.
o ~

-0.6

(b)
-0.8 ~
0.95 0.85 0,75 0.65
X

Fig. 65. Compositional dependence of the uniaxial constant of (a) Gd-Fe (Hansen et al. 1989) and
(b) Gd-Co (Taylor and Gangulee 1976) alloys prepared by evaporation. The full lines were calculated
from eq. (66) where the sublattice magnetizations were inferred from mean-field calculations.

(Kobayashi et al. 1983, Takeno et al. 1986) alloys. The anisotropy of amorphous T b -
Fe alloys is larger by more than one order of magnitude as compared to the Gd-
based alloys which is associated with the high single-ion contribution of the Tb. In
that case, a linear increase of Ku with the R content is expected and indeed can be
MAGNETIC AMORPHOUS ALLOYS 377

Tbl-x Fex

0.5 0.6 0.7 0,8 0,9 1.0


X

o /
6
• ~ o

b,.x(Gd,Fe}x Tb,_xlFe,,Co~x /

23 el o

,!
2 • i

¢
1

b) !

O( ] I i
0.1 0.2 0.3 0.4
1-x
Fig. 66. Compositional dependence of the uniaxial anisotropy constant at room temperature (a) for
magnetron sputtered Tb-Fe alloys at two substrate temperatures T~ (van Dover et al. 1985) and (b)
evaporated GdTb-Fe and Tb-Fe alloys (Hansen and Witter 1988, Hansen et al. 1989). Circles and
triangles in fig. (a) represent vibrating sample magnetometer and torque magnetometer measurements,
respectively.

observed (Sato et al. 1985, Hansen et al. 1989) as shown in fig. 66b. The slope depends
on the deposition parameters (Klahn et al. 1990b). The variation of the room-
temperature Ku as a function of the R element (Y. Suzuki et al. 1987) is shown in
fig. 67 for (Gdo.75Ro.2s)o.19Coo.81 alloys. A clear maximum of Ku appears around
Tb and Dy. The two curves in fig. 67a represent the measured Ku before (open circles)
378 P. HANSEN

(Gdo.75 R 1.25)0.19Co 0.81


6
5
g" /+
E
3
o
2
2
1

-1
{a)
-2
I I I I I I I I I I [ I I
La Ce Pr Nd PmSm Eu Gd Tb Dy Ho Er Tm Yb Lu
R

iI"
i
/theory
o amorphous
+ crystatline

1,~

<3_ 1
-2
-3

o
-5 -(b)
I f I I I I I I I I "~" [
LaCe Pr Nd PmSmEu Gd Tb DyHo Er TrnYb Lu
R
Fig. 67. (a) Uniaxial anisotropy constant for amorphous GdR-Co alloys, prepared by sputtering before
(open circles) and after (solid circles) removal of the substrate, versus rare-earth element and
(b) uniaxial anisotropy per rare-earth atom versus R element for G d R - C o alloys (Y. Suzuki et al. 1987).
The crosses in fig. (b) refer to crystalline compounds (Tajima 1971) and the full line was calculated from
the single-ion model [eq. (68)].

and after (solid circles) removal of the substrate, indicating that the stress-induced
contribution,
K~ = - ~-o-2=, (67)
does not play the dominant role, but in Tb-Fe-based films it was found responsible
up to 40 percent of the total anisotropy (Cheng et al. 1989). a and 2= denote the
stress- and saturation-magnetostriction constant, respectively. The single-ion contri-
bution of the S-state Gd is zero. The K, value of the G d - C o alloy thus can be
considered as a reference. Subtracting its K, from the stress-free data (solid circles),
the variation of the single-ion contribution is obtained leading to the variation shown
MAGNETIC AMORPHOUSALLOYS 379

in fig. 67b. The crosses refer to crystalline materials (Tajima 1971). The full line was
calculated from the equation (Y. Suzuki et al. 1987)
( l']e2q(r2)6
Ku -
3 esJ J - 2] a 3 x,
10
(68)

where J is the total angular momentum of the R atom and as is the second-order
Stevens factor (Stevens 1952). e, q, a and x are the electron charge, the valency of the
neighboring ion, the average distance between nearest neighbors and the R concentra-
tion, respectively. 6 is the local strain of the atomic environment produced by inelastic
deformation and (r 2) represents the average distribution of the 4f wave function.
Corresponding results were obtained for evaporated R-Co and R-Fe alloys
(Miyazaki et al. 1988, Takahashi et al. 1988a,b).
The temperature dependence of Ku strongly depends on composition. For many
alloys, the dipole equation [eq. (66)] accounts well for the measured results. This is
shown in fig. 68a for amorphous G d - F e and G d - C o alloys. The temperature varia-
tion for some Tb-containing alloys is given in fig. 68b, indicating a steeper slope of
K, in the high-temperature range (Hansen et al. 1989). However, also a sign change
can occur as it is observed for H o - C o alloys (Hansen et al. 1991). This is shown in
fig. 69 and can be discussed in terms of eq. (66) assuming different signs for the
coefficients. The temperature variation of Ku for ferromagnetically coupled N d - F e
and N d - C o alloys (Miyazaki et al. 1987a,b) is presented in fig. 70.

5.7. Magnetostriction
The magnetostrictive effects in amorphous alloys originate from the magnetoelastic
interactions associated with the local anisotropies and the local strains controlling
the local direction of the magnetic moments. The origin of the local strains was
discussed in terms of the single-ion model with random local axis (Cochrane et al.
1974, Ffihnle and Egami 1982, Suzuki and Egami 1983, Lachowicz and Szymczak
1984, O'Handley and Grant 1985, Furthmfiller et al. 1986, 1987a,b, Szymczak 1987,
F/ihnle and Furthmfiller 1988, 1989, Pawellek et al. 1988, Suzuki and Ohta 1988).
The macroscopic magnetostriction representing the relative changes in length and
volume is related to the macroscopic magnetoelastic constants which have to be
considered as an average over the local elastic contribution. The linear magneto-
striction can be expressed by

2 = 32s(cos20- ½), (69)

where 0 is the angle between the magnetization and the strain, 2s represents the
saturation magnetostriction constant

2s = ~(211 -- 2±), (70a)


where 2 LIand 2± refer to the relative changes in length parallel and perpendicular to
the direction of magnetization. The spontaneous volume magnetostriction cos can be
380 P. H A N S E N

TIK)
100 200 300 L+O0 500
O! , , , ~ j

-0.2 x=O.70 ° 9 , , , ~ " ~ " - f - jv 1~"'~'

~:~, -0.6 O.70,~.,..o,.~

-1.2 (o)

12 ~ m ~ F~&~o
10

£
~ 6
~, Tbo.21Feo.sz,C°0.15

(b) I
!
O0 100 200 300 &00 500 600
T(K)
Fig. 68. Temperature dependence of the uniaxial anisotropy constant for (a) amorphous Gd-Fe and G d -
Co alloys and (b) amorphous T b - F e , T b - F e C o and G d T b - F e alloys (Hansen et al. 1989) prepared by
evaporation. The full and broken lines in (a) were calculated from eq. (66) where the sublattice magnetiza-
tions were inferred from mean-field calculations.

written as
cos = 211 + 22±. (70b)
Some typical values for amorphous transition-metal-metalloid alloys are compiled
in table 14 and fig. 71. Fe-rich alloys exhibit a large positive saturation magnetostric-
tion in contrast to polycrystalline iron. Many Co-rich alloys exhibit a small negative
2s, in agreement with that of crystalline cobalt, but with increasing Co content, 2~
becomes positive. The different sign of 2~ in these alloys has led to the development
of zero-magnetostriction alloys which are of interest for different applications (Boll
et al. 1983, Tag0 et al. 1985). Low magnetostrietion can easily be reached using Co-
based alloys with some additives (Sherwood et al. 1975, O'Handley et al. 1976c,
Gyorgy 1978, O'Handley 1978a, O'Handley and Sullivan 1981, Fujimori et al. 1984,
MAGNETIC AMORPHOUS ALLOYS 381

800

/" I H°°2° C°°'8°


600
Ku
E
£ 2 400
2
Ms

200

0 0

0 200 400 600


T(K)
Fig. 69. Temperature dependence of the uniaxial anisotropy constant and saturation magnetization for
an amorphous Ho-Co alloy prepared by evaporation (Hansen et al. 1991).

Chen and Rao 1986, Shiba et al. 1986, Jergel et al. 1989). The presence of a
compositional point for 2s and o~s is also demonstrated in fig. 72 for amorphous B -
CoFe alloys (Jergel et al. 1989). In particular a negative 2s was reported for amorphous
M - C o with M = B, Nb, Ta or W, and a positive 2s for M = Ti, Hf or Zr (Fujimori
et al. 1984, Shiba et al. 1986). The room-temperature magnetostriction constant was
shown to scale with a 2 for various amorphous alloys (du Tr6molet de Lacheisserie
1982), which is shown in fig. 73. However, many Co-based alloys do not follow this
dependence. The temperature dependence of 2, can be well described in terms of the
single-ion model for uniaxial symmetry predicting the relation (Callen and Callen
1965)
2s(T) = 2 s(0)[s/2 (x), (71)
where [5/z(X) is the reduced Bessel function,
3 3
[5/2(x) = 1 + ~5 - x c°thx,

and x is related to the relative magnetization, ms(T) = M~(T)/Ms(O), by


1
ms(T) = - - + coth x. (72)
x
The temperature variation of the reduced linear saturation magnetostriction versus
reduced saturation magnetization is displayed fig. 74 for different amorphous alloys
(O'Handley 1978a, du Tr6molet de Lacheisserie 1982). The data follow well the line
calculated from eq. (71), confirming the dominance of single-ion contributions in
382 P. HANSEN

1.0 x=0.63
~£/'~ NdlxFex
.53
0.8

\IE 0.6
e~

0.4

0.2
(a)
100 200 300 400
T(K)
7

65 x=O~ Ndl-xC°x

4 0.69"~
E
o 3 0,85
2

Ib) 0.52
I I I I
_"
0 100 200 300 400
T(K)
Fig. 70. Temperature dependence of the uniaxial anisotropy constant for amorphous (a) Nd-Fe (Miyazaki
et al. 1987a,b) and (b) Nd-Co alloys (Takahashi et al. 1988a,b) prepared by evaporation.

these alloys. Co- and CoFe-based alloys again deviate from the pure single-ion
behavior and their temperature dependence suggests the presence of two-ion contri-
butions (O'Handley 1978a, Hernando et al. 1984, Chen and Rao 1986) which are
proportional to m2(T) (Callen and Callen 1965). Then, 2s can be expressed by
)~s(T) = 21 [5/2(x) + 22m2(T), (73)
with 21 < 0, 2 z > 0 and x defined by eq. (72). The occurrence of a compositional and
temperature compensation of 2s in CoFe-based alloys confirms the presence of these
two mechanisms contributing to 2s. However, also the competition between two
single-ion contributions was discussed to cause 2~ = 0 in B-CoMn alloys (du Tr6molet
MAGNETIC AMORPHOUS ALLOYS 383

TABLE 14
Room-temperature saturation magnetostriction for amorphous metal-metalloid alloys and rare-earth-
transition-metal alloys.
Alloy 2s(x 106) Refs.*
Bo.zoFeo.so 31.0 [1]
Po. 13Co.ovFeo.so 31.0 I-23
Po.14Bo.06 Feo.40Nio.4o 11.0 [3]
Bo.2oFeo.3oNio.so 8 [4]
Po.o9Coo.ga -4.3 [5]
Bo.2oCoo.so -4 [-4]
Bo.loSio.15Coo.75 -3 [-6]
Bo.2oFeo.o6Coo.v4 ~0 I-7]
Bo.2oNo.4oCoo.4o - 7 [4]
Pro.2o Feo.so 96 tl) [8]
Smo.zoFeo.so 140~a) [-8]
Gdo.33 Feo.67 20 [9]
Tbo.2o Feo.so 220 ~1) [8]
Tbo.33 Feo.67 310~z) [10]
DYo.aa Feo.67 38t2) [10]
Gdo.3oCoo.7o 5 [11]
Gdo.23 C0o.77 33 [11]
(1)At H = 1600kA/m.
(2)At H = 2000kA/m.
* References:
[1] O'Handley (1977b). [6] Arai et al. (1976).
[-2] Tsuya et al. (1975). [7] Jergel et al. (1989).
[3] O'Handley (1976). [-8] Ishio (1988a).
[4] O'Handley (1978a). [-9] Moorjani and Coey (1984).
[5] Simpson and Brambley (1971). [10] Clark (1980).
[11] Twarowski and Lachowicz (1979).

de Lacheisserie and Yavari 1988). The magnetostriction arising from a disorder


caused by randomly oriented easy axes follows a power law 2s(T) = 2s(0)m,(T)" with
n < 3 (Cullen and del Moral 1990).
The magnetostriction in amorphous rare-earth-transition-metal alloys is com-
posed of an R and a T contribution where in the case of non-S-state rare earths 2s
is primarily controlled by the strong single-ion contribution of the R atoms (Suzuki
et al. 1988, del Moral and Arnaudas 1989) in accordance with the anisotropy. Alloys
containing S-state rare earths are characterized by a magnetostriction which is
comparable in magnitude with that of the T-based alloys. This is demonstrated in
fig. 75, showing 2s versus x at T = 80 K for amorphous Y - T and G d - T alloys with
T = Fe or Co (Yoshino and Masuda 1988). In the case of Y-Fe and G d - F e (fig. 75a),
the variation of 2s reflects the behavior of Ms and Tc versus x. The value of 2s for
G d - F e is higher than that of Y - F e alloys due to the positive Gd contribution. At
the R-rich side, 2~ approaches zero for Y - F e or approaches the 2~ value for Gd in
the case of G d - F e alloys because the Fe contribution vanishes at low x. At the Fe-
rich side, 2s decreases due to the noncollinear structure of the Fe sublattice. Amor-
phous Y - C o and G d - C o alloys (fig. 75b) show a corresponding behavior at the
384 P. HANSEN

Bo.2o(COl_xF ex)o.eo i./~"~A\

P0.13C0.o71COl_xF ex}o.75 ,,/


2.--
g • S /"

lO ,"• /..Y
/ /;~
"- \

0
t SIo.15Bo.lo( C%xFex)o.7s
'<~ I I I I
0.2 0.4. 0.6 0.8 1.0
X

Fig. 71. Compositional dependence of the saturation Illagnetostriction for various amorphous metal-
metalloid alloys (Luborsky 1978): PBA1-FeCo (Brooks 1976), PC-FeNi (Arai et al. 1976), B-FeNi
(O'Handley 1977c, Luborsky 1978), PB-FeNi (Luborsky 1978), PC-FeCo, SiB-FeCo (Fujimori et al.
1976), and B-FeCo (O'Handley 1977c).

0 0.04 0.08 0,12 0.16 0.20


×

Fig. 72. Compositional variation of the saturation linear magnetostriction (solid symbols) and the
spontaneous volume magnetostriction (open circles) for amorphous B-FeCo and B-CoNi alloys (Jergel
et al. 1989).

R-rich side. The additional minimum of 2 S occurring at x --- 0.4 can be attributed to
the vanishing Co moment at the critical concentration xc ~ 0.5. The decrease of 2s
at small x is associated with the negative 2s for Co-rich alloys. Some magnetostriction
data for amorphous R - T alloys are compiled in table 14.
Amorphous alloys with non-S-state rare earths exhibit magnetostrictive effects
higher by more than one order of magnitude as it is demonstrated in fig. 76a. The
measurements were performed in fields up to 1.6 x 106 A/m (Ishio 1988a, Ishio and
Miyazaki 1988). At temperatures below Trio (R = Dy or Tb), the coercive field strongly
MAGNETIC AMORPHOUS ALLOYS 385

30

~<~ 20
%

10 J
/ 1 2 3

~
0 t I I
-z.o.-2(A2m~/kg2)

-10 i i
Fig. 73. Room-temperature saturation magnetostriction versus the square of relative saturation magnetiza-
tion for different amorphous transition-metal-metalloid alloys (du Tr6molet de Lacheisserie 1982): (O)
B-Fe (Narita et al. 1979), (A, V) B-FeNi, ( i ) PB-FeNi (O'Handley 1977c), ([]) B-FeNi (O'Handley
1978a), (©) Fe-based alloys (Itoh et al. 1980).

increases due to the presence of a compensation temperature preventing correct ,~


measurements. The sign of 2 corresponds to that found in the corresponding crystal-
line counterparts (Clark 1980). The variation of 2 in the series Ro.2oFeo.8o is presented
in fig. 76b and was explained in terms of the single-ion theory (Ishio 1988a, Suzuki
et al. 1988, Ishio et al. 1990).
However, the data shown in fig. 76 are measured in a nonsaturated state. Magnetic
saturation can only be reached at much larger magnetic fields as already discussed
in section 5.3. The field dependence is displayed in fig. 77, revealing an increase of 2
in fields up to 2 × 106 A/m (Clark 1980) which are known to be much too low to
achieve magnetic saturation due to the high locally varying anisotropy fields.
The high R contribution to 2s can be used to enhance or suppress the magnetostric-
tion, e.g., in amorphous B-Fe-based alloys (Kazama and Fujimori 1982, Gr6ssinger
et al. 1988). The large forced volume magnetostriction O~o/OH reveals comparable
magnitudes and dependencies on composition for R-Fe and Y-Fe, B-Fe alloys and
thus appears to be primarily controlled by the iron (Ishio 1988b).
The magnetostrictive effects originate from the local strains which correlate 2s and
a~s to the local atomic arrangement. This leads to a strong dependence of these
properties on the deposition parameters (Itoh et al. 1983).

5.8. Coercivity

The coercive field and the losses are controlled by the process of magnetization
reversal and thus depend on magnetic nucleation, rotation of magnetic moments and
386 P. HANSEN

1.0

0,8 o ~

.~ 0.6
,.<

o.z,

0.2

0 I I I I I I I I
0.8 0,6 0./+ 0.2 0
ms(T)
Fig. 74. Reduced saturation magnetostriction versus reduced specific saturation magnetization for various
amorphous transition-metal-metalloid alloys (O'Handley 1978a, du Tr6molet de Lacheisserie 1982). The
full line represents the single-ion model [eq. (71)]. (O) Bo.2oF%.so (O'Handley 1977a), (©) Bo.t4Feo.s6
(O'Handley et al. 1979), (A) Bo.2oNio.5oFeo.3o, ( , ) Bo.EoCOo.4oNio.#o (O'Handley 1978a), (~)
Co.toPo.lsFeo.75 (Berry and Pritchet 1978).

30
Gdl_xFex
,~20

10

30

20 Gdl-xC°x'~
10

0 cbl 0.20
,-'-'o,-"~,0.40 ", 0.60
,~
x
Fig. 75. Compositional variation of the saturation magnetostriction at T = 80 K for (a) amorphous Y-Fe
and Gd-Fe alloys and (b) Y-Co and Gd-Co alloys (Yoshino and Masuda 1988).
M A G N E T I C A M O R P H O U S ALLOYS 387

600 R°'2°Fe°'8°

400

200

-200

-400
(a) l I ,
100 200 300 400
T(K)
8OO
~ R°'2°Feo.so
6OO
4OO

K
20O
=o

-200
-400 V
-600 (b)
I I I I I I I I 1 I I I I
La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb

R
Fig. 76. (a) Temperature dependence of the magnetostriction at H = 1.6 x 106 A/m and (b) variation of the
magnetostriction at T = 77 K and H = 1.6 x 106 A/m as a function of the rare-earth element for amorphous
R - F e alloys (Ishio 1988a). At temperatures below THoin fig. (a), the coercivity strongly increases preventing
correct X measurements. The full line in fig. (b) represents the single-ion theory.

domain-wall motion. The losses additionally increase with frequency due to the
presence of eddy currents. The rotation of magnetization and domain-wall motion
are associated with the anisotropy, strain, exchange interaction, demagnetizing effects
and the presence of structural and surface inhomogeneities where the latter in
amorphous materials are sensitive functions of the preparation conditions and anneal-
ing treatments.
There exists two extreme pictures to describe the magnetization reversal: (i) coher-
388 P. HANSEN

300 ~ 6 7

,-< 200
~o /7 Tb°.l° DY°.23Fe°.67

100 f DYo.33Feo.67

0 ~ t i
0.5 1.0 1.5 2.0
H(10SA/m)
Fig. 77. Field dependence of the room-temperaturemagnetostriction of amorphous Tb-Fe, Dy-Fe and
TbDy-Fe alloys (Clark 1980).

ent rotation (Brown Jr 1945, Stoner and Wolhfarth 1948, Aharony 1962) involving
a simultaneous rotation of all magnetic moments and (ii) domain-wall motion. The
former process yields a coercive field determined by the nucleation field,
2K
Hc- -- Neff Ms, (74)
#oMs
where K is the anisotropy constant and Neff the effective demagnetizing factor. The
latter process involves local nucleation and domain expansion controlled by the
presence of defects, local material inhomogeneities, surface roughness or intrinsic
magnetic fluctuations caused by structural disorder (Dekker 1976, Gyorgy 1978,
Kronmfiller 1981 a,b, Kronm/iller and Gr6ger 1981, Mansuripur 1982, Ramesh and
Srikrishna 1988). The coercive field arising from these irregularities can be calculated
in terms of a statistical theory (Pfeffer 1967, Kronmfiller 1970, 1973)

_ 1 q dx/~ , (75a)

where q = (p/2F)In(teL~25) and the wall is assumed to move along the x axis. F, p
and L are the domain-wall area, the density of pinning centers and the domain width,
respectively. E denotes the energy fluctuation according to the present irregularities.
The brackets indicate the averaging over the statistical parameter. The domain wall
thickness, 5 = ~/-A/K, is determined by the micromagnetic exchange stiffness constant
A and the anisotropy constant K. The coercive field originating from domain wall
motion thus can be calculated from eq. (74) yielding
qp6"
Hc - (75b)
#oMs'
where p and n comprises the particular features of the material irregularity. Aniso-
tropy and exchange fluctuations give rise to n = - ½ and n -- - {, respectively (Kron-
MAGNETIC AMORPHOUS ALLOYS 389

mfiller 1981a,b). Surface roughness leads to n = 1 (Gyorgy 1978, Kronmfiller 1981b)


and in this case the magnitude of Hc estimated from eq. (75) is of the order of
0.7 A/m and thus represents one of the limiting factors for the coercivity of magnetic
glasses. Much higher coercivities arise from elastic stress centers because they repre-
sent much stronger obstacles for domain movement through the magnetostrictive
coupling. This contribution resulting from dislocation dipoles gives rise to n = - 1
(Grrger and Kronmfiller 1981) and p is proportional to the saturation magneto-
striction constant and the shear modulus. This contribution implies the dependence
of Hc ~ K1/4and was found to control H~ in Fe-based glasses which is demonstrated
in fig. 78a. In low-magnetostrictive Co-based alloys, surface irregularities and
relaxation effects are the relevant factors governing He. In these soft-magnetic
materials, the defect-controlled domain-wall motion accounts well for the observed
coercivity.
The other extreme where no free domain-wall displacement takes place is rep-
resented by hard-magnetic materials used for permanent magnets. In particular

1.75
Po.lzBo.o
, sFeo.~.oNio.z,o

~
o
1.50

1.25
/ B°'2°Fe°'I'i/

1.00

0.75 I
5
I I
6
I
7
I I I
8

Kl/t. (j/m3)1/z.
5'OOI(b) "S
1.00

:~
~,, 0"50I o~ "Tb-Fec°
I- 0

0.051- • ~.

0.o1 . ' ' - - '5'.o'"1'o.o


K.ulloSj/rn31
Fig. 78. Room-temperature coercive energy versus uniaxial anisotropy constant (a) for amorphous B-Fe,
B-FeNi and PB-FeNi alloys (Gr6ger and Kronmfiller 1981)and (b) for various sputtered and evaporated
amorphous Gd-Fe, GdTb-Fe and Tb-FeCo alloys (Klahn et al. 1990a,b).
390 P. HANSEN

polycrystalline N d - F e - B and Sm-Co are well-known as suitable materials for


permanent-magnet applications. In this case, the grain size is a further relevant
parameter determining Ho while in amorphous alloys the local structure controlled
by the deposition parameters is of primary importance. In this case, the coercive
field and its temperature dependence were discussed in terms of a generalized form
of the nucleation field (Kronmfiller et al. 1988, Givord et al. 1990)
2K
Hc = c t - - - Nef f M s , (76)
#o Ms
where ~ represents a microstructural parameter and Nef f a n averaged local demagne-
tizing factor. The temperature dependence of He for many crystalline hard-magnetic
materials can be well described in terms of eq. (76), treating ~ and Neff as adjustable
parameters (Hirosawa et al. 1986, Hirosawa and Sagawa 1988, Hirosawa 1989,
Givord et al. 1990). Another approach to calculate He for hard-magnetic materials
with a uniaxial anisotropy constant K, > 0 is based on the solution of the differential
equations for the case that the domain wall passes a planar barrier with different
material properties, yielding (Friedberg and Paul 1975, Hilzinger 1977, Paul 1976,
1982)

1.21A ( A K , _ Kb'] (77)


Ho- \ J"
A denotes the thickness of the barrier. A b and K b a r e the exchange stiffness constant
and the anisotropy constant inside the barrier, respectively. Equation (76) is restricted
to barriers satisfying A ~ 6.
For many magnetic materials, however, both local rotation controlled by the
nucleation field in a small volume and domain-wall propagation are responsible for
the observed He. The activation volume depends on temperature and thus the
temperature dependence of He is also determined by thermal activation processes
(Rio et al. 1987, Givord et al. 1988, Labrune et al. 1989, Givord et al. 1990). Rare-
earth-transition-metal alloys are characterized by rather high coercivities except for
those containing only S-state rare earths. Domains imaged in amorphous Tb-Fe
alloys indicate the importance of nucleation processes. In this case, Ku is related
linearly to the Tb content as demonstrated in fig. 66b.
According to eqs. (75)-(77), the coercivity is expected to increase with K,~ where n
is determined by the mechanism controlling H c. The linear increase of K, with the
R content in amorphous R - T alloys was well established and a plot of ~oMsHe
versus Ku presented in fig. 78b for Tb-containing alloys yields n - 1.3 (Klahn and
Hansen 1991). These data thus support neither a pure nucleation process (He ~ Ku),
nor just domain wall motion. In the latter case, the calculated Ho values for exchange
fluctuations (H e KSu/4) are much too low to account for the observed coercivities
~

and surface roughness, pinning through anisotropy fluctuations, clustering or stresses


on an atomic scale give rise to a much weaker Ku (n < 1) dependence (Kronmfiller
1973, 1981b, Kronmiiller and Gr6ger 1981). However, the local stresses may be quite
large as compared to the average macroscopic magnetoelastic interaction and, thus,
MAGNETIC AMORPHOUS ALLOYS 391

might also be important for the interaction with thin domain walls in particular in
Tb-containing alloys. The direct relation between Hc and Ku and the single-ion
origin of Ku varying with 2jd 2 [eq. (68)] for amorphous R-T alloys suggests Hc to
vary with the same factor which indeed was well confirmed experimentally for the
heavy rare-earth-transition-metal alloys (Komatsu and Fukamichi 1989, Komatsu
et al. 1989). Also, the Ku variation of Ho predicted for the domain-wall pinning
through a planar barrier (H¢ ~ K 3/2) is in reasonable agreement with the observed
dependence (fig. 78b).
The presence of a compensation temperature is a factor giving rise to severe
changes of the magnitude and the temperature profile of Ho. This is demonstrated
in fig. 79 for different amorphous alloys of composition (Gd, Tb)l_xFex with x
around 0.7 (Hansen and Witter 1988). The total T~ompshift corresponds to a change
in x of the order of 0.03. In the vicinity of T~omp,the energy of #oMsH¢ is almost
constant because Hc varies as M~-1 leading to a singularity of Hc at T~omp.The
resulting profile can be utilized to tailor these films for magneto-optical data storage
as discussed in section 8.2. In this case, also the Hc versus T characteristic in the
switching range near Tc is of significant importance. This is shown in fig. 79 for
GdTb-Fe and Tb-FeCo alloys with different Tb content and compensation temper-
atures. The strong anisotropy of Tb leads to a very steep temperature dependence
of H¢ near Tc and the position of T~omvaffects the curvature of He.
The interpretation of H~ data for alloys of almost identical composition requires
a detailed knowledge about the preparation conditions due to the high sensitivity of
H~ on structural disorder. Annealing of the alloys lead to a drastic reduction of Hc
due to structural relaxation and oxidation processes.

400
x=0.664 x=0.710 x=0.738 x=0.729 x=0.689
¢=0.111 y=~66 y=0.169 y=0.181 y=0.163

300

..,,. ----t ~_yb}l_, , ' \ ~


.<
Tbl_×(gel_yCoy)× q:~

::= 20C

,~o,76 '~
100

x= 0,760 ~ ~
y= 0.333 'K'
I I
~00 350 400 450 500
T(K)
Fig. 79. Temperature dependence of the coercive field for evaporated amorphous GdTb-Fe and
Tb-FeCo alloys with different compensation temperatures [(~7) T~omp=170K, (O) T~omp=181K,
(D) Tcomp= 395 K, (A) T~o,.p= 466 K] (Hansen and Witter 1988).
392 P. HANSEN

Most transition-metal-metalloid glasses are soft ferromagnets exhibiting techni-


cally interesting properties. In this respect, three groups of ribbons can be distin-
guished (Warlimont and Boll 1982): (i) Fe-based alloys with high Ms and low power
loss, (ii) FeNi-based alloys combining good soft-magnetic properties with intermedi-
ate Ms and (iii) Co-based alloys with low Ms, 2s and excellent soft-magnetic properties.
Typical coercive fields range from 0.24 to 8 A/m for BSi-CoNiFe and B-Fe alloys.
Typical Hc values for B-CoFe glasses (Gyorgy 1978) are displayed in fig. 80. Hc is
at a miminum when 2s passes through zero, indicating the importance of stresses.

5.9. M a g n e t i c e x c i t a t i o n s

Single-particle and collective-particle excitations contribute to the excitation


spectrum of magnetic materials. The spin waves or magnons are the excitations of
lowest energy and involve a deviation of the moment from the parallel or antiparallel
alignment of the ground state. This requires an exchange energy D k 2 according to
eq. (54) where D is the spin-wave stiffness constant which is related to the exchange
constants and k is the wave vector of the spin wave. D is a function of temperature
and can be expressed by

D ( T ) = D(O) 1- a , (78)

where n = ~ for an ideal Heisenberg ferromagnet (Dyson 1956, Izuyama and Kubo
1964) while the band theory predicts n = 2 (Izuyama and Kubo 1964, Mathon and
Wohlfarth 1968). Typical values for D(0) are compiled in table 15. D(0) is expected
to scale with Tc. This indeed was found for many Co-based alloys (Suran et al. 1981,
H/iller 1986) and is demonstrated in fig. 81. Different slopes were found for Co-
metalloid alloys and Co-transition-metal alloys, which probably can be attributed
to the different bonding responsible for the alloy formation. For Fe-based alloys, the
D(0) also increase with Tc but the data do not follow straight lines (H/iller 1986)

8 25
E B°'2°Fes"e°'xCOx 20

H¢ 15

0i i l i i I -5
0.50 0.60 0.70 0.80
X

Fig. 80. Compositional variation of the room-temperature coercivity and saturation magnetostriction of
amorphous B-FeCo alloys (O'Handley et al. 1976c).
MAGNETIC AMORPHOUS ALLOYS 393

TABLE 15
Spin-wave stiffness constant at T = 0 K and Curie temperature for amorphous alloys and crystalline iron,
cobalt and nickel.
Alloy Tc (K) D(0) (10 - 23 eV m 2 ) References
Fe(cryst.) 1040 311 Aldred and Froehle (1972)
Co(cryst.) 1400 580 Pauthenet (1982a,b)
Ni(cryst.) 630 359 Aldred (1975)
Bo.l,~Feo.86 556 118 Ishikawa et al. (1981)
Bo.18Feo.s2 617 167 Rhyne et al. (1982)
Po.26 Feo.74 580 115 Hfiller and Dietz (1985)
Bo.t3Sio.o9Feo.78 710 156 Yu et al. (1988)
Bo.17Wo.o5 Feo.75 450 59 Kaul and Mohan Babu (1989)
Bo.EsFeo.4oNio.35 637 120 Kaul and Mohan Babu (1989)
Co.ovCro.loPo.13Feo.7o 360 60 Xianyu et al. (1982)
Po.16Bo.o6Alo.o3Feo.75 630 134 Birgenau et al. (1978)
PoA6Bo.o6Alo.o3Feo.37Nio.38 482 91 Birgenau et al. (1978)
Po.25 Coo.7s 480 105 Hfiller and Dietz (1985)
Bo.25Sio.o5Coo.7o 539 115 Maskiewicz (1982)
Bo.llSio.llFeo.osCoo.7o 738 221 Swierczek and Szymura (1988)
Tio.229Coo.771 550 164 Suran et al. (1981)
Po.aoNio.9o 280 140 Hfiller and Dietz (1985)
Yo.33Coo.67 470 480 Fukamichi et al. (1986b)
(Gdo.26Coo.74)o.95Moo.o5 - 470 Macsymowicz et al. (1985)

40(3

Ml_xCox
~'E 30C

200

I00 ~ "~°~¢°

t
0 500 1000
Tc(K)
Fig. 81. Spin-wave stiffness constant at T = 0 K versus Curie temperature for various amorphous
alloys (Luborsky 1980, H/iller 1986). (+) M = W (Hfiller et al. 1985), ( . ) M = Ti (Suran et al. 1981),
(C]) M = Bi (Maeda et al. 1976), ( 0 ) M = P (Hiiller and Dietz 1985), ((3) M = BSi (Maskiewicz 1982).

which is probably associated with the noncollinear magnetic structure for Fe-rich
alloys and those with compositions near xc. D(T) is correlated to the coefficient B
of the T 3/2 term [eq. (55)] governing the low-temperature variation of the saturation
magnetization. D(T) can be determined from the spin-wave excitations using standing
spin waves in thin films, from Brillouin scattering, magnetization data or neutron-
scattering experiments. The spin-wave resonance condition for the resonance field
394 P. HANSEN

applied normal to the film plane reads

D (n~z~2] (79)
c°=?IH"--4~Ms +gltB \ h J j"
h is the film thickness and H, the resonant field of the nth spin-wave mode. The
temperature dependence of D(T) obeys the T 5/2 law for some amorphous alloys such
as B-Fe (Rhyne et al. 1982) or G d - C o M o (Macsymowicz et al. 1985), while for other
alloys D(T) deviates from this relation. The strong temperature dependence of D(T)
measured by inelastic neutron scattering is presented in fig. 82 for an amorphous
BSi-Fe alloy (Yu et al. 1988).
D(T) is related to the micromagnetic exchange stiffness constant A(T) by
MdT) ....
A(T)- 2 ~ B t)(l), (80a)

where g#B = 7h and 7 denotes the gyromagnetic ratio. A(T) is related to the domain
wall energy aw of a 180° Bloch wall by aw = 4A,,/-~. Thus, D(T) also can be estimated
from wall energy measurements. A(T) can be expressed in terms of the exchange
constants (Gangulee and Kobliska 1978b)
A(T) = { ~ N,S,(T) Z "]iJZur2Sj(T)' (80b)
J

where Ni, z u and rij are the number of i atoms per unit volume, the number of
nearest neighbors and the atomic distance between the atoms i and j, respectively.
Combinations of eqs. (80a) and (80b) yields a relation between D(T) and the sublattice
magnetizations in ferrimagnets. The temperature dependence of A(T) obtained from
standing spin-wave spectra in an amorphous G d - F e film (Vittoria et al. 1978) is
displayed in fig. 83.
The presence of non-S-state rare earths gives rise to a strong coupling between

150

Bo,13 Sio.o9 F eo.Ta~~


E 100
o
'o
I.-
50

I
200
I I
h00
I I
600
I
800
I

T(K)
Fig. 82. Temperature dependence ofthe spin-wave stiffness constant for amorphous BSi-Femeasured by
inelastic neutron scattering (Yu et a1.1988).
MAGNETIC AMORPHOUS ALLOYS 395

\
\
0 I I I I
100 200 300 400 500
T{K]
Fig. 83. Temperature dependence of the micromagnetic exchange stiffness constant for an amorphous
Gd-Fe alloy evaluated from standing spin-wave spectra (Vittoria et al. 1978).

the spin system and the lattice via spin-orbit coupling yielding high linewidth in the
resonance spectra (Lubitz et al. 1976) which prevents the determination of D(T).
It should be noted that strong differences were observed for D(0) values evaluated
from magnetization and neutron-diffraction experiments. This discrepancy was attrib-
uted to additional low-lying magnetic excitations (Fernandez-Baca et al. 1987).

5.10. Annealing effects

The amorphous structure represents a metastable state and thus any thermal treat-
ment causes a continuous change of the atomic arrangement. Different processes can
be distinguished: (i) corrosion and oxidation, (ii) structural relaxation and (iii) crystal-
lization. The latter describes the transition to the crystalline state and this process
requires an activation energy of the order of 2.5 eV. This is briefly discussed in
section 4.1. The first two processes affect the structural, magnetic and electrical
properties in a specific way (Egami 1984). The structurally sensitive properties like
the compensation temperature, the uniaxial anisotropy, the coercivity or the conduc-
tivity thus are suitable parameters to investigate the time and temperature dependence
of these processes. The oxidation and structural relaxation of magnetic glasses is less
important because they exhibit an even higher corrosion resistance than the corre-
sponding crystalline alloys (Waseda and Aust 1981). However, annealing in a mag-
netic field leads to strong changes in the magnetic properties (Luborsky 1978, 1980,
Egami 1984) (see fig. 64). The evaporated or sputtered rare-earth-transition-metal
films exhibit a low resistivity against corrosion and oxidation due to the high oxygen
affinity of the rare earths (Saito et al. 1986). The degradation by oxidation and
structural relaxation was investigated for different amorphous R - T alloys (Luborsky
et al. 1985, van Dover et al. 1986, Hartmann et al. 1987, Klahn et al. 1987, Greidanus
and Klahn 1989, Klahn 1990). The time dependence of the reduced coercive field for
a GdTb-Fe film with and without protection layers is shown in fig. 84. The bare
396 P. HANSEN

1.0 o,,, •

GdTb-Fe

i-\ 0.8 • mn •

o.7
• •

0.6 ibJ

i n Intalal I I lUlltll I I Ilnlnnl i , nlnllll I , nttnhnl

°%, I102 1031 I10 S I10


1hour 1day 1week 1month 1year
t(min}

Fig. 84. Degradation of the reduced coercive field for amorphous GdTb-Fe films (Greidanus and Klahn
1989). (O) Structural relaxation, (mm)trilayer in dry atmosphere, (A) oxidation of the bare film. The inset
illustrates the typical trilayer stack used for optical recording: (a) glass substrate, (b) dielectric layer, (c)
GdTb-Fe layer, (d) metal reflector, (e) organic protection layer.

film (triangles) shows a drastic reduction of H c due to the selective oxidation of the
rare earths. A typical trilayer configuration (insert of fig. 84) used for magneto-optical
recording protects the magnetic film and leads to a significantly higher stability
(squares). The circles represent the degradation of Hc due to structural relaxation.
In this case, the G d T b - F e film was coated with an 50 nm aluminium layer to avoid
any oxidation effects (Klahn et al. 1987). The structural relaxation process can be
described by a spectrum of activation energies (Gibbs et al. 1983, Fish 1985) according
to the various atomic rearrangements possible in an amorphous material (see
section 4.2). However, the structural degradation causes only small changes in the
magnetic and electronic properties as compared to the oxidation process that is
associated with a very low activation energy. Electrical conductivity measure-
ments performed on evaporated and sputtered R-Fe, R-Co, G d T b - F e , Tb-FeCo,
D y - F e C o , with R = Gd, Tb, Dy or Ho, films reveal activation energies ranging from
0.18 to 0.31 eV (Allen and Connell 1982, Klahn et al. 1988, Klahn 1990). The variation
of the normalized conductivity with annealing time for bare evaporated R - F e and
R - C o alloys is presented in fig. 85a. The Fe-based alloys clearly indicate a much
stronger structural degradation. Arrhenius plots for in-situ measured uncoated films
are shown in fig. 85b. Oxygen depth profiles indicate that an oxide layer grows into
the film leading to three different layers (van Dover et al. 1986, Aeschlimann et al.
1988): (i) a 10nm thick surface layer where both R and T atoms are completely
oxidized, (ii) a layer where mainly the R atoms are oxidized and (iii) the bulk layer.
The presence of oxidation processes, therefore, are very important in the interpreta-
tion of the properties of thin films.
The resistivity of thin R - T films against corrosion and oxidation can be improved
using dense films obtained by magnetron sputtering (Hong et al. 1986) and by small
additions of Ti, Nb, Ta, Cr (Imamura et al. 1985, Niihara et al. 1988), Hf, Mo
(Kobayashi et al. 1988), In (Iijima 1987, 1988) or Pt (Hatwar and Majumdar 1988).
MAGNETIC AMORPHOUS ALLOYS 397

1.0

/Dy-Co
0.9 Tb-Co

"6
0.8

0.7

(a)
0.6 I I I f I i I I I
20 40 60 80 100
ta(h}

T (C°)
160 110 90 65 25
I I I I ,~A

(b)
=0.31eV
-5

Tbo.22 Feo.3t, Co o.t,,~

:O,23eV
-4

-3

-2

3
110-3 K 1

Fig. 85. (a) Variation of the normalized conductivity with annealing time for bare amorphous R-T alloys
prepared by evaporation (Klahn 1990). (b) Arrhenius plot of in-situ measured conductivities for three
days for amorphous 80 nm thick Tb-FeCo films prepared by rf diode sputtering (open circles) and
evaporation (solid circles) (Klahn et al. 1988). Annealing of the R-T alloys in fig. (a) was performed in
dry nitrogen-oxygen atmosphere at 383 K.
398 P. HANSEN

6. Magneto-optical properties

6.1. Phenomenological theory

The interaction of polarized light with a magnetic material leads to a change in the
state of polarization depending on the direction of magnetization. The magneto-
optical effects originating from these light-induced electronic transitions associated
with the spin-orbit coupling can be observed either in transmission (Faraday effect)
or in reflection (Kerr effect). In both cases, the incident linearly polarized light
interacting with the magnetic material is transformed into elliptically polarized light.
This can be described phenomenologically in terms of a complex rotation ~b(M)=
- ~ b ( - M ) resulting from different velocities v+_ =c/N+_ of right and left circularly
polarized light in the magnetic medium. N÷ and N_ are the complex refrective
indexes

N_+ =n_+ -ik_+, (81)

where k+ and k_ are the extinction coefficients which are related to the optical
absorption by k+_ = (2/4n)a+. + and - refer to right and left circularly polarized
light, respectively. The specific Faraday rotation and ellipticity, and Kerr rotation
and ellipticity then are defined by

1 1
OF= ~ Re(q~v), OF = ~ Im(~bF), (82a)

OK= Re(qSK), ~K = Im(CkK), 82b)

where

fD
~bF= 2c (N_ -- N+ ), (83a)

i(N_ - N+ )
~bi~= (83b)
N+N_ - 1

The complex refrective index is related to the dielectric tensor e(M) or the optical
conductivity tensor tr(M) which are correlated by the expression

~(M) = 1 - 4~i ~(M). (84)


(D

For M II z, the conductivity tensor reduces to the form


\
trxx tTxy 0 \
,r(M) = -- axy ff xx 0 ). (85)
0 0 azz
MAGNETIC AMORPHOUS ALLOYS 399

From MaxweU's equations, one obtains

N~ = 1 - 4z~(iaxx + a~,). (86)


co
Combining eqs. (81)-(86) and using the real and imaginary parts of the complex
..(1).T ia!z) with i, k = x, y, z, the rotations and ellipticities
tensor components a,k = "~k
are given by the expressions
2n (1) (2)
OF = c(n2 "-}-k2 ) (na~ r - kaxr ), (87a)

2re
O, , 2 . - 2 , ( k a ~ ) + na~)), (87b)
ctn + K )

and
4 ~ k ( - k 2 + 3 n 2 - 1 ~ a (1)+n(n 2 3k 2 --_(2)
OK = _ _ ' xy - 0o~y - (88a)
co n2(n 2 3k 2 1)2+k2(k 2 - 3 n 2 - 1 ) 2 '
- -,.,(2)
47rn(n 2 - 3k 2 - 1)a(x~) - k ( - k 2 + 3n 2 - l),.xy (88b)
~ti(= co n2(n 2 -- 3k 2 - 1) 2 + k 2 ( k 2 - 3 n 2 _ 1) 2 ,

where n = ½ (n + + n_ ) and k = ½(k + + k_ ). Sign, magnitude and spectral dependence


of the Faraday and Kerr rotation and ellipticity thus are controlled by the off-
diagonal components of g which are determined either by interband or intraband
transitions and by the optical constants. These band transitions in crystalline metals
and alloys give rise to broad 0F or OK spectra (Krinchik and Artem'ev 1968, Erskine
1975, Katayama and Hasegawa 1982, van Engelen and Buschow 1986), in contrast
to oxides as, e.g., bismuth-substituted garnets (Hansen and Krumme 1984) or lantha-
nides and chalcogenides (Reim and Schoenes 1990). This is demonstrated in figs. 86
and 87 showing the spectral dependence of coOxy-(2)(Krinchik and Artem'ev 1968) and
OK(Weller et al. 1988b). The major features of these spectra are even less pronounced
for structurally disordered metals (Prinz et al. 1981).
Experimentally, the tensor elements or complex rotations can be obtained from
ellipsometric measurements (Clemens and Jaumann 1963, Prinz et al. 1981, Allen
and Connell 1982). Ov data have the advantage to represent the bulk properties.
However, the high optical absorption restricts Ov measurements to thin films and
also corrections due to multiple reflections have to be taken into account. OK data
are affected by the surface properties and for film thicknesses below 100 nm also by
the optical constants of the substrate. In this case, significantly enhanced OK values
can be measured (Weller and Reim 1989b).

6.2. M i c r o s c o p i c models

6.2.1. Interband transitions


The magneto-optical properties for energies above 1 eV are mainly governed by
(1) and ~,xr
interband transitions. Their contributions to axr ,,(2) can be expressed in terms
400 P. H A N S E N

•... Ni

3 .': ,,zCo

2~ ." I \X~J
" \.~', ~ /Fe
/ - 1 "U.. ..Jr...... .\../.,....

ti-! "J

0 1 2 3 4 5
Photon energy (eV)
Fig. 86. Absorptive component of the magneto-optical conductivity of Ni, Co and Fe versus photon
energy (Krinchik and Artem'ev 1968). The vertical scale is correct for Ni. The scale for Co and Fe have
to be multiplied by a factor of five. The chain line represents the contribution from intraband transitions.

-0,
COl- x F e x

-05

A -(?.2 d =0.s2 Z
x*x t:r..,,,,x~,~,Jm,~
n~J

~ -o.4 ,- ×=oT
-O.i
%.."
-O,i

I 1 1 I I
-0.7 1 2 3 4 5 6
Photon energy (eV)
Fig. 87. Room-temperature Kerr rotation for different polycrystalline Co 1_xFe~ films (Weller et al. 1988b).

of the absorption rates for left and right circularly polarized light by (Bennett and
Stern 1965)
(1) 7~e2 m2- - Me+
cr~y
- ~ ...-5---75.2,
2 h o g m 2 V s ~,~, ¢O~p - - 09
(89a)
MAGNETICAMORPHOUSALLOYS 401

7~e 2
(2) ~ (M z_ - MZ+)a(co,~- co), (89b)
a~r = 2hcom 2 V~ ,,p

where M+ = (//In+ I~> and the operators rt+ = rt~ _+i~r are linear combinations of
the electron kinetic momentum operator defined by
h
n = p + 8-~5mc
z [s x gV(r)]. (90)

V, is the sample volume and VV(r) is the electric field controlling the motion of the
electron with spins s and momentum p. [e> and [fl> are linear combinations of wave
functions with spin-up and spin-down states and the summation in eq. (89) extends
over the occupied and unoccupied states. The second term in eq. (90) represents the
spin-orbit contribution.
Different approaches for the calculation of-'")
~'ik and t~ik
~ ( 2 ) were reported (Wang and

Callaway 1973, Erskine and Stern 1973a, Singh et al. 1975, Laurent et al. 1979,
Misemer 1988). In addition to the many features of the band structure, the magnitude
of the spin-orbit coupling and the spin polarization are the significant parameters
determining the magneto-optical effects. Some features of the measured spectra were
reproduced by the theory for the crystalline metals (Fe, Ni, Co, Gd), however, parts
of the spectra show a less satisfactory agreement between theory and experiment.
Also relativistic band structure calculations were performed yielding only minor
changes with respect to the nonrelativistic results (Ebert et al. 1988).

6.2.2. Intraband transitions


Normal electron scattering and skew scattering processes are responsible for the
intraband transitions (Erskine and Stern 1973a,b, Voloshinskaya and Fedorov 1973,
Voloshinskaya and Bolotin 1974, Reim et al. 1984). They contribute to the low-
energy part of the magneto-optical spectra. The frequency dependence originating
from these processes yield (Erskine and Stern 1973a).
cog f Q Fico(Y + i@-]~
ax,=-g(a,>~f~+t/k ~j jj, (91)

where f(co) = g22 + (y + ico)2. co2 = &re2N/m , and <az > are the plasma frequency and
the spin polarization, respectively. N and m* denote the concentration and the
effective mass of conduction electrons, respectively. O denotes the skew scattering
frequency and ~ = 1/7 is the normal scattering lifetime, t/ is proportional to the
strength of the spin-orbit coupling. In the high-frequency limit, defined by (2)> 7 and
(2)>>g2, eq. (91) yields

t- '-rx( 1y ) - - -- ~/']<O'z> (92a)

(A) 2
~xrn
(2) -- - ~ ~/(a= > 7(2) (92b)
402 P. HANSEN

These expressions indicate that the skew scattering process is negligible at high
photon energies. The normal scattering process is proportional to the spin-orbit
coupling and the spin polarization. Therefore, the contribution of d electrons is more
significant than that of p electrons due to the large exchange spitting of the
d-electron bands and their relatively large spin-orbit splitting. At high energies, the
contribution of this process to axy is expected to be small and, in particular, coa~Zy) is
independent of ~o (chain line in fig. 86) in contrast to interband transitions.

6.3. One-subnetwork alloys

The crystalline metals Fe, Co, Ni are characterized by a negative Kerr rotation and
a positive Faraday rotation (Clemens and Jaumann 1963, Krinchik and Artem'ev
1968, Buschow et al. 1983, Buschow 1988, Weller et al. 1988b), except for Ni,
exhibiting a sign change below 1 eV and above 4 eV. In all cases, the typical features
of the spectra are determined by two minima (maxima) of OK (0v) located between
1 and 2eV and 4 and 5eV as shown in fig. 87 for Fe, Co and Feo.szCoo.48 (Weller
et al. 1988b). Intra 3d alloys like Col-xFex reveal an increase of the magneto-
optical effects reaching a maximum around x = 0.5. This strong variation of 0,: with
x indicates a significant change of the spin polarization which is in accordance with
the Slater-Pauling curve (fig. 26) showing a maximum of the magnetic moment
around x ~ 0.65. Also, a shift of the OK minimum towards higher energies was
observed which might be attributed to a change in the majority d-state excitations
(Weller et al. 1988b). The magneto-optical properties of crystalline alloys and inter-
metallics which have been discussed in a recent review (Buschow 1988) reveal a
reduction of the main peak between 1 and 2 eV as compared to the pure elements
and the typical features of the spectra are less pronounced (van Engelen and Buschow
1986, Buschow et al. 1983, Buschow 1988). Some s or p metals like A1, Ga and Sn
affect primarily the high-energy side of the spectrum for Co-based alloys. Pt and Pd
enhance also the magneto-optical effects at high energies which was attributed to
spin polarized 4d and 5d electrons for Pd and Pt, respectively. The larger effect of
the Pt was associated with the spin-orbit coupling of the Pt being higher than that
of Pd (Buschow et al. 1983, Buschow 1988).
High magneto-optical effects were found in some Mn-based alloys such as MnBi
(Chen et al. 1973), MnSb (Sawatzky and Street 1971, Buschow et al. 1983) and
PtMnSb (van Engen et al. 1983, Inukai et al. 1986, Ohnuma et al. 1988). The latter
reaches a OK value of --1.27 ° at 1.72 eV (720 nm) and room temperature. This large
Kerr rotation was assigned to the unusual band structure of this half-metallic
ferromagnet (de Groot et al. 1983, 1984, de Groot and Buschow 1986). However, for
most other compositions, an interpretation of the magneto-optical spectra in terms
of band calculations was not yet available.
Amorphous alloys principally show a similar spectral dependence of OK and OF as
compared to their crystalline counterparts. However, for many alloys, the rotation
is reduced and the structure of the spectra is much less pronounced due to the
influence of the structural disorder on the spin-up and spin-down bands and on
the magnetic order. The former influence was interpreted in terms of a change
MAGNETIC AMORPHOUS ALLOYS 403

of the balance between spin-up and spin-down transitions due to the loss of the
k-selection rule of vertical transitions (Weller and Reim 1988). This causes a reduction
of the rotation and a broadening of the transitions. The latter effect was shown to
play an important role for various Fe-rich alloys where competing positive and
negative exchange interactions lead to a drastic reduction of Ms and Tc with increas-
ing Fe content.
One-subnetwork amorphous alloys were not investigated very intensively in con-
trast to R - T alloys. The Kerr rotation for B-Fe glasses are shown in fig. 88 for air-
incident (curve B) and substrate-incident (curve A) light (Buschow and van Engen
1981a). The two rotations are related via the substrate refractive index. The OKvalues
for the amorphous alloys appear to be slightly below the crystalline data (solid
circles). The decrease of OK for the Fe-rich alloys that is not observed from the
magnetization of B-Fe glasses possibly indicate a small deviation from the collinear
structure. This effect was shown to occur more pronounced for Zr-Fe (fig. 29b), Ce-
Fe (fig. 30) or Y-Fe (fig. 31) alloys with respect to the magnetic properties and,
therefore, these alloys are expected to exhibit much lower rotations as compared to
their crystalline counterparts. The compositional and spectral variation of OKfor M -
Fe compounds (Buschow et al. 1983) and the corresponding amorphous alloys
(Buschow and van Engen 1981a) behave similar for M = Si, Ge or Sn and follow the
trend observed for Ms. This was also observed for amorphous F - F e (Sugawara et
al. 1989). The opposite behavior where OK is larger in the amorphous state than in
the crystalline state was reported for Si-Fe (Afonso et al. 1980).
The magneto-optical behavior for Co-based alloys reflects the compositional and
temperature dependencies of M s which are based on a strong ferromagnetic coupling.
The room temperature Kerr rotation is shown in fig. 89a for B-Co glasses at 2 =
633 nm (1.95 eV) and 830 nm (1.49 eV). The full line represents the corresponding Ms
variation. The values for Si-Co were found to agree with the B-Co data within

-0.6

-0.5

-0.4
0J
-0.3

-0.2

-0.1

I I I I I
°0 4 0.6
X
0.8 1.0

Fig. 88. Compositional variation of the room temperature Kerr rotation at 2 = 633nm (1.95eV) for
amorphous B-Fe alloys prepared by vapor quenching (Buschow and van Engen 1981a). Curve (a)
represents 0~ data measured for substrate-incident light and curve (b) for air-incident light. The solid
circles indicate OK values for crystalline materials.
404 P. HANSEN

-0.5
150
-0.4
-0.3 100 A

-0.2
50 <
-0.1 b

0
(a)
I I I I
0.1
o16 x
o18 1.0
-0,5

-O.L
Ti 1-xCox o° o f
~ -0.3

(I~ -0.2

-0.1

0 I I I I I I I I
400 800 1200 1600
[v~(kA/m)

-0.5
-0.4
Zrl-x Cox o°3°o~O
"~ -0.3
(~ -0.2
-0.1

i 4001 I 8001 J 12100 I 16100


Ms(kA/m)
Fig. 89. (a) Compositional variation of the room temperature Kerr rotation at 2 = 633 nm (1.95eV) (open
circles) and ~.= 830nm (1.49eV) (solid circles) for amorphous B-Co alloys prepared by evaporation
(Buschow and van Engen 1981a).The data were measured at the substrate interface. (b) Room temperature
Kerr rotation at 2=780nm (1.59eV) versus saturation magnetization for amorphous Ti-Co and (c)
Zr-Co alloys prepared by sputtering (Honda and Yoshiyama 1988b). The data were measured at the
air-film interface.

experimental error (Buschow and van Engen 1981c). This behavior is confirmed for
a m o r p h o u s T i - C o and Z r - C o alloys ( H o n d a and Yoshiyama 1988b). These results
are displayed in figs. 89b and c, demonstrating that the r o o m temperature OK at 2 =
780 n m (159 eV) scales well with Ms. In the vicinity of zero magnetization, a deviation
from linearity is expected since the reduction of b o t h Ms and Tc cause a decrease of
0i~. Also, OK data for a m o r p h o u s Y - C o alloys (Choe et al. 1987) fit into the OK versus
MAGNETIC AMORPHOUS ALLOYS 405

x variation shown in fig. 89a. Crystalline Cr-Co alloys exhibit also a linear variation
with Ms (Tsutsumi et al. 1983, Honda and Yoshiyama 1988b).
The magneto-optical behavior of the Co-based alloys thus can be interpreted in
terms of a weakly structured spectral dependence typically like that reported for
amorphous Y-Co alloys (Choe et al. 1987). The nonmagnetic diluting elements
primarily tend to decrease the spin polarization and the strength of the exchange
coupling resulting in a linear decrease of Or: with x or Ms except for temperatures
close to Tc. Deviations between the compositional dependence of 0~: and Ms were
reported for Mg-Co (Buschow and van Engen 1981b).

6.4. Two-subnetwork alloys

The magneto-optical activity of rare-earth (R)-transition-metal (T) alloys originates


from the R and T contributions. The heavy rare earths reveal strong transitions
between 4 and 5 eV which is shown in fig. 90 for crystalline RFe2 compounds
(Katayama and Hasegawa 1982). Gd exhibits a positive Kerr rotation in the entire
spectral range investigated, with a peak around 295 nm (4.2 eV) due to interband p -
d transitions (Erskine and Stern 1973a). Tb, Dy, Ho and Er exhibit a strong negative
peak around 280nm (4.4eV). This feature is much less pronounced for the corre-
sponding amorphous alloys (Togami et al. 1983, Y. Suzuki et al. 1987). These
transitions were assigned to interband 4 f ~ 5d transitions (Erskine and Stern 1973a).
The ferrimagnetic order of the heavy R-T alloys leads to a sign change of OKexcept
for Gd where the Gd and T contributions add due to the opposite sign of the Gd
sublattice rotation. This implies that the R and T contribution can be roughly
considered as independent, provided no common bands are formed between the 4f,

Photon energy (eV)


5.0 4.0 3.0 2.4 2,0 1.8
I I I I I f

0.2 .~ /GdFe 2 ~.~-.:~


/ \

(I~ \ ~ ,/~ ~HoFe2


-0,2 , \ // ~ D y F e 2

-0.4 ~ Z b F e 2

I I I I I
20( 300 400 500 600 700 800
~k(nm}
Fig. 90. Wavelength dependence of the room-temperature Kerr rotation for crystalline RFe2 compounds
(Katayama and Hasegawa /982).
406 P. HANSEN

5d R electrons and the 3d T electrons. This was confirmed for amorphous G d - F e


and G d - C o alloys where the sublattice rotations were estimated from the Kerr
rotation of amorphous Gd-Ni, Y-Fe and Y-Co alloys (Tsunashima et al. 1989).
The spectral dependence of amorphous R - T alloys with R = Gd or Tb and T = Fe,
Co or FeCo is shown in fig. 91. The magnitude of the room temperature 10KI for
Gd0.e4Feo.a7Coo.39 is almost reduced by a factor of two (fig. 91a) as compared to
the data for crystalline Feo.50Coo.50 (Weller et al. 1988a,b). Part of this reduction is
due to the reduced Fe and Co moment and the lower Tc in the R - T alloys, but the
remaining reduction in 10KI originates from the influence of the structural disorder
on the spin-up and spin-down transitions. The comparison of the amorphous binaries
G d - F e and G d - C o (Hansen et al. 1989) (fig. 91b) with the crystalline pure elements
(fig. 87) indicate a lower reduction in 10KI. In fig. 91b, the absolute value was plotted
to eliminate the influence of the magnetic compensation on the sign of 0K. This figure
also shows the influence of the temperature on 0K which is much more pronounced
for the R-Co alloys than for the R-Fe alloys due to the much stronger shift of Tc
with the Co content.
The light R - T alloys behave differently concerning two aspects: (i) their moments
are aligned parallel to the T moment and (ii) their wavelength dependence is different
with respect to sign and spectral variation (Suzuki and Katayama 1986, Choe et al.
1987, 1988, Gambino et al. 1986, McGuire and Gambino 1987b, McGuire et al.
1987, Y. Suzuki et al. 1987, Takahashi et al. 1987, Honda and Yoshiyama 1988a,
Weller et al. 1988a, Tsunashima et al. 1989). The different signs occurring for 0K and
0V in R - T alloys are summarized in table 16 using the common sign convention
(Kahn et al. 1969). This leads to a positive Faraday (Kerr) rotation for the magnetiza-
tion parallel to the propagation of the transmitted (reflected) light when the displace-
ment vector D turns counter clockwise for an observer facing the light. It should be
noticed that the presence of a compensation composition or temperature leads to a
sign change in OK or 0F due to the reversal of the net magnetization in the presence
of an external field. Therefore, this sign change represents no intrinsic material
property but is important when 0K or OF data of the ferrimagnetically ordered alloys
are compared. The room temperature spectral dependence of amorphous N d - C o
alloys is displayed in fig. 92a. The Nd contributes to 0K via f ~ d interband transitions
(Weller et al. 1988a) leading to an increase of 10KI in the short-wavelengths range
(Choe et al. 1989). The low-energy peak of 10KI, originating from Co, decreases
continuously with increasing Nd content due to a reduced Co content and spin
polarization. A further rise of the Nd content causes a strong reduction of Tc resulting
in a strong overall reduction of Tc and the room-temperature spectrum vanishes for
x - 0.38 (Tc = 295 K) (Honda and Yoshiyama 1988a,b). A corresponding influence of
Nd and Pr was observed for Fe-based alloys (Suzuki et al. 1988). Their spectral
variation together with that of other alloys is shown in fig. 92b. A comparison of the
other various spectra reveals that the highest rotations in the visible and ultraviolet
region can be reached with amorphous N d - F e C o and Pr-FeCo alloys which is of
interest for magneto-optical recording operating at short wavelengths. However, the
other rare earths, except for Eu, give rise to very small contributions to axr (1) and
a~2)
x y in the visible due to the large binding energies of the 4f electrons. Also, the
MAGNETIC AMORPHOUS ALLOYS 407

Photon energy (eV)


1 2 3 L 5

-0.1
Tb°'22F e ° ' 6 ~

-0.2

do,2z,Feo.37Co 0.39
-0.3
d

-0.4

-0•5

(al
-0.6 I [ I
1500 700 500 300
~k(nm)
0,8
Gdl_xTx x=xO~
T= LOK ..~ ~"~

0.6 ..... T= 295K Z 5 8 4 )

xJ 0.L

0.2
T=Co
(b)
0 5•0 10~0 15100
(nm)
Fig. 91. Wavelength dependence of (a) the room-temperature Kerr rotation for crystalline FeCo and
amorphous Tb-FeCo and Gd-FeCo alloys (Weller et al. 1988a,b) and (b) the Kerr rotation for amorphous
G d - F e and G d - C o alloys at two temperatures (Hansen et al. 1989).

optical constants are large limiting the attainable Kerr rotation in amorphous R - T
alloys (Weller and Reim 1989a).
The compositional variation of the Faraday rotation at 2 = 633 nm (1.95 eV) for
amorphous R - T alloys with R = Gd or Tb and T = Fe or Co is presented in fig. 93
408 P. HANSEN

TABLE 16
The sign of the Faraday rotation and the polar Kerr rotation in the visible and infrared, assuming the
direction of the net magnetization parallel to the applied field. The sublattice magnetizations are aligned
antiparalM in the case of heavy rare earths (Gd, Tb, Dy) and parallel for light rare earths (Nd, Pr) as
indicated by the arrows.
R~ -xT~ Or OK
R('i~) T(T) x= 0 Xeomp~ X~ 1 X< Xeomp X= 0 Xcomp~ X~ 1 X< Xeomp
Gd Fe,Co - ~ + 6]" - "D~ + 'i) - o]" + ~'~
Tb,Dy Fe,Co + "i) + 0~" - / + "ff$ -- "D - O~" +/-- @~

R T x~0 x=l 0<x<l x=0 x=l 0<x<l


Nd,Pr Fe,Co + Q~ + ~ + ~i)T - ~ - 1" - "~]"

(Hansen et al. 1989). The compositional compensation [Ms(xcomp)=0] for the


different alloys is indicated by the arrows. The low-temperature OF data given in
fig. 93a reveal a dilution of the G d contribution for x < 0.5. In this range of composi-
tions, the Co carries no magnetic m o m e n t and the Fe m o m e n t and exchange is small
and, therefore, both behave like nonmagnetic elements like Cu or A1, indicated by
the crossed symbols. For x > 0.5, the spin polarization and the exchange coupling
strongly increases and the rotations from the R and T elements add, giving rise to
a strong rise of OF. This supports the picture of two independent sublattice contribu-
tions which approximately is valid for this wavelength that is between, and just
sufficiently away from, the centers of the R and T transitions. The strong temperature
dependence of the magnetic and magneto-optical properties gives rise to a significant
change of the compositional variation of OF and OK at high temperature which is
shown in fig. 93b. At the R-rich side, the Curie temperatures passing the room-
temperature limit the rotation. For the Fe-rich alloys, the strong turndown of 0v is
caused by the low Tc of the possibly speromagnetically ordered Fe-rich R - F e alloys
causing the typical m a x i m u m of IOF[ and IOKIaround x ~ 0.75 (Hansen 1987, Hansen
et al. 1987, 1989). A similar variation of OK and OF occurs for alloys with R = Dy or
Ho and T = Fe or Co (Hansen et al. 1991). The m a x i m u m disappears when Co is
added, which raises Tc due to the high C o - F e exchange coupling. A comparison of
the high- and low-temperature data suggest that the R contribution is of minor
importance at high temperatures. This is confirmed for various amorphous R - C o
alloys where the room temperature OK at 2 = 633 nm (1.95eV) was shown to scale
well with the Co sublattice magnetization inferred from mean-field results (Honda
and Yoshiyama 1988b). These results are displayed in fig. 94 for R -- G d or Dy. A
plot of 0K versus x for R = Y, Gd, Tb or Ce also reveals an almost R-independent
room-temperature rotation (Choe et al. 1987).
The temperature dependence of the Faraday and Kerr rotation is determined by
eqs. (87) and (88) where axy _(2) are proportional to the spin polarization (az)
(1) and oxr
for both interband and intraband transitions. For G d - T alloys, the magneto-optical
transitions from the G d and the transition metal can be assumed to be independent
because of the 4f and 5d G d electrons and the 3d T electrons form no common band.
Then, ( a z ) is proportional to the sublattice magnetization and the combination of
MAGNETIC AMORPHOUS ALLOYS 409

4~'u+u ~ ~'=~++~ ~_+++++ Ndl-x Cox

0 "~:F+I-
-0.1 /
x=0.66 !

°°°%°Oo.......~0.75 ~p
-0,2
oJ

-0.3 ;;; -'. Z / • o

-0.4
../
(o)
-0.5 I I I I I
1 2 3 4 5
Photon energy (eV)

0.1

-0.1
~ ~ ....... ~DY 0.22FeoTa
. . . . Tbo.15 Nd020 FeoJ,5sCo 0.195
-0.2 - ~-.~.~---~'='~ ~Tbo.15 Ndo.15 Fe0.595Co 0.105
;,~--"~-. . . . . . . . . . . . . . . . . . . . . -~.~Tbo.22 Feo.Te
qc~ ~.-..~ . . . . . . _ . . . . --.,~Tb0.,~ Nd0.,0(FeCol ....
-0.3
"

; \ "~."~,~. ,,,,,~ Ndo.z,oFeosl Coo.og


'\ • ~ ,,,,,,.,,,," __ Tbo.ts (FeCo)o.Bs
-0.4 j ~.I " ~ . ~ .~Pr o.~,oFeo6o
'Its"/ .,I'./ Ndol.o
• Coo15Feoz=5
-0.5 = f
./
-0.6
:b)
".~..~-'I"
-0.7 I I I I
300 500 700 900
7, (nm)
Fig. 92. Room-temperature Kerr rotation for (a) amorphous N d - F e alloys versus photon energy (Weller
et al. 1988a) and (b) for various amorphous R - T alloys versus wavelength (Suzuki et al. 1988).

eqs. (87), (88) with eqs. (89) or (91) yields


Ov = AMod(T) + BMTM(T), (93)
and a corresponding equation for OK. The magneto-optical coefficients A and B in
principle are determined by eqs. (87)-(91), but this requires an evaluation of the
410 P. HANSEN

T=Z,.2K /,
A
E

-o 3
%

co,
~.Z~,. •

• ". Xeomp

I I I I
0,5
I "", I
1.0
3.5
(b)

o •
3.0 T=295K

2.5

"E~ 2.0 Gdl-x Fex"-~..~..l~,.-"


Tbl_x F e x , ~ ( I~ ~;~

,o ;// / I I I II

# / ~ /'Tb,_xCo, I!
I I I'l x°o,,, II
i /" / /TbFe--. /TbCo II

0.3 0.L 0.5 0.6 0.7 0.8 0.9 1.0


x
Fig. 93. Compositional variation of the Faraday rotation/or binary R - T alloys with R = Gd or Tb and
T = Fe or Co at (a) T = 4.2 K and (b) T = 295 K (Hansen et al. 1989). (©) Gd-Co, ([]) Od-Fe (Hansen
et al. 1989), (0) Gd-Co (Gambino and McGuire 1986), (n) Gd-Fe (Gambino et al. 1986), (1) Tb-Co
(McGuire and Hartmann 1985) ( ~ Gd-Cu, (z~) Gd-A1 (McGuire and Gambino 1987a) in fig. (a); in fig.
(b), (©) Gd-Co, ([]) Gd-Fe (Hansen et al. 1989), (0) Tb-Fe (Urner-Wille 1981), (1) Tb-Co (Heitmann
et al. 1985). The arrows indicate the compensation composition [eq. (59)] of the respective alloys.
MAGNETIC AMORPHOUS ALLOYS 411

0.5 (a)

0.4 Gd-Co oZ

"~ 0,3
n~

O__~0.2

0.1

0
o I I
400
I I
800
I

Mco (kA/rn)
I
1200
I i
1600

0.5
(b)
0.4

0,3

0.2

0.1

I i I I I I
0
0 400 800 1200 1600
Mco (kA/rn)
Fig. 94. Room-temperature Kerr rotation at 2 = 780rim (1.59eV) versus cobalt sublattice magnetization
Mc, for amorphous Gd-Co and Dy-Co alloys (Honda and Yoshiyama 1988b). Me, was inferred from
the fit of the mean-fieldtheory to the measured saturation magnetization.

respective matrix elements. This problem has not yet been solved and thus A and B
have to be regarded as adjustable parameters.
The measured temperature dependence of OF at 2 ----633 nm (1.95 eV) was investi-
gated for different amorphous G d - T alloys (Hansen and Urner-Wille 1979, McGuire
and Hartmann 1985, Hansen and Hartmann 1985, Hansen et al. 1989) and is shown
in fig. 95 for amorphous G d - F e and G d - C o alloys. Both systems exhibit a sign
change of Ovat Xoomp,resulting in a different sign of Ovfor Gd-rich and T-rich alloys.
The full lines were calculated from eq. (93), where the sublattice magnetizations were
inferred from the fit of the measured saturation magnetizations. The good agreement
achieved between measured and calculated rotations with temperature-independent
magneto-optical coefficients suggests again that the assumption of independent sub-
lattice rotations is a reasonable approach for these alloys. The evaluated sublattice
rotations confirm the dominance of the Fe and Co sublattice magnetization at this
wavelength and for high temperatures. At low temperatures, the Gd contribution is
essential and dominates for Gd-rich alloys as is obvious from fig. 93a. The noncollin-
ear moment distribution of the non-S-state rare earths leads to a reduced sublattice
rotation as reported for amorphous T b - C o alloys (McGuire and Hartmann 1986).
Various attempts were made to improve the magneto-optical properties by small
additions of further elements. The influence of rare earths and transition metals on
412 P. HANSEN

Gdl-x Fex
3 ~ 8 5 - - ca[culoA-ed

[] [] []

0 I I I I | II
/ 100 200 T(K, 3 ~ 400 500

_11_

2
Gdl_xCOx
- - colcutated

o T(K)
100 200 300 400 500 600 700
I i i i I I I

-2
"0.75
(b)

Fig. 95. Temperature dependence of the Faraday rotation at 2 = 633 nm (1.95 eV) for amorphous (a) G d -
Fe and (b) G d - C o alloys (Hansen et al. 1989). The full lines were calculated in terms of the sublattice
magnetizations [eq. (93)] inferred from mean-field results.

the room temperature OK at 2 = 633nm (1.55eV) is demonstrated in fig. 96 for


amorphous Tb-Fe-based alloys (Imamura et al. 1985). The influence of the R addi-
tions can be attributed to changes of Tc except for Nd or Pr, giving rise to additional
contributions via f ~ d transitions. Many transition metals, such as Ti, V, Cr, Mn
and Cu, and metals (A1 and Au) tend to decrease Tc and the spin polarization
resulting in a low OK and 0r (Tsujimoto et al. 1983, Aratani et al. 1985, Imamura et
al. 1985, McGuire and Hartmann 1985, Hansen and Hartmann 1986) while Co raises
Tc and affects the band structure (figs. 60 and 87). Other elements, like Bi, Pb, Sn
MAGNETIC A M O R P H O U S ALLOYS 413

0,30
Tb,Fe)l_x Rx + R=Gd
0.25

~ 0.20
.8
0.15

0.10

0.05
Ill) I I r I
0

0.30

0.25
, v .dc u
"-~ 0.20
.8
~:~ 0.15

0.10

0.05
(b) p , I I
0.02 0.04 0.06 0.08 0.1o
X

Fig. 96. Room-temperature Kerr rotation at 2 = 633 nm (1.95 eV) for amorphous Tb-Fe-based alloys with
small additions of (a) rare earths and (b) transition metals (Imamura et al. 1985).

(Hansen and Urner-Wille 1979, Urner-Wille et al. 1980, Hartmann et al. 1984b,
Masui et al. 1984) or In (Iijima et al. 1989) give rise to a small increase of the
magneto-optical properties. Also Ni, Pt (Imamura et al. 1985) and U (Dillon Jr et
al. 1987) were studied with respect to their magnetic and magneto-optical properties.
The influence on the structural disorder and composition by the preparation
parameters gives rise to significant changes of OK and 0V (Sato and Togami 1983,
Tsujimoto et al. 1984, Shieh and Kryder 1985, Heitmann et al. 1987b), as expected
from the corresponding influence on the magnetic properties.

7. Transport properties

7.1. Resistivity and magnetoresistance

Structurally disordered alloys exhibit much larger resistivities than crystalline mater-
ials due to different scattering processes reducing the conduction-electron mean free
path. The resistivity values for amorphous alloys range typically from 100 to
300 g~ cm. Data for some M-T, R - M and R T alloys with M = Cu or Au and T =
Fe, Co or Ni are compiled in tables 17-19. Experimental and theoretical work was
discussed in different reviews (Campbell and Fert 1982, Buschow 1984a).
The resistivity tensor p couples the components of the electric field E and the
414 P. HANSEN

TABLE 17
Resistivity of some amorphous alloys.
Alloy Magnetic T(K) p(gf~ cm) References
order
Sio.2oPdo.so Diam. 4.2 76 K/ister et al. (1980)
Lao.76Gao.24 Diam. 4.2 219 Shull et al. (1978)
8no.21Feo.79 Ferrom. 4.2 120 McGuire and Hartmann (1985)
Sno.59Feo.41* Ferrom. 200 235 Geny et al. (1982)
Bio.14Feo.s6 Ferrom. 4.2 227 McGuire and Hartman (1985)
Gdo.26Feo.74 Ferrim. 4.2 274 McGuire and Hartman (1985)
Sno.szCoo.48* Ferrom. 200 265 Geny et al. (1982)
Gdo.67Coo.33 Ferrim. 4.2 280 Durand and Poon (1977)
Po.26Nio.v6 Param. 4.2 162 Cote (1976)
Tio.52Nio.4s* Param. 4.2 325 Buschow (1983)
Yo.oTNio.93 Ferrom. 4.2 86 Cochrane and Str6m-Olsen (1978)
Sno.6oNio.4o Param. 200 140 Geny et al. (1982)
Gdo,68Nio.32 Ferrom. 4.2 197 Durand and Poon (1977)
Gdo.55Auo.4s Ferrom. 4.2 288 McGuire and Hartmann (1985)
GdsoAuo.2o Ferrom. 4.2 312 Cote (1976), Durand and Poon (1977)
* Composition at maximum resistivity.

TABLE 18
Curie temperature Tc, saturation magnetization M s at T = 4.2 K, electrical resistivity, p and Hall resistivity
Pn at T = 4.2K for amorphous R Au and R-Cu alloys (McGuire and Gambino 1979, McGuire et al.
1980a).
Alloy Tc Ms p(p~ cm) PH
(K) (kA/m) T = 4.2 K T = 297 K (gf~ cm)

Pro .47Auo.s3 < 4.2 127 167 168 0.96


Ndo,385Auo.615 10 159 230 234 2.24
Smo,43Auo.57 48 173 177 0.83
Euo.2oAuo.so < 4.2 302 213 198 0.30
Gdo.537 Auo.463 99 401 205 210 - 5.66
Tbo.47 Auo.53 > 4.2 748 190 197 - 1.27
Dyo,46 Auo.s4 15 812 132 138 - 0.40
Hoo.4.,Auo.56 11 74 93 95 -0.30
Ero.s8 Auo.42 9.5 787 191 199 - 0.03
Tmo.62Auo.38 636 194 207 --0.09
Pro .4oCuo .6o < 4.2 183 86 82 0.43
Ndo.44 Cuo. 56 < 4.2 238 114 116 2.07
Gdo.42 Cuo.58 77 1180 176 175 - 1.96
Tbo.503Cu0.497 23 621 147 151 - 0.26
Dyo,4s Cuo.55 16 1020 141 145 0.15
HooA.4Cuo.56 8 939 220 0.45
Ero.s2 Cuo.49 < 4.2 668 176 178 0.52
Tmo.61 Cuo.39 812 149 159 0.31
MAGNETIC AMORPHOUS ALLOYS 415

TABLE 19
Electrical resistivity and Hall resistivity Pn for amorphous R-T alloys.
x T p P, Ref.*
(K) (g~ cm) (gfl cm)
Y1 - xFex 0.71 77 226 4.22 [1]
Y~ _xCox 0.75 77 4.4 [2]
0.67 77 244 3.3 [1]
Yz _~Nix 0.936 4.2 140 -1.42 1-3]
0.83 4.2 90 -0.33 [3]
Gd 1_xFe~ 0.74 4.2 274 - 11.5 [4]
0.79 77 195 11.25 [1]
0.734 77 211 - 11.6 [1]
(Gdo.26 Feo. 74)0.89 BioA1 4.2 400 -22.0 1-4]
Gd t _~,Co:, 0.84 77 141 4.1 [1]
0.82 77 162 -4.6 [1]
0.79 77 250 -5 1-5]
GdoA45 Coo.72MooA] 77 172 -1.4 I-1]
Gdo.19Coo.7oAuoAo 77 138 -4.3 [1]
Gda _xNi~ 0.84 4.2 152 0.88 I-6]
0.64 4.2 400 0.37 1-6]
0.59 4.2 197 -0.19 [6]
0.40 4.2 245 -2.28 1-6]
Tba _xFex 0.75 295 200 11.0 [7]
Hol _xCox 0.67 77 -3.2 [2]
* References:
[1] yon Molnar et al. (1981). [5] Shirakawa et al. (1976).
1-2] Asomoza et al. (1977b). 1-6] McGuire and Gambino (1978).
[3] McGuire and Gambino (1978). 1-7] Malmh/ill (1983).
[4] McGuire and Hartmann (1985).

current density J by E = par. The components of p depend on B = # o ( H + M) due


to the Lorentz force and scattering mechanisms. For an isotropic medium in a
magnetic field applied to the z direction, p can be expressed in the form

P= H Pi 0 , (94)
0 Pll

where P ll and p± are the resistivities for the electrical current applied parallel and
perpendicular to the direction of magnetization. Both can be split into a spontaneous
or extraordinary part and an ordinary part. The off-diagonal element PH is the
spontaneous or extraordinary Hall resistivity. The average resistivity for an dis-
ordered material than is given by
j0 = ~Pll-t-
1 ~z p ± , (95)

where iS(T, 0) = p(T).


The magnetoresistance describes the change of the resistivity by the applied m a g -
416 P. HANSEN

netic field. The isotropic and anisotropic part of the magnetoresistance thus can be
expressed by
Api(T,/-/) = ~(T, H) - ¢5(T,0), (96)
Apa(T, H) = [Pll (T, H) - p±(T, H)]. (97)
Apart from the high resistivity, some further characteristic features of amorphous
alloys were observed. A very low temperature dependence of p at H = 0 was found
with Ap/p(O)= [ p ( T ) - p(O)]/p(O) below 10% for 0 ~ T ~<300 K. Many alloys exhibit
a low-temperature resistivity minimum with a large logarithmic upturn as shown in
fig. 97a for PBA1-Fe and PBA1-FeNi glasses (Rapp et al. 1978). This temperature
behavior was also found for FeNi-based alloys (Babi6 et al. 1978, Steward and
Phillips 1978, Rao et al. 1979), Co-based alloys (Cochrane et al. 1975, Marzwell 1977,
Rao et al. 1979) or Ni-based alloys (Cochrane et al. 1975, Berrada et al. 1978,
Cochrane et al. 1978a,b). It is very similar to that observed for Kondo systems
involving spin-flip scattering of the conduction electrons on localized magnetic
inpurities as demonstrated for amorphous Pd-Si alloys containing Cr, Mn, Fe or
Co impurities (Hasegawa and Tsuei 1971a,b). However, the presence of this upturn
in nonmagnetic amorphous alloys like La0.66Alo.34 (Mueller et al. 1980) and the field
independence of the logarithmic resistivity found for many alloys suggest that this
behavior originates from a nonmagnetic mechanism. A model treating the electron
scattering in terms of a two-level system (Cochrane et al. 1975, Tsuei 1978) or by
tunneling (Anderson et al. 1972) leads to a temperature dependence of p of the form

p(T) = p(O) + c In 1 + , (98)

where A is the mean value of the energy separation between the two levels. This
logarithmic temperature dependence predicted by eq. (98) is confirmed in several
amorphous alloys such as La-A1 (Mueller et al. 1980), P-Co, P-Ni (Cochrane et al.
1975, Cochrane and Str6m-Olsen 1977, Berrada et al. 1978), Y-Ni (Cochrane et al.
1978a,b) and Tb- or Pb-based alloys (Cornelison and Sellmyer 1983), while the
agreement with experimental results with other theoretical treatments (Kondo 1976,
Black and Gyorffy 1978) yielding a (ln T) 2 dependence is less satisfactory. For
amorphous G d - C o alloys, only the low-temperature range can be described by
eq. (98) (Okuno et al. 1981).
A further feature is a maximum of the resistivity and a minimum of the temperature
coefficient c~= (1/p)dp/dT occurring in the compositional variation. This is shown in
fig. 97b for amorphous Sn-Fe (Geny et al. 1982) and was also reported for Sn-Cu
(Korn et al. 1972), Sn-T with T = Fe, Co or Ni (Geny et al. 1982), Sn-Au (Blasberg
et al. 1979), Ti-Ni (Buschow 1983) and Hf-Ni (Buschow and Beekmans 1979b). In
many alloys, the maximum cannot be observed due to the limited glass-forming
range.
These concentration dependencies of p and c~are commonly encountered in liquid
alloys suggesting that there exists a close similarity in the resistivity behavior between
amorphous and liquid alloys (Gfintherodt et al. 1978, Geny et al. 1982). These results
M A G N E T I C A M O R P H O U S ALLOYS 417

/Po
z~ ~xa'-~axa '16 Bo '06 A[O '03 Feo,75 A
z~
144,4 162.1

00(300000 O0 A A
162.0
1/,/*.3 0000 Az~ z~
" •14.4.2 O~Oo ,xA
oo
",
161.9 "~
:&
o

o.. 1/.4,1
% 161.8 o.,
o

j Oo 2I
14/*.0 Po.15
I B 0.oe A[o.o3 Feo, 60 Ni0.1s S g- 161.7
o o
o oo
(a) %~6
1/.3,9 I I I ' 161.6
0,01 0.1 1.0 10 100
T(K)

250 40
p

20C 20

15C 02"-
,,?
o

o, 100 -20

50
J -40

(b)
I -60
0.5 1.0
X
Fig. 97. (a) Resistivity versus T (logarithmic scale) for amorphous PBA1-Fe and PBA1-FeNi alloys
(Rapp et al. 1978). (b) Compositional dependence of the average resistivity and the resistivity coefficient
at T = 200 K for S n - F e alloys (Geny et al. 1982).

suggest that the theoretical model for liquid metals (Ziman 1961, Faber and Ziman
1965) based on the nearly free electron model and the Boltzmann transport equation
should be extendable to the amorphous alloys (Cote 1976, Nagel 1977, Cote and
Meisel 1977, Esposito et al. 1978). This theory relates p to the resistivity structure
factor which relates its maximum at q = ~ 2kF where q is the electron scattering
vector and kF the Fermi wave vector. This maximum of the structure factor also
determines that of p and the minimum of a. These theoretical considerations are in
418 P. HANSEN

qualitatively agreement with the experimental results and accounts for the different
resistivity behavior of related alloys.
The compositional variation of the resistivity for R-Feo.50Coo.o5 alloys (Weller
and Reim 1989) is presented in fig. 98.
A significant anisotropy of the resistivity was reported for amorphous Au-T alloys
with T = Fe, Ni or Co (Bergmann and Marquardt 1978) and Co-based alloys (Shiba
et al. 1986).
Another reason for the occurrence of a temperature resistivity minimum was
associated with the onset of magnetic ordering and was ascribed to coherent exchange
scattering by the rare-earth spins (Asomoza et al. 1977a,b, Fert et al. 1977). This
model predicts a resistivity
CJpm [ 1
P = J + 1 L + cm(2kF)], (99)

where Pm is the maximum resistivity and c the concentration of magnetic atoms.


m(2kF) is the spin correlation function. The first term on the right-hand side represents
the contribution from the scattering of each atom. The second term is proportional
to the number of pairs and arises from coherent exchange scattering. In amorphous
alloys, only the spin correlations between neighbor sites contribute to m(2kF) and,
thus, the resistivity resulting from this process is controlled by the local magnetic
order. In this case where the magnetic order is of longer range than the structural
order, m(2kF) can be expressed by (Fert and Asomoza 1979)
m(2kF) = [a(2kF)- 1]j 2, (100)

250

[] o [3

200

150
E
o
C:~

o_ 100 []

• Tb
+ Tm .: \

50

h-

i I I !
0,20 0.4.0 0.60 0.80 1,00
x

Fig. 98. Room-temperature resistivity versus composition for amorphous R-FeCo alloys prepared by
sputtering (Weller and Reim 1989a).
MAGNETIC AMORPHOUS ALLOYS 419

where j2 = (Ji" Jk )/j2 describes the local magnetic order and a(2kv) is the structure
factor. The low-temperature resistivity upturn and thus the occurring minimum of
p is caused by the competition between elastic exchange scattering and inelastic spin-
flip processes.
A resistivity minimum due to magnetic ordering was observed, e.g., for amorphous
R-Ni alloys (Fert et al. 1977, Asomoza et al. 1977a, Fert and Asomoza 1979) or U -
T alloys with T = Fe, Ni, Gd, Tb or Yb (Freitas et al. 1988). As an example, the
results for amorphous Dy-Ni (Asomoza et al. 1979a, Fert and Asomoza 1979) are
shown in fig. 99. The resistivity upturn at low temperatures thus can be explained
assuming m(2kv) to be positive, i.e., a(2kv)> 1.
The isotropic part of the magnetoresistance is displayed in fig. 100 for amorphous
Dy-Ni and turns out to be positive in accordance with eq. (100) for a(2kv) > 1 and
j > 0. The much smaller anisotropic part of the magnetoresistance is presented in
fig. 100b. This anisotropy was ascribed to the quadrapole moment of the rare earth
(Fert and Asomoza 1979, Asomoza et al. 1979a). The magnetoresistance was found
to be positive for Dy-Ni, Ho-Ni (Asomoza et al. 1979b) and Ce-Co (Felsch et al.
1982) and negative for Er-Ni (Asomoza et al. 1979b) and DyGd-Ni (Amaral et al.
1988). Amorphous R - U with R = Gd or Tb reveal a sign change of Ap with temper-
ature (Freitas et al. 1988). Below the spin-freezing temperature, an increasing portion
of antiferromagnetic interactions become important, leading to a negative magneto-
resistance according to eq. (100) with (Ji',lrk) < 0 in contrast to amorphous U - F e
with collinear or random ferromagnetism where Ap > 0 due to a positive <Ji. Jk>.
The latter is in agreement with results for amorphous Si-Fe (Shimada and Kojima
1978). Also, a negative magnetoresistance was observed in the spin-glass-like amor-
phous D y - U and N d - U systems (Freitas et al. 1988).

297

296 Dy0.2s N i 0 . 7 / /

o 29/.
295 ~ ~ ~ 6 ~0 kA/m
:zl.
~o.. 293

292 "et~z/"'-. ~ '2/.00 kA/m


•" ~'H= 0 kA/m
~,.+'
291 "" Tc
/

290 ~ q i t i
0 20 /*0 60 80 100
T(K}
Fig. 99. Temperaturedependenceof the resistivityfor amorphous Dy-Ni at differentmagneticfields(Fert
et al. 1977, Asomozaet al. 1977a,b, 1979b).
420 P. HANSEN

2.5
(a}
Dyo.25 Nio.75

2.0

/H=2400 kA/m
1.5 1600 k A / m
C~
640 k A / m

~1.0

0.5

I
10 20 30 40
T(K)

(b) T=I.2K
4.2
D Yo.25Nio.75 7.0
3.0 10.0

15.0

<~2.0
".4"
o • • ~, 25.0

• 30.0
1.0
40.0

' r I I
0 0.5 1.0 1.5 2.0 2.5
H(106A/m)

Fig. 100. (a) Isotropic magnetoresistance versus temperature for amorphous Dy-Ni at different magnetic
fields and (b) anisotropy of the magnetoresistance for different temperatures for amorphous Dy-Ni
(Asomoza et al. 1979a,b, Fert and Asomoza 1979).

7.2. Hall effect

The Hall resistivity is determined by the off-diagonal elements of the resistivity tensor
and is composed of the ordinary and the spontaneous or extraordinary part. PH thus
can be expressed by
PH = #o(RoH + RsMs). (101)
Ro is the ordinary and Rs the spontaneous Hall coefficient. The ordinary Hall effect
arises from the Lorentz force acting on the moving electrons, and the spontaneous
Hall effect is caused by asymmetric scattering of the conduction electrons by the
magnetic atoms via skew scattering (Smit 1955, 1958, Fert and Friedrich 1976) or
side jumb scattering (Berger 1970, 1972, 1973) with contributions proportional to p
M A G N E T I C A M O R P H O U S ALLOYS 421

and p2, respectively. Thus, Rs can be expressed by (Majumdar and Berger 1973)
Rs = ap + bp 2, (102)
where a and b are constants. The spontaneous contribution usually dominates in
amorphous alloys. Typical values for Gd, evaluated from amorphous G d - M alloys
with M = Cu, Au (McGuire and Gambino 1979, Gambino et al. 1981, McGuire and
Gambino 1987a) or M = Ge (Gambino and McGuire 1983) are Ro = -0.03 I~)cm/T
and Rs = 5 ~f)cm/T. For alloys with Ro ~ R~, the Hall resistivity is proportional to
the saturation magnetization, which implies the presence of a Hall hysteresis corre-
sponding to magnetization or magneto-optical measurements. This is displayed in
fig. 101 for amorphous Gd-Au and G d - F e alloys (McGuire et al. 1977), demonstrat-
ing that pn is negative for the Gd spin parallel aligned to the applied field. The study
of amorphous R-Cu and R-Au alloys reveals a positive PH for light rare earths and
a negative PH for the heavy rare earths, which is shown in fig. 102. This suggests that
the spontaneous Hall effect is correlated to the rare-earth spin. In the case of the
heavy rare earths, the spin and orbital moments are parallel aligned to the applied
magnetic field while for the light rare earths the spin moment is oppositely aligned
to the magnetic field due to their antiparallel coupling. This is in accordance with
the side jumb scattering process which is proportional to the conduction electron
polarization caused by the rare-earth spin.
Amorphous Fe-based (Shimada and Kojima 1978, Stobiecki and Kowalski 1984,
McGuire and Hartmann 1985), Co-based (Asomoza et al. 1977b, McGuire et al.
1980a, Shiba et al. 1986) and Ni-based (McGuire and Gambino 1978, McGuire and
Taylor 1979) alloys exhibit a positive spontaneous Hall effect. This is in agreement
with T-rich G d - T alloys as reported for G d - F e (McGuire et al. 1980a, Stobiecki
and Kowalski 1984, Honda et al. 1985), Gd-Co (Ogawa et al. 1975, Asomoza et al.
1977b, Okuno et al. 1981, Stobiecki and Kowalski 1984) and Gd-Ni (Mimura
et al. 1976a, McGuire and Gambino 1978, McGuire and Taylor 1979, Asomoza et


0 O

-4 -2

-8
-1,s -1'.o -o'.s 015 1.~0 1.5
H (106A/m]
Fig. 101. Hall resistivity at T = 4.2 K as a function of applied magnetic field for amorphous G d - A u and
G d - F e alloys (McGuire et al. 1977).
422 P. H A N S E N

12

10

8
f\
6

°52 R-Cu

-2

-4

-6
Ce Pr Nd PmSm Eu Gd Tb Dy Ho Er Tm Yb
R
Fig. 102. Tangent of the Hall angle at T = 4.2 K and H = 1.6 × 106 A/m for a m o r p h o u s R - C u and R - A u
alloys (McGuire and G a m b i n o 1979).

al. 1979b). The results for some of the one-subnetwork alloys identify the side jumb
scattering as the dominant mechanism controlling the spontaneous Hall effect. This
is demonstrated in fig. 103 displaying the room-temperature spontaneous Hall co-
efficient as a function of resistivity for various amorphous M-Co alloys. The logarith-
mic plot reveals a p 2 dependence of Rs (Shiba et al. 1986, Jen and Yang 1988), as
predicted by the side jumb process. Similar results are obtained for amorphous Gd-
Au alloys (Gambino et al. 1981). For two-subnetwork R-T alloys, the resistivity
dependence of Rs indicates that the skew scattering mechanism is also of importance
(Fert and Friedrich 1976, 1977, Lachowicz 1984).
Heavy rare-earth-transition-metal alloys exhibit a magnetic compensation. Thus,
compositions with x < Xcomv show a sign of change of Pn as compared to those with
x > Xcompdue to the reversed direction of the sublattice magnetizations which also
applies when passing through the temperature compensation. These results suggest
that the spontaneous Hall effect can be composed in analogy to the magneto-optical
properties by an R and T contribution leading to the equation
Pn = fro(Roll + RsT MT + RsR MR). (103)
Combining eqs. (10 l) and (103) yields for the spontaneous Hall coefficient of binary
alloys
R~T MT + R~R M R
R~ = , (104)
[M R -- MT[

which indicates that R~ undergoes a singular behavior at the compositional or


temperature compensation and is observed for different R-T alloys, and shown for
MAGNETIC AMORPHOUS ALLOYS 423

/,,
5 m

M1-xC°x ¢
3

1
E 0.5
t3

:&
0.3
/ [] M=Y
• Hf
/'~ o Zr
/ • Ta
0'1~o / A Nb
0.05 / :
0'03 F
I I [ I
10 30 50 100 300
p (/.l,~ cm)
Fig. 103. Logarithmic plot of the spontaneous Hall coefficient as a function of the resistivity at room
temperature for various amorphous Co-based alloys (Shiba et al. 1986). The full line represents a
p2 dependence.

amorphous Gd-Co alloys in fig. 104. The spontaneous Hall effect of R-T alloys thus
can be considered as the sum of the R and T contribution, because in the case of
heavy rare earths the positive T contribution and the negative R contribution add
due to the antiparallel alignment of the R and T sublattice and in the case of light
rare earths both R and T contribution are positive and are also additive due to the
parallel alignment of the sublattices. This behavior corresponds to that of the Faraday
sublattice rotations. This aspect was discussed for various R-T alloys (McGuire and
Hartmann 1985, McGuire et al. 1986, McGuire and Gambino 1987a,b). The linear
relation between 0v and the Hall angle OH= arctan(pn/p) is presented in fig. 105 for
amorphous R - M alloys with M = A1, Cu or Au (McGuire and Gambino 1987a).

8. Technological applications

8.1. Metallic glasses

The class of amorphous M - T alloys with M = B, P, C, A1 or Si and T = Fe, Co or


Ni have received much attention with respect to their favorable properties concerning
various commercial applications such as power supplies, transformers, magnetic
sensors and transducers, magnetic heads, magnetic shielding or magnetometers
(Luborsky et al. 1978, Boll et al. 1983, Mohri 1984, Moorjani and Coey 1984,
Hilzinger 1985, Fish and Smith 1986, O'Handley 1987a, Hilzinger 1990). Generally,
these alloys are characterized by a high electrical resistivity, high mechanical strength,
424 P. HANSEN

60
Gdl_xCOx
40 °o
oo
20
o
:& o
0

-20

-40

-60 Xcomp
/
-80 I I I I~ I i t
0.60 0,70 0.80 0.90 1.00
X
Fig. 104. Compositional dependence of the spontaneous Hall coefficient at T = 295K for amorphous
G d - C o alloys (McGuire et al. 1980a).

3 R°'6°M°'40 R=N..~d
/

er......_/E~
2
E
Dy~/8°
Z I

0
Tb...~

-2 -1 0 1
OH(deg}
Fig. 105. Faraday rotation at 2 = 633nm and T = 1.6K as a function of the Hall angle for amorphous
R - M alloys with M = A1, Cu or Au (McGuire and Gambino 1987a).

good corrosion resistance, the absence of crystalline anisotropy, structural defects


and grain boundaries due to the noncrystalline state. The magnetic properties such
as saturation flux density, Curie temperature, magnetostriction and induced anisot-
ropy can be controlled by the alloy composition and a subsequent temperature
MAGNETIC AMORPHOUSALLOYS 425

treatment. The Fe-rich alloys exhibit the highest saturation flux density and the Co-
based alloys are characterized by low magnetostriction, very high permeabilities and
low magnetic losses (Yagi et al. 1988).
The high electrical resistivity and the small thickness of the melt-quenched ribbons
lead to low eddy current losses. In combination with the low hysteresis losses, this
results in very low core losses which is of interest for power electronics at high
frequencies (Pfeifer and Kunz 1982, Boll and Hilzinger 1983, Lupi 1988). Magnetic
core materials for transformers operating at frequencies above 20 kHz require low
power losses and a high flux density to achieve a high output power (Grfitzer 1978).
Disadvantages arise from the lower attainable induction of metallic glasses (Luborsky
and Johnson 1981) as compared to the grain-oriented Si-Fe and from the lower
packing factor of the core material.
Metallic glasses were also utilized in particle accelerators (Birx et al. 1983, Raskine
and Smith 1983) and for pulse compression systems to obtain high-power pulses
(Chu et al. 1982, Pacala et al. 1984, Smith 1988).
Materials for magnetic heads need a high saturation flux density, a high magnetic
permeability, thermal stability and a good resistance to wear and corrosion. Amor-
phous Co-based alloys are attractive candidates for this application (Takahashi et
al. 1983) and are used to manufacture audio heads for tape recorders.
Another field of application for amorphous alloys are sensors and transducers
(Mohri 1984) utilizing primarily the low coercivity of nonmagnetostrictive alloys for
magnetometer applications or the stress dependence of the hysteresis loop in mag-
netoelastic sensors for displacement, stress or torque measurements. The small eddy
current losses of the ribbons permit to operate the magnetometers at frequencies
above 100kHz, leading to designs with small coils, low power consumption and a
quick response (Mohri 1983). The combination of favorable mechanical and magnetic
properties of metallic glasses has led to the development of magnetoelastic transduc-
ers. The stress-induced anisotropy K~ = -~-o-2s gives rise to a strong influence on
the permeability and the hysteresis. This can be used to design force and displacement
transducers using a single amorphous core multivibrator with a DC output (Mohri
and Sudoh 1981). Other transducer designs provide a high linearity (Meydan and
Overshott 1982), magnetic torque transducers offer the possibility of a rapid and
contactless response and low magnetostriction amorphous alloys permit the develop-
ment of high-frequency transducers.
Metallic glasses are also suitable soft-magnetic materials for magnetic shielding
due to their high permeability which applies primarily for Co-based alloys like
(MoSiB)o.aoFe0.04Coo.66 (Boll and Borek 1980, Warlimont and Boll 1982). Such
alloys are insensitive to strain and shock and reveal better shielding factors at low
fields than crystalline materials.

8.2. Magneto-optical recording

8.2.1. Storage principle


The structural disorder in magnetic alloys has generated a number of interesting
features which can be tailored for different applications. The amorphous rare-earth-
426 P. HANSEN

transition-metal films deposited by magnetron sputtering represent a suitable class


of materials for magneto-optical information storage and, therefore, these materials
have received much attention in the past decade. The magneto-optical recording
represents one of the most advanced storage techniques (Hartmann et al. 1984a,
Imamura et al. 1985, Kryder 1985, Meiklejohn 1986, Connell 1986, Hansen and
Heitmann 1989, Hansen 1990, Klahn et al. 1990b).
It combines the merits of magnetic and optical techniques. The optical disk systems
offer the unlimited cyclability of the magnetic media, contactless write, erase and
read operations, high storage capacity and removability of the optical disk. This has
led to strong efforts concerning the tailoring of the amorphous R-T alloys and with
respect to the development of optical recorder with polarization sensitive detection.
At present, amorphous R-T alloys of general composition GdTb-Fe, Tb-FeCo and
Dy-FeCo are the most promising candidates for magneto-optical storage. They fulfill
all requirements concerning the magnetic and magneto-optical properties and with
respect to those imposed by the recording system.
The thermomagnetic switching process is based on a few simple principles. The
write/erase process utilizes the temperature characteristics of the magnetic properties
where the temperature profile of the coercivity plays a dominant role. A typical
temperature variation of the coercive field Hc is sketched in fig. 106 (Hansen 1987,
Hansen and Witter 1988, Craseman et al. 1989). In the room-temperature range, the
high Ho fixes any domain configuration, while at temperatures above T~ the material
becomes magnetically soft permitting to orient the direction of magnetization parallel
to that of an applied magnetic field Hs. This different behavior at room temperature
and at high temperatures is reflected by the corresponding magnetic hysteresis loops

T=295K

300

30 . . . . . . . . . J....... _N~-ILls ÷H s
l \
29s t t5~o
&To
T(K)
Fig. 106. Schematic representation of the temperature variation of the coercive field for an amorphous
rare-earth-transition-metal alloy. The material changes from a magnetically 'hard' state around room
temperature to a magnetically 'soft' state at high temperatures where the direction of magnetization can
be reversed by weak switching field H s as indicated by the hysteresis loops.
M A G N E T I C A M O R P H O U S ALLOYS 427

also shown in fig. 106. The temperature rise can be achieved with a laser beam
heating the film locally as sketched in fig. 107. The maximum temperature reached
in the magneto-optical film is determined by the thermal constants of the total film
stack and the laser pulse energy. The indicated asymmetry of the temperature profile
results from the disk velocity. The writing process is performed either by laser
modulation at constant magnetic field or by field modulation at constant laser power.
The latter permits direct overwrite but is limited in the switching frequency to the
range <10 MHz due to the operation margins of the coil generating the magnetic
field. The information thus obtained is stored by magnetic domains. Their size is
determined by the temperature profile and is typically of the order of 1 ~tm. The
written domains can be imaged by Lorentz microscopy (Suits et al. 1986, Rugar et
al. 1987, Suits et al. 1987, 1988, Ichihara et al. 1988a,b, Greidanus et al. 1989a,b,c,
Zeper et al. 1989a,b), scanning electron microscopy (Aeschlimann et al. 1990) or by
magnetic force microscopy (den Boef 1990, Rugar et al. 1990) which is shown in
fig. 108 for domains written by laser modulation and field modulation. The domains
in fig. 108c (left) indicate the accurrance of subdomains.
Reading of the stored information is performed utilizing the magneto-optical Kerr
effect that represents the rotation of the plane of polarization of linearly polarized
light as a function of the direction of magnetization when reflected at the surface of
a magnetic material. Thus, the written information can be read using polarization
optics detecting the difference in polarization of the reflected light for the two possible
directions of magnetization in a uniaxial material with respect to the film normal.
Some further aspects are discussed in section 8.2.5 in connection with the optical
recorder.

objective ~ ~ t
,e o+

,/n
li ( .O-,oye
d°moir~-~+l I II ~reftective
watt H,~--.~ I- -I - - I - -I---,~.---.~ foyer
I'~++/I ~ I I l~,~l
.F~'+++..J J i i t l . / " ' - J
coit/r-~--l-]
~-1~m-~:
',/3',
position
Fig. 107. Principle of thermomagnetic writing or erasure. The magneto-optical layer is locally irradiated
by a laser beam reducing the coercivity in the heated volume according to fig. 106, and the magnetization
can be switched by an external magnetic field.
428 P. HANSEN

(Ct) ' ' 5,u,m (b) ' " 5/~m

(c} ' J 1.5 ,u,m

(d) .... 5,u,m


Fig. 108. (a), (b) Magnetic domains in amorphous GdTb-Fe films imaged by Lorentz microscopy (Greida-
nus et al. 1989a,b). The domains in (a) were written by laser modulation at different magnetic fields. The
laser power was 6.8mW and the linear disk velocity 2.1 m/s. The pulse frequency was 1 MHz and the
pulse time 400 ns. The domains in (b) were written by field modulation at different frequencies: The laser
power was 3.7 roW, the linear disk velocity 0.5 m/s and the applied field 24.7 kA/m. (c) The domains in
the Tb-FeCo films were imaged by magnetic force microscopy (den Boer 1990). The domains were written
by laser modulation. (d) Domains in a Co/Pt multilayer imaged by Lorentz microscopy (Greidanus et al.
1989c, Zeper et al. 1989a,b).

8.2.2. Material selection


The write a n d r e a d processes are b a s e d on a n u m b e r of magnetic, m a g n e t o - o p t i c a l
a n d o p t i c a l p r o p e r t i e s which have to be well c o n t r o l l e d to meet the r e q u i r e m e n t s
i m p o s e d o n the m a g n e t o - o p t i c a l material:
- p r e s e n c e of a uniaxial m a g n e t i c a n i s o t r o p y with a positive a n i s o t r o p y constant,
Ku, with Ku > ~1 t o M s ,2 .
- large r o o m - t e m p e r a t u r e coercivity to p r o v i d e a high d o m a i n stability a n d s t o r a g e
density;
- a Curie t e m p e r a t u r e r a n g i n g between 400 K <~ T < 550 K;
MAGNETIC AMORPHOUSALLOYS 429

-squareness of the hysteresis loop in the high-temperature range to guarantee


perfect switching characteristics;
-high optical absorption, ~, to produce a sufficiently high local temperature rise;
- high figure of merit RO 2 (R = reflectivity);

- long-term stability.

There are many magnetic materials fulfilling some of these requirements but at
present there are only the amorphous R-T alloys and to some extent also a few
magnetic oxides (Abe and Gomi 1987) and Pt-Co multilayers (Zeper et al. 1989a,b,
Greidanus et al. 1989c, Hashimoto and Ochiai 1990) satisfying all these requirements
together with some additional conditions inferred from the recording system.
As discussed in sections 5 and 6, the magnetic and magneto-optical properties can
be well controlled by the composition and the deposition parameters. A high uniaxial
anisotropy can be obtained in Tb-containing alloys (fig. 66b). The presence of Tb
induces also a sufficiently high coercivity (fig. 78b) and the films are characterized by
a good squareness of the hysteresis loops. The Curie temperature for Co-rich alloys
is too high (fig. 57) while that for Tb-Fe alloys is too low. However, in the latter
case, small additions of Co can be used to control Tc where Tc is increased by
roughly 7 to 10 K per at.% Co depending on the R-Fe alloy. The presence of a
compensation temperature used to optimize the temperature profile of Hc can be
adjusted for R1-xTx alloys by the ratio (1 -x)/x. The optical absorption in metals
is very high and, thus, the local heating using a laser leads to no principal problems.
The optical Kerr rotation is not large but the figure of merit R02 is sufficient to
reach good read-out characteristics. An optimal set of magnetic and magneto-optical
parameters thus can be achieved for ternary GdTb-Fe, Tb-FeCo or Dy-FeCo
alloys. The high corrosivity of these alloys requires a protection against air and can
be achieved by suitable coatings as outlined in section 8.2.4.

8.2.3. Thermomagnetic switching process


The storage of information in amorphous alloys is based on thermomagnetic writing
and erasure. They are primarily controlled by the laser-induced temperature profile,
the radial and time dependence of the magnetic properties and the domain nucleation
and domain-wall motion (Heitmann et al. 1985, Kryder 1985, Imamura et al. 1985,
Connell 1986, Hansen and Heitmann 1989, Hansen 1990). The understanding of the
thermomagnetic switching process involves the treatment of three basic problems: (i)
the solution of the equations of heat conduction for a multilayer structure leading
to the temperature distribution in the magneto-optic films; (ii) the calculation of the
temperature and radial dependence of the relevant magnetic parameters for ternary
or quarternary alloys; and (iii) the investigation of the conditions for domain-wall
stability (Huth 1974, Mansuripur and Connell 1983, 1984, Mansuripur 1987, Hansen
1987, 1988b, McDaniel and Mansuripur 1987, Nagato et al. 1988, Takahashi et al.
1988, Sato et al. 1988a,b, Suits et al. 1988).
The basic requirement for thermomagnetic switching is the presence of a uniaxial
anisotropy with Ku >~poMs
1 2 and of sufficiently high coercivity to guarantee high
stability of any magnetization configuration at room temperature. Only then two
430 P. HANSEN

stable directions of M, parallel and antiparallel to the film normal, occur which is
necessary to represent the two logical states '0' and '1'.
The formation of a magnetic domain in the presence of an applied field then is
possible at temperatures close to Tc where Hc is low (figs. 80 and 106). The size and
shape of the domain is controlled by the driving force F = -~E/~R, where E is the
total energy and R the position of the domain wall. F can be expressed in the form

F -- IF~ + FD + Fwl - Fno, (105)


where Fn, FD, Fw, Fn~ are the forces associated with the field energy, demagnetizing
energy, domain-wall energy and the coercive energy. The forces are functions of the
radial profile and thus depend on the temperature. They can be expressed for the
case of cylindrical symmetry by (Huth 1974, Hansen 1987, 1988b)

Fn = -T- 4rcohRHoMs(R), (106a)


Fno = 4ZC#ohRHc (R)Ms(R), (106b)
Fw = 2nh[aw(R) + R~aw(R)/~R], (106c)
FD = - - 47r/2ohRMs (R)/Td(R). (106d)
Hd(R) is the z-averaged demagnetizing field and in general requires numerical
computation (Huth 1974, Suits et al. 1988). An analytical expression for/Td(R ) can
be obtained using a rough approximation where the radial dependence of Ms(R) is
replaced by its average value M(R) (Hansen 1987, 1988b). The Ha(R) can be written
in the form /qd(R)= ffls(R)f(R, rwoomp, h) where f(R, rr . . . . , h) can be expressed in
terms of elliptic integrals. Ho, h, R, r and rTcom p a r e the applied magnetic field, the
film thickness, the domain radius, the radial coordinate and the compensation
temperature radius [Ms(rrcomp)=O], respectively, r r . . . . appears explicitly in
f(R, rrcomp, h) only for films with T~omp above ambient temperature, T~. The - and
+ sign in eq. (106a) refer to writing and erasing, respectively. Domain expansion
occurs for Fn + FD + Fw < 0 and domain contraction for Fn + FD + Fw > 0, provided
F > 0. The laser-induced temperature profile T(r) causes a strong radial dependence
of M s, K,, He and aw and thus also with respect to the force balance. The general
problem of the calculation of the temperature distribution in a stack of layers can
be solved numerically (Mansuripur et al. 1982, Bartholomeusz 1989, Holtslag 1989).
However, for R-T-coated films also an approximation can be used treating layers
with similar thermal behavior as one effective layer. This model yields the relation
(Holstslag 1989)
r(r) = T~ + A(t) exp(-- r2/r2), (107)
where t is the pulse duration, rt = ~ o + x t, ro the radius of the laser beam and
the diffusivity of the effective layer. From eqs. (106) and (107), the radial dependence
of the forces acting on the domain wall can be calculated where the radial dependence
of aw can be obtained combining eqs. (105), (106) and (107) and using the sublattice
magnetization inferred from mean field results. A typical variation of the forces as a
function of the wall position is shown in fig. 109 for a GdTb-Fe alloy with T~omp
MAGNETIC AMORPHOUS ALLOYS 431

1.0
Sf '-
O. 5 /t fJ
// R{,~m) /
E J'0.2 O.Z, 0.6~ 0.8
o
o

R=r./cr ~% ~ ~.
~ f eff
. . . .
J
o,5 \ \ /
"• ~"' fH
1.0 ~fHe
!

Fig. 109. Radial dependence of the reduced forces f = F/4~h 2 acting on the domain wall, where h is the
film thickness. The written domain radius R is determined by the coercive radius rno obtained from the
condition feff (r//°) = ff/c (rH¢) where feff =ftt + fo + fw.

above room temperature. The laser beam diameter is 1 ~tm and the maximum temper-
ature in the spot center is slightly higher than the Curie temperature. Thus, the
material is paramagnetic for radii smaller than the radis rrc where the Curie temper-
ature is reached. The forces represented in fig. 109 are normalized according to f - -
F/4~h 2. During writing, the stable domain radius is reached at the coercive radius
R = r~c defined by F(rnc) = 0. At this radius, the effective force Feff = Fn + FD + Fw
driving the domain wall is balanced by the coercive force F~o. For a rotating disk
the domain length can be approximated by
L = 2rno + vt, (108)
where v is the linear disk velocity of typically 5 m s 1 and t is the pulse duration of
30 to 50ns. This leads to a domain length of 1 gm for the case presented in fig. 109.
When the laser is switched off, the heated spot cools down very rapidly and the
domain is frozen due to the increasing coercivity. The magnetization in the unheated
region remains undisturbed. The erase process can be considered as the reversed
writing process with the opposite direction of the applied magnetic field. The optimum
write/erase conditions are mainly determined by the magnetic field, the compensation
temperature and the Curie temperature where the latter has to be adapted to the
available laser power.
Strong deviations from this ideal switching behavior can result from different
sources such as local material inhomogeneities, effects on submicron scale, fluctua-
tions of the focussing and tracking system, laser-induced thermal fluctuations, etc.
In particular, the occurrence of subdomains as shown in fig. 108c or irregular shaped
domains as reported for Tb-Fe and Tb-FeCo films (Suits et al. 1987, 1988, Greidanus
et al. 1989b) require a different model of domain formation in terms of the nucleation
field present (Suits et al. 1987, 1988).
432 P. HANSEN

8.2.4. Magneto-optical disk


The amorphous R - T films are preferentially prepared by magnetron sputtering,
yielding dense films at high sputtering rates without substrate cooling (Klahn et al.
1990a). However, the bare films are not sufficiently stable in air and thus need a
protection which can be achieved by a suitable coating. A typical film stack is
sketched in fig. 110. It consists of a pregrooved PC or glass substrate, an antireflective
dielectric layer (80 nm), the magneto-optical layer (45 nm), a reflective metal layer
(30 nm) and finally a protective polymer layer (some I~m). The pregroove structure
serves for tracking. From accelerated life tests with disks of this type of multilayer
structure, life times of more than 10 years were extrapolated (Klahn et al. 1987).
However, the multilayer structure serves also for an enhancement of the carrier-to-
noise ratio which requires a proper adjustment of the thicknesses of the intermediate
antireflective and the magneto-optical layer. Finally, the switching sensitivity depends
on the thermal constants of the different films and thus the stack additionally has
to be optimized in this respect. Therefore, deposition processes for the different layers
have to be well controlled to reach optimum switching characteristics and to keep
the raw byte error rate below 10 .4 to 10 -5. The error correction codes and inter-
leaving reduce the byte error rate to a level of 10 -16 to 10 -17.
Direct overwrite capability is a further requirement for many storage applications.
Magnetic field modulation is one possibility (Tanaka et al. 1987, Nakao et al. 1987,
Ando et al. 1988, Miyamoto et al. 1989) to solve this problem. However, the switching
power and a suitable spacing between disk and coil limits the modulation frequencies
below 10 MHz. Therefore, direct overwrite methods operating without a modulated
magnetic field are of interest. Different techniques have been reported (Osato et al.
1987, Schultz and Kryder 1989, Schultz and Kryder 1990), but further development
seems to be necessary to fulfill the present recording requirements.

8.2.5. Recorder requirements


A magneto-optic recorder is very similar to other optical disk systems as, e.g., a
digital audio compact disc unit (Carasso et al. 1982, Deguchi et al. 1984). However,

laser beam

magr
Fig. 110. Cross section of a magneto-optical disk with a pregrooved substrate.
MAGNETIC AMORPHOUS ALLOYS 433

in the read-out light path, polarizing optical elements have to be added. Furthermore,
in a magneto-optic recorder differential detection of the polarization direction has
to be applied (Hartmann et al. 1985a), as sketched in fig. 111. The light emitted from
a laser, L, is collimated to a parallel beam which passes two neutral beam splitters,
NBS, and is focussed onto the disk, Di, by an objective lens of numerical aperture,
NA = 0,5. The light reflected from the disk is deflected by the beam splitters to the
detection arrangement and to the tracking and focussing circuitry, Tr, Fo. The
detection arrangement consists of a polarizing beam splitter, PBS, which splits the
light into two perpendicularly polarized components that are detected by PIN
photodiodes, D1 and D2. Their output is connected to a differential amplifier. The
2/2 waveplate is used to adjust the bias level of the differential detection circuit to a
minimum by setting the azimuth of the polarization plane to 45 ° with respect to the
characteristic beam splitter axes (see insert). The modulation of the azimuth due to
the Kerr rotation _+OK produces amplitude variations between A~ and Ai- and
A + and A2 at the diodes D1 and D2, respectively. The coil, C, provides magnetic
fields for thermomagnetic switching.
To compete with other magnetic recording systems, the magneto-optical system
has to fulfill the following requirements:
-high storage density;
-low access time;
-high write and erase sensitivity;
- l o w switching field;
-high carrier-to-noise ratio;
- l o w bit error rate;
-direct overwrite capability.
The tailoring of these recording parameters, improvement of the bit error character-
istics (Yamamoto and Yamada 1988) and the understanding of their relation to the
magnetic and magneto-optical properties are key aspects in the sussessful develop-

/x;:i [
T,Fo II/I D

Di

Fig. 111. Differential detection arrangement for magneto-optical recording (Hartmann et al. 1984a).
(L) laser, (Di) magneto-optical disk, (D1) and (D2) PIN photodiodes, (C) coil, (NBS) beam splitter,
(PBS) polarizing beam splitter, (Tr), (Fo) tracking and focussing. The modulation of the azimuth due to
the Kerr rotation produces amplitude variations between A 1,2 + and A~2 at the diodes Dx and D2,
respectively.
434 P. HANSEN

ment of the magneto-optical system. Recorder measurements performed for various


disks based on GdTb-Fe and Tb-FeCo films reveal a characteristic dependence on
the magnetic parameters (Crasemann et al. 1989, Crasemann and Hansen 1989)
which can be interpreted in terms of the switching model outlined in one of the
previous sections. An example is displayed in fig. 112 showing the decrease of the
domain length, L, with increasing Tc at constant laser power. The full line was
calculated from eqs. (105) and (106) using an average radial coercivity profile. The
scatter in the data primarily has to be associated with the different T~ompvalues of
the films. Some typical data for a magneto-optical disk are compiled in table 20.
With these material data, the magneto-optical technology offers a wide field for
applications such as the data storage, digital audio recording or video recording.
Also, the magneto-optic system is expected to be a good candidate as a multipurpose
recording system due to its large storage density and relatively short access time.

1'65[ o Tb-FeCo
L • • GdTb-Fe
1,4~-~ • • ~theory

oOo" "
.o
o ooO

i:2t , , ,\
400 420 440 /.60 480 500 520
Tc[K)

Fig. 112. Domain length as a function of Curie temperature at a writing field of 16kA/m and writing
energy of 0.6nJ/pulse (Crasemann et al. 1989, Crasemann and Hansen 1989).

TABLE 20
Typical performance data of a magneto-optical disk (Hansen 1990).
Alloy Tb-FeCo, GdTb-Fe, Dy-FeCo
Write power (50 ns pulse) <10mW
Erase power (CW) <10mW
Read power ~<1mW
Write/erase field 16-32 kA/m
Archival life >10 years
CNR >50dB (1 MHz, 30kHz bandwidth)
Bit size (recording density) 1.0-1.5 pm ( ~ 25000 bpi)
Track width (track density) 1.6 pm ( ~ 17000 tpi)
Speed 220 rpm
Capacity (5¼ inch, per side) 500 MByte (3 x 10v bit/cm2)
Access time (average latency) 44ms (14ms)
Transfer rate 1.1 MByte
MAGNETIC AMORPHOUS ALLOYS 435

9. Summary

The structural disorder of magnetic M-T, M - R and R - T alloys (T: magnetic


transition metals, R: rare earths, M: not T and R) gives rise to significant changes
of the mechanical, electrical, magnetic and magneto-optical properties as compared
to the crystalline counterparts. Although it is not yet possible to explain many
properties in terms of the relevant band structure of amorphous alloys, many new
concepts and theories were generated and an overwhelming amount of experimental
work was performed, leading to new physical insights concerning the atomic structure
and short-range order, the relation between chemical bonding and magnetism, mag-
netic structures, anisotropic magnetic properties and transport properties on the one
hand and the development of materials with a unique combination of properties on
the other hand which makes these alloys attractive for a variety of applications.
Generally, amorphous Fe-based alloys behave differently from Co- and Ni-based
alloys due to their stronger (sp)-d hybridization and weaker covalent p - d bonding,
resulting in broader valence bands. Also, they exhibit a sensitive dependence of the
exchange coupling on the atomic distance, leading even to antiferromagnetic bonds
for Fe-Fe distances below 0.25 nm. The distribution of atomic distances inferred
from the radial distribution function suggests the presence of a concentration-depen-
dent portion of negative exchange interactions. Thus, various amorphous M1-xFex
alloys show noncollinear magnetic structures, and for M = Zr, Hf, Y, La, Ce or Lu
even speromagnetic or spin-glass-like behavior occurs with a tricritical point for
x ~>0.9 and Curie temperatures ranging between 100 and 200K.
In amorphous M - C o alloys, small Co-Co distances are favorable for the magnetic
moment formation and tend to increase the exchange interaction and thus Tc.
Therefore, amorphous Co-based alloys are predominantly strong ferromagnets, al-
though significant differences in the magnetic properties are observed for Co-based
alloys containing metalloids or other elements. The magnetic moment variation was
interpreted in terms of the magnetic valence model or the environment model. In
both cases, the experimental data were well described for certain classes of alloys.
Most transition-metal-metalloid alloys exhibit good corrosion resistance, high
electrical resistivity, good mechanical properties and are soft-magnetic materials. Fe-
rich alloys exhibit the highest saturation flux density, and the Co-based alloys show
low magnetostriction, high permeabilities and very low magnetic losses. These proper-
ties make various magnetic glasses attractive candidates for commercial applications
such as power supplies, transformers, sensors, transducers, magnetic heads, magnetic
shielding or magnetometers.
Amorphous rare-earth-transition-metal alloys reveal pronounced differences in
their magnetic properties as compared to M - T alloys due to the different electronic
structure of the rare earths and the presence of two magnetic sublattices formed from
elements of different groups. The negative exchange coupling between the 5d rare-
earth electrons and the 3d transition-metal electrons leads to a parallel alignment of
the R and T moments for the light rare earths and an antiparallel alignment for the
heavy rare earths. A further difference in the amorphous R - T alloys with respect to
M - T alloys is the strong influence of the structural disorder on the local direction
436 P. HANSEN

of the R moments which are coupled via the strong spin-orbit coupling to the
randomly varying axes of the electrostatic field. This leads to sperimagnetic structures
except for the Gd (S-state) based alloys exhibiting ferrimagnetic order. Amorphous
R-Fe and R-Co alloys reveal the same differences as observed for M - F e and M -
Co alloys. Fe-rich R-Fe alloys exhibit a very low Tc, below 200 K, due to competing
positive and negative exchange interactions. Therefore, all Rl_~Fex alloys show a
maximum in the concentration dependence of Tc around x ~ 0.7, followed by a
strong turndown of Tc at large x, in contrast to R-Co alloys revealing a steep
increase of Tc for alloy compositions with x above the critical composition.
Amorphous R-T alloys containing non-S-state rare earths are characterized by
strong uniaxial anisotropies and high coercivities but low magnetizations in the case
of sperimagnetic order. The presence of a compensation temperature for the antiferro-
magnetically coupled alloys gives rise to strong influences on the temperature and
the concentration dependence of the magnetic, magneto-optical and transport proper-
ties. Amorphous R-T alloys also show a large extraordinary Hall effect arising from
side jumb and skew scattering processes.
The favorable magnetic and magneto-optical properties have led to the develop-
ment of GdTb-Fe, Tb-FeCo or Dy-FeCo alloys for magneto-optical data storage.

References

Abe, M., and M. Gomi, 1987, J. Magn. Soc. Jpn. Amamou, A., 1980, Solid State Commun. 33,
11, Suppl., p. 299. 1029.
Aboaf, J.A., and E. Klokholm, 1981, J. Appl. Amamou, A., and G. Krill, 1979, Solid State
Phys. 52, 1844. Commun. 31, 971.
Aeschlimann, M., G.L. Bona, F. Meier, M. Stam- Amamou, A., and G. Krill, 1980, Solid State
poni, A. Vaterlaus, H.C. Siegmann, E.E. Commun. 33, 1087.
Marinero and H. Notarys, 1988, IEEE Trans. Amaral, V.S., J.M. Moreira, J.B. Sousa, B. Bar-
Magn. MAG-24, 3180. bara, J. Filippi and B. Dieny, 1988, J. de Phys.
Aeschlimann, M., M. Scheinfein, J. Unguris, (France) 12, Suppl. C8-1237.
F.J.A.M. Greidanus and S. Klahn, 1990, J. Anderson, P.M., and A.E. Lord, 1980, Mater. Sci.
Appl. Phys. 68, 4710. & Eng. 44, 279.
Aharony, A., 1962, Rev. Mod. Phys. 34, 227. Anderson, P.W., B.I. Halperin and C.M. Varma,
Alameda, J.M., Y. Berthier, F. Briones, M.C. 1972, Philos. Mag. 25, 1.
Contreras, D. Givord, A. Li6nard and H. Ando, R., K. Fujiie, T. Yoshida, K. Watanabe
Rubio, 1985, J. Magn. & Magn. Mater. 54-57, and T. Nagaki, 1988, IEEE Transl. J. Magn.
233. Jpn. 3, 633.
Aldred, A.T., 1975, Phys. Rev. B 11, 2597. Arai, K.I., N. Tsuya, M. Yamada, H. Shirae, H.
Aldred, A.T., and P.H. Froehle, 1972, Int. J. Fujimori, H. Saito and T. Masumoto, 1976,
Magn. 2, 195. in: Proc. 2nd Int. Conf. Rapidly Quenched
Alfonso, C.N., A.R. Lagunas, F. Briones and S. Metals, eds N.J. Grant and B.C. Giessen (MIT
Gir6n, 1980, J. Magn. & Magn. Mater. 15-18, Press, Cambridge, MA) p. 489.
833. Arajs, S., B.L. Tehan, E.E. Anderson and A.A.
Algra, H.A., K.H.J. Buschow and R.A. Henskens, Stelmach, 1970, Int. J. Magn. 1, 41.
1980, J. Physique 41, Suppl. C8-646. Aratani, K., T. Kobayashi, S. Tsunashima and S.
Allen, R., and G.A.N. Connell, 1982, J. Appl. Uchiyama, 1985, J. Appl. Phys. 57, 3903.
Phys. 53, 2353. Arrese-Boggiano, R., J. Chappert, J.M.D. Coey,
Alperin, H.A., J.R. Cullen and A.E. Clark, 1976, A. Li6nard and J.P. Rebouillat, 1976, J. Phy-
AIP Conf. Proc. 29, 186. sique 37, Suppl. C6-771.
MAGNETIC AMORPHOUS ALLOYS 437

Asomoza, R., I.A. Campbell, A. Fert and R. nathan and T.R. Anantharaman, 1982, Solid
Meyer, 1977a, J. Phys. F 7, L327. State Commun. 44, 905.
Asomoza, R., I.A. Campbell, H. Jouve and R. Bhatnagar, A.K, B.B. Prasad and R. Jaganna-
Meyer, 1977b, J. Appl. Phys. 48, 3829. than, 1984, Phys. Rev. B 29, 4896.
Asomoza, R., I.A. Campbell and A. Fert, 1979a, Bhattacharjee, A.K., R. Julien and M.J. Zucker-
J. Physique 40, Suppl. C5-225. mann, 1977a, J. Phys. F 7, 393.
Asomoza, R., I.A. Campbell, A. Fert, A. Li~nard Bhattacharjee, A.K., B. Coqblin, R. Julien and
and J.P. Rebouillat, 1979b, J. Phys. F 9, 349. M.J. Zuckermann, 1977b, Physica B 91, 179.
Aur, S., T. Egami, A.E. Berkowitz and J.L. Waiter, Biesterbos, J.W.M., A.G. Dirks, M.A.J.P. Furla
1982, Phys. Rev. B 26, 6355. and J.P. Grundy, 1979, Thin Solid Films 58,
Avrami, M., 1939, J. Chem. Phys. 7, 1103. 259.
Avrami, M., 1940, J. Chem. Phys. 8, 212. Birgeneau, R.J,, J.A. Tarwin, G. Shirane, E.M.
Avrami, M., 1941, J. Chem. Phys. 9, 177. Gyorgy, R.C. Sherwood, H.S. Chen and C.L
Babir, E.E., R. Krsnik and B. Leontic, 1970, J. Chien, 1978, Phys. Rev. B 18, 2192.
Phys. E 3, 1014. Birx, D.L., E. Cook, S. Hawkins, L.L Reginato,
Babir, E.E., Z. Marohnic and J. Ivkov, 1978, J. Schmidt and M. Smith, 1983, IEEE Trans.
Solid State Commun. 27, 441. Nucl. Sci. NS-30, 2763.
Barker, J.A., J.L. Finney and M.R. Hoare, 1975, Black, J.L., and B.L. Gyorffy, 1978, J. Physique
Nature 257, 120. 37, Suppl. C6-941.
Bartholomeusz, B.J., 1989, J. Appl. Phys. 65, 262. Blasberg, E., D. Korn and H. Pfeifle, 1979, J.
Becket, J.J., F.E. Luborsky and J.L. Walter, 1977, Phys. F 9, 1821.
IEEE Trans. Magn. MAG-13, 988. Bloch, D., and R. Lemaire, 1970, Phys. Rev. B 2,
Bedell, J.R., 1975, U.S. Patent No. 3, 862, 658 2648.
(Jan. 28, 1975). Bloch, D., D.M. Edwards, M. Shimizu and J.
Beebey, J.L., 1964, Phys. Rev. A 135, 130. Voiron, 1975, J. Phys. F 5, 1217.
Beloritzky, E., M.A. Frrmy, J.P. Gavigan, D. Boll, R., and L. Borek, 1980, NTG-Fachberichte
Givord and H.S. Li, 1987, J. Appl. Phys. 61, 76, 178.
3971. Boll, R., and H.R. Hilzinger, 1983, IEEE Trans.
Bennett, C.H., 1972, J. Appl. Phys. 43, 2727. Magn. MAG-19, 1946.
Bennett, H.S., and E.A. Stern, 1965, Phys. Rev. Boll, R., H.R. Hilzinger and H. Warlimont, 1983,
A 137, 448. in: Metallic Glasses, Chemical and Structural
Berger, L., 1970, Phys. Rev. B 2, 4559. Properties, ed. R. Hasegawa (CRC Press, Boca
Berger, L., 1972, Phys. Rev. B 5, 1862. Raton) p. 183.
Berger, L., 1973, Phys. Rev. B 8, 2351. Boswell, F.G., 1980, J. Thermal Anal. 18, 353.
Berger, L., 1977, Physica B 91, 31. Boucher, B., 1976, J. Physique 37, L345.
Berger, L., and G. Bergmann, 1980, in: Hall Effect Boucher, B., 1977, IEEE Trans. Magn. MAG-13,
and its Applications, eds C.L. Chien and C.R. 1601.
Westgate (Plenum Press, New York) p. 55. Brett, M.J., 1989, J. Mater. Sci. 24, 623.
Bergmann, G., and P. Marquardt, 1978, Phys. Brooks, H.A., 1976, J. Appl. Phys. 47, 344.
Rev. B 18, 326. Brown Jr, W.F., 1945, Rev. Mod. Phys. 17, 15.
Bernal, J.D., 1960, Nature 188, 410. Brunsch, A., and J. Schneider, 1978, IEEE Trans.
Berrada, A., M.F. Lapierre, B. Loegel, P. Panis- Magn. MAG-14, 731.
sod, C. Robert and J. Beille, 1977, Physica B Buschow, K.H.J., 1977, Rep. Prog. Phys. 40, 1179.
86-88, 790. Buschow, K.H.J., 1980a, J. Appl. Phys. 51, 2795.
Berrada, A., M.F. Lapierre, B. Loegel, P. Panissod Buschow, K.H.J., 1980b, J. Magn. & Magn.
and C. Robert, 1978, J. Phys. F 8, 845. Mater. 21, 97.
Berrada, A., J. Durand, N. Hassanain and B. Buschow, K.H.J., 1980c, in: Handbook on Ferro-
Loegel, 1979, in: Rare Earths in Modern Sci- magnetic Materials, Vol. 1, ed. E.P. Wohlfarth
ence and Technology, Vol. 2 (New York). (North-Holland, Amsterdam) p. 297.
Berry, R.S., and W.C. Pritchet, 1978, Solid State Buschow, K.H.J., 1981a, J. Less-Common Met.
Commun. 26, 827. 79, 9.
Bethe, H.A., 1935, Proc. R. Soc. London A 150, Buschow, K.H.J., 1981b, J. Appl. Phys. 51, 3319.
552. Buschow, K.H.J., 1982a, Solid State Commun.
Bhatnagar, A.K., B.B. Prasad, N. Ravi, R. Jagan- 43, 171.
438 P. HANSEN

Buschow, K.H.J., 1982b, J. Appl. Phys. 53, 7713. Capellmann, H., 1974, J. Phys. F 4, 1966.
Buschow, K.H.J., 1983, J. Phys. F 13, 563. Carasso, M.G., J.B.H. Peek and J.P. Sinjou, 1982,
Buschow, K.H.J., 1984a, in: Handbook on the Philips Tech. Rev. 40, 151.
Physics and Chemistry of Rare Earth, Vol. 7, Cargill III, G.S., 1970, J. Appl. Phys. 41, 2248.
eds K.A. Gschneidner Jr and L. Eyring (Else- Cargill III, G.S., 1975a, Solid State Phys. 30, 227.
vier, Amsterdam) p. 265. Cargill III, G.S., 1975b, Solid State Phys. Vol. 30,
Buschow, K.H.J., 1984b, Proceedings of MRS eds H. Ehrenreich, F. Seitz and D. Turnbill
Conference, ed. M. yon Allmen, Strasbourg, (Academic Press, New York) p. 227.
1984 (Les l~dition des Physique, Les Ulis) Cargill III, G.S., 1981a, in: Diffraction Studies in
p. 313. Non-Crystalline Substances, eds I. Hargittai
Buschow, K.H.J., 1985, Philips Tech. Rev. 42, 48. and W.J. Orville-Thomas (Elsevier, Amster-
Buschow, K.H.J., 1988, in: Handbook on Ferro- dam) p. 731.
magnetic Materials, Vol. 4, eds E.P. Wohlfarth Cargill III, G.S., 1981b, in: Diffraction Studies in
and K.H.J. Buschow (North-Holland, Amster- Non-Crystalline Substances, eds I. Hargittai
dam) p. 493. and W.J. Orville-Thomas (Elsevier, Amster-
Buschow, K.H.J., and N.M. Beekmans, 1979a, dam) p. 781.
Phys. Rev. B 19, 3843. Cargill III, G.S., 1983, in: Amorphous Materials,
Buschow, K.H.J., and N.M. Beekmans, 1979b, J. ed. V. Vitec (Warrendale, Pennsylvania) p. 15.
Appl. Phys. 50, 6348. Cargill III, G.S., and R.W. Cochrane, 1974, J.
Buschow, K.H.J., and N.M. Beekmans, 1980, Physique 35, Suppl. C4-269.
Phys. Status Solidi A 60, 193. Cargill III, G.S., and S. Kirkpatrick, 1976, AIP
Buschow, K.H.J., and P.H. Smit, 1981, J. Magn. Conf. Proc. 31, 339.
& Magn. Mater. 23, 85. Cargill III, G.S., and T. Mizoguchi, 1978, J. Appl.
Buschow, K.H.J., and W.W. van den Hoogenhof, Phys. 49, 1753.
1979, J. Magn. & Magn. Mater. 12, 123. Chappert, J., J.M.D. Coey, A. Li6nard and J.P.
Buschow, K.H.J., and A.M. van der Kraan, 1981, Rebouillat, 1981, J. Phys. F 11, 2727.
J. Magn. & Magn. Mater. 22, 220. Chaudhari, P., and D.C. Cronemeyer, 1975, AIP
Buschow, K.H.J., and P.G. van Engen, 1980, J. Conf. Proc. 29, 113.
Physique 41, Suppl. C8-650. Chen, D., G.N. Otto and F.M. Sehmit, 1973,
Buschow, K.H.J., and P.G. van Engen, 1981a, J. IEEE Trans. Magn. MAG-9, 66.
Appl. Phys. 52, 3557. Chert, D.X., and K.V. Rao, 1986, IEEE Trans.
Buschow, K.H.J., and P.G. van Engen, 1981b, Magn. MAG-22, 451.
Solid State Commun. 39, 1. Chen, H.S., 1978a, in: Metallic Glasses, eds J.J.
Buschow, K.H.J., and P.G. van Engen, 1981c, Gilman and H.J. Leamy (Metal Park, Ohio)
Mater, Res. Bull. 16, 1177. p. 75.
Buschow, K.H.J., M. Brouha, J.W.M. Biesterbos Chert, H.S., 1978b, J. Appl. Phys. 49, 3289.
and A.G. Dirks, 1977, Physica B 91, 261. Chen, H.S., 1980a, Rep. Prog. Phys. 43, 353.
Buschow, K.H.J., A.M. van Diepen, N.M. Beek- Chert, H.S., 1980b, Rep. Prog. Phys. 43, 394.
mans and J.W.M. Biesterbos, 1978, Solid State Chen, H.S., and C.E. Miller, 1970, Rev. Sci.
Commun. 28, 181. Instrum. 41, 1237.
Buschow, K.H.J., H.A. Algra and R.A. Heskens, Cheng, S.-C.N., M.H. Kryder and M.C.A.
1980, J. Appl. Phys. 51, 561. Mathur, 1989, IEEE Trans. Magn. MAG-25,
Buschow, K.H.J., P.G. van Engen and R. Jonge- 4018.
breur, 1983, J. Magn. & Magn. Mater. 38, 1. Chi, M.C., and R. Alben, 1977, J. Appl. Phys. 48,
Cahn, R.W., 1980, Contemp. Phys. 21, 43. 2987.
Callen, E.R., and H.B. Callen, 1965, Phys. Rev. Chi, M.C., and T. Egami, 1979, J. Appl. Phys. 50,
A 139, 455. 1651.
Callen, E.R., Y.J. Liu and J.R. Cullen, 1977, Phys. Chien, C.L., and R. Hasegawa, 1977, Phys. Rev.
Rev. B 16, 263. B 16, 2115.
Campbell, I.A., 1972, J. Phys. F 2, L47. Chien, C.L., and K.M. Unruh, 1981, Phys. Rev.
Campbell, I.A., and A. Fert, 1982, in: Handbook B 24, 1556.
on Ferromagnetic Materials, Vol. 3, ed. E.P. Chien, C.L., and K.M. Unruh, 1982, Nucl.
Wohlfarth (North-Holland, Amsterdam) Instrum. & Methods 199, 193.
p. 747. Chien, C.L., D.P. Musser, F.E. Luborsky, J.J.
MAGNETIC AMORPHOUS ALLOYS 439

Becker and J.L. Walter, 1977, Solid State Connell, G.A.N., 1986, J. Magn. & Magn. Mater.
Commun. 24, 231. 54-57, 1561.
Chien, C.L., D.P. Musser, E.M. Gyorgy, R.C. Corb, B.W., 1985, Phys. Rev. B 31, 2521.
Sherwood, H.S. Chert, F.E. Luborsky and J.L. Corb, B.W., R.C. O'Handley and N.J. Grant,
Walter, 1979, Phys. Rev. B 20, 283. 1982, J. Appl. Phys. 53, 7728.
Chien, C.L., K.M. Unruh and S.H. Liou, 1983, J. Corb, B.W., R.C. O'Handley and N.J. Grant,
Appl. Phys. 53, 7756. 1983, Phys. Rev. B 27, 636.
Chipman, D.R., L.D. Jennings and B.C. Giessen, Cornelison, S.G., and D.J. Sellmyer, 1983, J. Appl.
1978, Bull. Am. Phys. Soc. 23, 467. Phys. 53, 7722.
Choe, Y.J., S. Tsunashima, T. Katayama and S. Cote, P.J., 1976, Solid State Commun. 18, 1311.
Uchiyama, 1987, J. Magn. Soc. Jpn. 11, 273. Cote, P.J., and V. Meisel, 1977, Phys. Rev. Lett.
Choe, Y.J., S. Tsunashima and S. Uchiyama, 39, 102.
1988, J. Physique 49, Suppl. C8-1953. Crasemann, J., and P. Hansen, 1989, Thin Solid
Choe, Y.J., K. Nagase, S. Tsunashima and S. Films 175, 261.
Uchiyama, 1989, IEEE Transl. J. Magn. Jpn. Crasemann, J., P. Hansen, M. Rosenkranz and
4, 251. K. Witter, 1989, J. Appl. Phys. 66, 1273.
Chu, E.Y., B. Hofmann, H. Kent and T. Bern- Croat, J.J., 1981a, J. Appl. Phys. 52, 2509.
hardt, 1982, IEEE Conferenced Record of the Croat, J.J., 1981b, J. Magn. & Magn. Mater. 24,
15th Power Modulator Symposium, Balti- 125.
more, MD (IEEE, New York) p. 32. Croat, J.J., 1982, J. Appl. Phys. 53, 6932.
Chudnovsky, E.M., 1989, J. Magn. & Magn. Cullen, J., and A. del Moral, 1990, J. Magn. &
Mater. 79, 127. Magn. Mater. 83, 157.
Clark, A.E., 1980, in: Handbook on Ferromag- Czjzek, G., J. Fink, F. Gotz, H. Schmidt, J.M.D.
netic Materials, Vol. 1, ed. E.P. Wohlfarth Coey, J.P. Rebouillat and A. Li6nard, 1981,
(North-Holland, Amsterdam) p. 531. Phys. Rev. B 23, 2513.
Clark, A.E., and E. Callen, 1968, J. Appl. Phys. Dai, D.-S., R.-Y. Fang, L.-T. Tong, Z.-X. Lui, Z.-
39, 5972. J. Zhou and Z.-H. Lin, 1985, J. Appl. Phys.
Clemens, H.K., and J. Jaumann, 1963, Z. Phys. 57, 3589.
173, 135. Dai, D.-S., R.-Y. Fang, Z.-X Lui, H. Wang, J.
Cochrane, R.W., and J.O. Str6m-Olsen, 1977, J.
Lan, X.-L. Rao and Y.-P. Ji, 1986, Acta Phys.
Phys. F 7, 1799.
Sin. 35, 1502.
Cochrane, R.W., and J.O. Str6m-Olsen, 1978, J.
Davies, H.A., 1976, Phys. Chem. Glasses 17, 159.
Appl. Phys. 49, 1677.
Davis, LA., 1976, in: Rapidly Quenched Metals,
Cochrane, R.W., R. Harris and M. Plischke, 1974,
Section 1, eds N.J. Grant and B.C. Giessen
J. Non-Cryst. Solids 15, 239.
Cochrane, R.W., R. Harris, J.O. Str6m-Olsen and (MIT Press, Cambridge, MA) p. 369.
M.J. Zuckermann, 1975, Phys. Rev. Lett. 35, de Gennes, P.G., 1962a, J. Phys. Radium 23, 510.
676. de Gennes, P.G., 1962b, J. Phys. Radium 23, 630.
Cochrane, R.W., R. Harris and M.J. Zuckermann, de Groot, R.A., and K.H.J. Buschow, 1986, J.
1978a, Phys. Rep. 48, 1. Magn. & Magn. Mater. 54-57, 1377.
Cochrane, R.W., J.O. Str6m-Olsen, G. Williams, de Groot, R.A., F.M. Mueller, P.G. van Engen
A. Li6nard and J.P. Rebouillat, 1978b, J. Appl. and K.H.J. Buschow, 1983, Phys. Rev. Lett.
Phys. 49, 1677. 50, 2025.
Coey, J.M.D., 1978, J. Appl. Phys. 49, 1646. de Groot, R.A., F.M. Mueller, P.G. van Engen
Coey, J.M.D., and D.H. Ryan, 1984, IEEE Trans. and K.H.J. Buschow, 1984, J. Appl. Phys. 55,
Magn. MAG-20, 1278. 2151.
Coey, J.M.D., D. Givord, A. Li6nard and J.P. Deguchi, T., H. Katayama, A. Takahashi, K.
RebouiUat, 1981, J. Phys. F 11, 2707. Ohta, S. Kobayashi and T. Okamoto, 1984,
Coey, J.M.D., D.H. Ryan and Yu. Boliang, 1984, Appl. Opt. 23, 3972.
J. Appl. Phys. 55, 1800. Dekker, P., 1976, IEEE Trans. Magn. MAG-12,
Collins, M.F., V.J. Minki6wicz, R. Nathans and 311.
G. Shirane, 1969, Phys. Rev. 179, 417. del Moral, A., and J.I. Arnaudas, 1989, Phys.
Colvin, R.V., and S. Arajs, 1965, J. Phys. Chem. Rev. B 39, 9453.
Solids 26, 435. den Boef, A.J., 1990, Appl. Phys. Lett. 56, 2045.
~0 P. HANSEN

Deschizeaux, M.N., and G. Develey, 1971, J. Erskine, J.L., and E.A. Stern, 1973b, Phys. Rev.
Physique 32, 319. Lett. 30, 1329.
Dey, S., U. Gorres, H.J.V. Nielsen, M. Rosenberg Esposito, E., H. Ehrenreich and C.D. Gelatt Jr,
and M. Sostarich, 1980, J. Physique 41, Suppl. 1978, Phys. Rev. B 18, 3913.
C8-678. Faber, T.E., and J.M. Ziman, 1965, Philos. Mag.
d'Heurle, F.M., and J.M.E. Harper, 1989, Thin 11~ 153.
Solid Films 171, 81. F~ihnle, M., 1980, Phys. Status Solidi B 99, 547.
Dillon Jr, J.F., R.B. van Dover, M. Hong, E.M. Ffihnle, M., and T. Egami, 1982, J. Appl. Phys.
Gyorgy and S.D. Albiston, 1987, J. Appl. 53, 2319.
Phys. 61, 1103. F~ihnle, M., and J. Furthmfiller, 1988, J. Magn.
Dirks, A.G., and H.J. Leamy, 1977, Thin Solid & Magn. Mater. 72, 6.
Films 47, 219. F~ihnle, M., and J. Furthmfiller, 1989, Phys.
Dirks, A.G., and H.J. Leamy, 1978, J. Appl. Phys. Status Solidi A 116, 819.
49, 1735. F~ihnle, M., and H. Kronmiiller, 1980, Phys.
Dirks, A.G., J.W.M. Biesterbos and K.H.J. Status Solidi B 98, 219.
Buschow, 1977, Physica B 86-88, 761. F/ihnle, M., G. Herzer, H. Kronmfiller, R. Meyer,
Domb, C., and D i . Hunter, 1965, Proc. Phys. M. Saile and T. Egami, 1983, J. Magn. &
Soc. (London) 86, 1147. Magn. Mater. 38, 240.
du Tr6molet de Lacheisserie, E., 1982, J. Magn. F/ihnle, M., J. Furthmfiller and G. Herzer, 1988a,
& Magn. Mater. 25, 251. J. Physique 49, Suppl. C8-1329.
du Tr6molet de Lacheisserie, E., and R. Yavari, Ffihnle, M., P. Braun, R. Reisser, M. Seeger and
1988, J. de Phys. (France) 12, Suppl. C8-1327. H. Kronmfiller, 1988b, J. Physique 49, Suppl.
Durand, J., and S.J. Poon, 1977, IEEE Trans. C8-1201.
Magn. MAG-13, 1556. Felsch, W., 1969, Z. Phys. 219, 280.
Durand, J., and M. Yung, 1977, in: Amorphous Felsch, W., 1970a, Z. Angew. Phys. 29, 218.
Magnetism II, eds R.A. Levy and R. Hase- Felsch, W., 1970b, Z. Angew. Phys. 30, 275.
gawa (Plenum Press, New York) p. 275. Felsch, W., S.G. Kushniv, K. Samwer, M.
Dyson, J.F., 1956, Phys.Rev. 102, 1217, 1230. Schr6der, R. van den Berg and H. von L6h-
Ebert, H., P. Strange and B.L. Gyorffy, 1988, J. neysen, 1982, Z. Phys. B 48, 99.
de Phys. (France) 12, Suppl. C8-31. Ferchmin, A.R., and S. Kobe, 1983, Amorphous
Egami, T., 1978a, J. Mater. Sci 13, 2587. Magnetism and Metallic Magnetic Material
Egami, T., 1978b, Mater. Res. Bull. 13, 2589. Digest, in: Selected Topics in Solid State Phys-
Egami, T., 1981, Ann. NY Acad. Sci. 371, 238. ics, ed. E.P. Wohlfarth (North-Holland,
Egami, T., 1983a, J. Magn. & Magn. Mater. Amsterdam).
31-34, 1571. Fernandez-Baca, J.A., J.W. Lynn, J.J. Rhyne and
Egami, T., 1983b, in: Amorphous Metallic Alloys, G.E. Fish, 1987, Phys. Rev. B 36, 8497.
ed. F.E. Luborsky (Butterworths, London) Ferrer, R., R. Harris, D. Zobin and M.J. Zucker-
p. 100. mann, 1978, Solid State Commun. 26, 451.
Egami, T., 1984, Magnetic amorphous alloys, in: Fert, A., and R. Asomoza, 1979, J. Appl. Phys.
Rep. Progr. Phys. 47, 1601. 50, 1886.
Egami, T., P.J. Flanders and C.D. Graham Jr, Fert, A., and I.A. Campbell, 1978, J. Phys. F 8,
1975, AIP Conf. Proc. 24, 697. L57.
Egami, T., C.D. Graham Jr, W. Dmowski, P. Fert, A., and A. Friedrich, 1976, Phys. Rev. B 13,
Zhou, P.J. Flanders, E.E. Marinero, H. 397.
Notarys and C. Robinson, 1987, IEEE Trans. Fert, A., and A. Friedrich, 1977, Physica B 86-88,
Magn. MAG-23, 2269. 491.
Elliott, R.J., J.A. Krumhansl and P.L. Leath, Fert, A., R. Asomoza, I.A. Campbell, R. Meyer
1974, Rev. Mod. Phys. 46, 465. and H. Jouve, 1977, Comptes-Rends ~ l'Aca-
Englich, J., H. Lfitgemeier, M.W. Pieper, V. Nek- d6mie des Sciences B 285, 113.
vasil and P. Nov~tk, 1985, Solid State Com- Fibich, M., 1990, J. Phys. C: Condens. Matter 2,
mun. 56, 825. 2987.
Erskine, J.L., 1975, AIP Conf. Proc. 24, 190. Fiedler, R., and G. Schirmer, 1988, Thin Solid
Erskine, J.L., and E.A. Stern, 1973a, Phys. Rev. Films 167, 281.
B 8, 1239. Finney, J.L., 1970, Proc. R. Soc. A 319, 479.
MAGNETIC AMORPHOUS ALLOYS 441

Fish, G.E., 1985, IEEE Trans. Magn. MAG-21, Fukamichi, K., H. Komatsu, T. Goto and H.
1986. Wakabayashi, 1988, Physica B 149, 276.
Fish, G.E., and C.H. Smith, 1986, in: Soft and Fukamichi, K., H. Komatsu and H. Wakabayashi,
Hard Magnetic Materials with Applications, 1989a, 4th Int. Conf. on Phys. and Magn.
ed. J.A. Salsgiver (ASM, Metals Park, OH) Mater. (ICPMM), Poland, 1988, World Scien-
p. 7. tific, Singapore, 1989.
Foo, E., H. Amar and M. Ausloos, 1971, Phys. Fukamichi, K., H. Komatsu, T. Goto and H.
Rev. B 4, 3350. Wakabayashi, 1989b, MRS Int. Mtg. Adv.
Foo, E.-N., and D.-H. Wu, 1972, Phys. Rev. B 5, Mater., Vol. 11, eds M. Homma and Y. Ima-
98. koa (Materials Research Society, Pittsburgh,
Freitas, P.P., T.S. Plaskett, J.M. Moreira and V.S. PA) p. 285.
Amaral, 1988, J. Appl. Phys. 64, 5453. Furthmfiller, J., M. Ffihnle and G. Herzer, 1986,
Fr6my, M.A., D. Gignoux and A. Li6nard, 1984, J. Phys. F 16, L255.
J. Magn. & Magn. Mater. 44, 263. Furthmfiller, J., M. F/ihnle and G. Herzer, 1987a,
Friedberg, R., and D.I. Paul, 1975, Phys. Rev. J. Magn. & Magn. Mater. 69, 79.
Lett. 34, 1234. Furthmfiller, J., M. Ffihnle and G. Herzer, 1987b,
Friedel, J., 1958, Nuovo Cimento, Suppl. 7(2), J. Magn. & Magn. Mater. 69, 89.
287. Gambino, R.J., and J.J. Cuomo, 1978, J. Vac. Sci.
Friedt, J.M., M. Maurer, J.P. Sanchez, A. Berrada, Technol. 15, 296.
A. Qachao, P. Panissod and J. Durand, 1980, Gambino, R.J., and T.R. McGuire, 1983, IEEE
J. de Phys. 41, C8-638. Trans. Magn. MAG-19, 1952.
Fujimori, H., M. Kikuchi, Y. Obi and T. Masu- Gambino, R.J., and T.R. McGuire, 1986, J. Magn.
moto, 1976, Sci. Rep. Instrum. (Tohoku Uni- & Magn. Mater. 54-57, 1365.
versity) A 26, 36. Gambino, R.J., P. Chaudhari and J.J. Cuomo,
Fujimori, H., H. Morita, Y. Obi and S. Ohta, 1974, AIP Conf. Proc. 18, 578.
1977, in: Amorphous Magnetism, Vol. II, eds Gambino, R.J., T.R. McGuire and K. Fukamichi,
R.A. Levy and R. Hasegawa (Plenum Press, 1981, J. Appl. Phys. 52, 2190.
New York) p. 393. Gambino, R.J., T.R. McGuire, T.S. Plaskett and
Fujimori, H., K. Nakanishi, H. Hiroyoshi and W. Reim, 1986, IEEE Trans. Magn. MAG-
N.S. Kazama, 1982, J. Appl. Phys. 53, 7792. 22, 1227.
Fujimori, H., N.S. Kazama, K. Hirose, J. Zhang, Gangulee, A., and R.J. Kobliska, 1978a, J. Appl.
Phys. 49, 4169.
H. Morita, I. Sato and H. Sugawara, 1984, J.
Gangulee, A., and R.J. Kobliska, 1978b, J. Appl.
Appl. Phys. 55, 1769.
Phys. 49, 4896.
Fukamichi, K., and H. Hiroyoshi, 1985, Sci. Rep.
Gangulee, A., and R.C. Taylor, 1978, J. Appl.
RITU (Sci. Rep. of the Res. Inst., Tohoku
Phys. 49, 1762.
University) A 32, 154.
Garoche, P., and A.P. Malozemoff, 1984, Phys.
Fukamichi, K., M. Kikuchi, H. Hiroyoshi and T. Rev. B 29, 226.
Masumoto, 1978, in: Rapidly Quenched Gaskell, P.H., 1983, in: Glassy Metals, eds. H.
Metals III, Vol. 2, ed. B. Cantor (The Metal Beck and H.-J. Gfintherodt (Springer, Berlin)
Society, London) p. 325. p. 5.
Fukamichi, K., H. Hiroyoshi, T. Kaneko, T. Geny, J.-F., G. Marchal, Ph. Mangin, Chr. Janot
Masumoto and K. Shirakawa, 1982, J. Appl. and M. Piecuch, 1982, Phys. Rev. B 25, 7449.
Phys. 53, 8107. Gibbs, M.R.J., and J.E. Evetts, 1982, Proc. 4th
Fukamichi, K., H. Hiroyoshi, K. Shirakawa, T. Int. Conf. on Rapidly Quenched Metals, Sen-
Masumoto and T. Kaneko, 1986a, IEEE dai, 1981, eds T. Masumoto and K. Suzuki
Trans. Magn. MAG-22, 424. (Japan Institute of Metals) Vol. 2, p. 479.
Fukamichi, K., T. Goto, Y. Satoh, T. Sakakibara, Gibbs, M.R.J., J.E. Evetts and J.A. Leake, 1983,
S. Todo, U. Mizutani and Y. Hoshino, 1986b, J. Mater. Sci. 18, 278.
IEEE Trans. Magn. MAG-22, 555. Gignoux, D., D. Givord and A. Li6nard, 1982, J.
Fukamichi, K., T. Goto and U. Mizutani, 1987a, Appl. Phys. 53, 2321.
IEEE Trans. Magn. MAG-2, 3590. Givord, D., P. Tenaud and T. Viadieu, 1988,
Fukamichi, K., Y. Satoh and H. Komatsu, 1987b, IEEE Trans. Magn. MAG-24, 1921.
IEEE Trans. Magn. MAG-23, 2548. Givord, D., Q. Lu, M.F. Rossignol, P. Tenaud
442 P. HANSEN

and T. Viadieu, 1990, J. Magn. & Magn. Hansen, P., and M. Hartmann, 1986, J. Appl.
Mater. 83, 183. Phys. 59, 859.
Glang, R., 1970, in: Handbook of Thin Film Hansen, P., and H. Heitmann, 1989, IEEE Trans.
Technology, eds L.I. Maissel and R. Glang Magn. 25, 4390.
(McGraw Hill, New York) p. 1. Hansen, P., and J.-P. Krumme, 1984, Thin Solid
Graczyk, J.F., 1978, J. Appl. Phys. 49, 1738. Films 114, 69.
Gr~itzer, D., 1978, J. Magn. & Magn. Mater. 9, Hansen, P., and M. Urner-Wille, 1979, J. Appl.
91. Phys. 50, 7471.
Greidanus, F.J.A.M., and S. Klahn, 1989, Z. Hansen, P., and K. Witter, 1988, IEEE Trans.
Angew. Chem. 101, 243. Magn. MAG-24, 2317.
Greidanus, F.J.A.M., B.AJ. Jacobs, F.J.A. den Hansen, P., M. Hartmann and K. Witter, 1987,
Broeder, J.H.M. Spruit and M. Rosenkranz, J. Magn. Soc. Jpn. 11, Suppl., 257.
1989a, Appl. Phys. Lett. 54, 963. Hansen, P., C. Clausen, G. Much, M. Rosenkranz
Greidanus, F.J.A.M., B.A.J. Jacobs, J.H.M. Spruit and K. Witter, 1989, J. Appl. Phys. 66, 756.
and S. Klahn, 1989b, IEEE Trans. Magn. Hansen, P., S. Klahn, C. Clausen, G. Much and
MAG-25, 3524. K. Witter, 1991, J. Appl. Phys. 69.
Greidanus, F.J.A.M., W.B. Zeper, FJ.A. den Harris, A.B., 1974, J. Phys. C 7, 1671.
Broeder, W.F. Godlieb and P.F. Carcia, Harris, R., 1980, J. Phys. F 10, 2545.
1989c, Appl. Phys. Lett. 54, 2481. Harris, R., M. Plischke and M.J. Zuckermann,
Gr6ger, B., and H. Kronmfiller, 198l, Appl. Phys. 1973, Phys. Rev. Lett. 31, 160.
24, 287. Hartmann, M., J.J.M. Braat and B.A.J. Jacobs,
Gr6ssinger, R., H. Sassik, R. Wezulek and T. 1984a, IEEE Trans. Magn. MAG-20, 1013.
Tarnoczi, 1988, J. Physique 49, Suppl. C8- Hartmann, M., P. Hansen and P. Willich, 1984b,
1337. J. Appl. Phys. 56, 2870.
Gubbens, P.C.M., J.H.F. Appeldorn, A.M. van Hartmann, M., B.A.J. Jacobs and J.J.M. Braat,
der Kraan and K.H.J. Buschow, 1974, J. Phys. 1985a, Philips Tech. Rev. 42, 37.
F 4, 921. Hartmann, M., K. Witter and P. Willich, 1985b,
Gfintherodt, H.J., M. Mfiller, R. Oberle, C. IEEE Trans. Magn. MAG-21, 2044.
Hauser, H.U. Kfinzi, M. Liard and R. Mfiller, Hartmann, M., S. Klahn and P. Willich, 1987,
1978, Inst. Phys. Conf. Ser. 39, 436. IEEE Trans. Magn. MAG-23, 2946.
Gyorgy, E.M., 1978, in: Metallic Glasses, eds J.J. Hasegawa, H., 1979, J. Phys. Soc. Jpn. 46, 1504.
Gilman and HJ. Leamy (American Society Hasegawa, H., 1980a, J. Phys. Soc. Jpn. 49, 178.
for Metals, Metals Park, OH) p. 275. Hasegawa, H., 1980b, J. Phys. Soc. Jpn. 49, 963.
Hadjipanayis, G., D.J. Sellmyer and B. Brandt, Hasegawa, H., 1983, J. Phys. F 13, 2655.
1981, Phys. Rev. B 23, 3349. Hasegawa, H., 1984, J. Phys. F 14, 1235.
Hajjar, R., and M. Mansuripur, 1989, IEEE Hasegawa, R., 1975, J. Appl. Phys. 46, 5263.
Trans. Magn. MAG-25, 4021. Hasegawa, R., and R. Ray, 1978, J. Appl. Phys.
Handrich, K., 1969, Phys. Status Solidi 32, K55. 49, 4174.
Handrich, K., 1972, Phys. Status Solidi 53, K17. Hasegawa, R., and R. Ray, 1979, J. Appl. Phys.
Handrich, K., and S. Kobe, 1980, Amorphe 50, 1586.
Ferro- and Ferrimagnetika (Physik-Verlag, Hasegawa, R., and C.C. Tsuei, 1971a, Phys. Rev.
Weinheim). B 2, 1631.
Hansen, P., 1987, J. Appl. Phys. 62, 216. Hasegawa, R., and C.C. Tsuei, 1971b, Phys. Rev.
Hansen, P., 1988a, in: Landolt-B6rnstein, Numer- 3, 214.
ical Data and Functional Relationships in Hasegawa, R., R.J. Gambino and R. Ruf, 1975a,
Science and Technology, New Series, Group Appl. Phys. Lett. 27, 512.
III, Vol. 19, ed. H.P.J. Wijn (Springer, Berlin) Hasegawa, R., B.E. Argyle and L.-J. Tao, 1975b,
p. 136. AIP Conf. Proc. 24, 110.
Hansen, P., 1988b, J. Appl. Phys. 63, 2371. Hasegawa, R., R.C. O'Handley, L. Tanner, R.
Hansen, P., 1990, J. Magn. & Magn. Mater. 83, Ray and S. Kavesh, 1976, Appl. Phys. Lett.
6. 29, 219.
Hansen, P., and M. Hartmann, 1985, in: Physics Hashimoto, S., and Y. Ochiai, 1990, J. Magn. &
of Magnetic Materials (World Scientific, Magn. Mater. 88, 211.
Singapore) p. 158. Hashimoto, S., Y. Ochiai, M. Kaneko, K. Wata-
MAGNETIC AMORPHOUS ALLOYS 443

nabe and K. Aso, 1987, IEEE Trans. Magn. Hirscher, M., T. Egami and E.E. Marinero, 1990,
MAG-23, 2278. J. Appl. Phys. 67, 4932.
Hatwar, T.K., and D. Majumdar, 1988, IEEE Hoffmann, D.W., and J.A. Thornton, 1980, J. Vac.
Trans. Magn. MAG-24, 2449. Sci. Technol. 17, 380.
Hauser, J.J., 1975, Phys. Rev. B 12, 5160. Holtslag, AM.M, 1989, J. Appl. Phys. 66, 1530.
Hauser, J.J., 1986, Phys. Rev. B 34, 3212. Honda, S., and M. Yoshiyama, 1988a, Jpn. J.
Heiman, N., and N. Kazama, 1978a, Phys. Rev. Appl. Phys. 27, 1687.
B 17, 2215. Honda, S., and M. Yoshiyama, 1988b, Jpn. J.
Heiman, N., and N. Kazama, 1978b, J. Appl. Appl. Phys. 27, 2073.
Phys. 49, 1686. Honda, S., M. Nawate, M. Ohkoshi and T.
Heiman, N., and N. Kazama, 1979, Phys. Rev. B Kusuda, 1985, J. Appl. Phys. 57, 3204.
19, 1623. Hong, M., E.M. Gyorgy, R.B. van Dover, S.
Heiman, N., and K. Lee, 1975, AIP Conf. Proc. Nakahara, D.D. Bacon and P.K. Gallagher,
24, 108. 1986, J. Appl. Phys. 59, 551.
Heiman, N., and K. Lee, 1976, AIP Conf. Proc. Hoshi, Y., M. Naoe and S. Yamanaka, 1982, J.
34, 319. Appl. Phys. 53, 2344.
Heiman, N., K. Lee and I. Potter, 1976a, AIP Hubbard, J., 1963, Proc. R. Soc. A 276, 238.
Conf. Proc. 29, 130. Hubbard, J., 1979a, Phys. Rev. B 19, 2626.
Heiman, N., K. Lee, I. Potter and S. Kirkpatrick, Hubbard, J., 1979b, Phys. Rev. B 20, 4584.
1976b, J. Appl. Phys. 47, 2634. Hubbard, J., 1981a, Phys. Rev. B 23, 5974.
Heitmann, H., M. Hartmann, M. Rosenkranz and Hubbard, J., 1981b, in: Electron Correlation and
H.J. Tolle, 1985, J. Physique 46, Suppl. C6-9. Magnetism in Narrow-Band Systems, ed. T.
Heitmann, H., A.M.J. Spruijt, P. Willich and H. Moriya (Springer, Berlin) p. 29.
Wilting, 1987a, J. Appl. Phys. 61, 3343. Huggins, H.A., and M. Guvitcb, 1983, J. Vac. Sci.
Heitmann, H., M. Hartmann, S. Klahn, M. A 1, 77.
Rosenkranz, H.J. Tolle and P. Willich, 1987b, Hfiller, K., 1986, J. Magn. & Magn. Mater. 61,
J. Appl. Phys. 61, 3331. 347.
Henderson, D., M.H. Brodsky and P. Chaudhari, Hfiller, K., and G. Dietz, 1985, J. Magn. & Magn.
1974, Appl. Phys. Lett. 25, 641. Mater. 50, 250.
Herd, S.R., 1977, Phys. Status Solidi A 44, 363. Hfiller, K., G. Dietz and H. Bestgen, 1985, J.
Herd, S.R., 1978, J. Appl. Phys. 49, 1747. Magn. & Magn. Mater. 53, 275.
Huth, B.G., 1974, IBM J. Res. Develop. 18, 100.
Herd, S.R., 1979, J. Appl. Phys. 50, 1645.
Ichihara, K., Y. Mizusawa and N. Yasuda, 1988a,
Hernando, A., V. Madurga, C. Nnnez de Villavi-
J. Appl. Phys. 63, 5154.
cencio and M. V~squez, 1984, Appl. Phys.
Ichihara, K., Y. Mizusawa, N. Yasuda and T.
Lett. 45, 802.
Suzuki, 1988b, J. Appl. Phys. 63, 5154.
Herzer, G., M. Ffihnle, T. Egami and H. Kron- Ichikawa, T., 1975, Phys. Status Solidi A 29, 293.
mfiller, 1980, Phys. Status Solidi B 101, 713. Iijima, T, 1987, Appl. Phys. Lett. 50, 1835.
Herzer, G., M. Ffihnle, T. Egami and H. Kron- Iijima, T., 1988, J. Appl. Phys. 64, 5170.
mfiller, 1981, J. Appl. Phys. 52, 1794. Iijima, T., O. Ishii and I. Hatakeyama, 1989,
Hilzinger, H.R., 1977, Appl. Phys. 12, 253. Appl. Phys. Lett. 54, 2376.
Hilzinger, H.R., 1985, IEEE Trans. Magn. MAG- Imamura, N., S. Tanaka, F. Tanaka and Y.
21, 2021. Nagao, 1985, IEEE Trans. Magn. MAG-21,
Hilzinger, H.R., 1990, J. Magn. & Magn. Mater. 1607.
83, 370. Inukai, T., M. Matsuoka and K. Ono, 1986, Appl.
Hirosawa, S., 1989, IEEE Trans. Magn. MAG- Phys. Lett. 49, 52.
25, 3437. Ishikawa, Y., K. Yamada, K. Tajima and K.
Hirosawa, S., and M. Sagawa, 1988, J. Appl. Fukamichi, 1981, J. Phys. Soc. Jpn. 50, 1958.
Phys. 64, 5553. Ishio, S., 1988a, J. Physique 49, Suppl. C8-1347.
Hirosawa, S., K. Tokuhara, Y. Matsuura, H. Ishio, S., 1988b, J. Physique 49, Suppl. C8-1345.
Yamamoto, S. Fujimura and M. Sagawa, Ishio, S., and T. Miyazaki, 1988, IEEE Trans. J.
1986, J. Magn. & Magn. Mater. 61, 363. Magn. Jpn. 3, 434.
Hiroyoshi, H., K. Noguchi, K. Fukamichi and Y. Ishio, S., S. Takahashi, T. Miyazaki and S. Kado-
Nakagawa, 1985, J. Phys. Soc. Jpn. 54, 3554. waki, 1990, IEEE Transl. J. Magn. Jpn 5, 458.
444 P. HANSEN

Itoh, A., H. Uekusa, Y. Tarusawa, F. Inoue and Kazama, N.S., N. Heiman and R.L. White, 1978b,
K..Kawanishi, 1983, J. Magn. & Magn. Mater. J. Appl. Phys. 49, 1706.
35, 241. Kazama, N.S., N. Fujimori and H. Watanabe,
Itoh, S., K. Aso, Y. Makino and S. Uedaira, 1980, 1980, J. Magn. & Magn. Mater. 15-18, 1423.
Appl. Phys. Lett. 37, 665. Keffer, F., 1966, in: Ferromagnetism, ed. H.P.J.
Izuyama, T., and R. Kubo, 1964, J. Appl. Phys. Wijn (Springer, Berlin) Vol. XVIII, 2, p. 1.
35, 1074. Kirchmayr, H.R., and W. Steiner, 1971, J. Phy-
Jaccarino, V., and L.R. Walker, 1965, Phys. Rev. sique 32, Suppl. C1-45.
Lett. 15, 259. Kissinger, H.E., 1957, Anal. Chem. 29, 1702.
Jgger, E., 1977, Phys. Status Solidi B 80, K81. Klahn, S., 1990, Thesis, Hamburg, FRG.
Jen, S.U., and S.M. Yang, 1988, J. Appl. Phys. Klahn, S., and P. Hansen, 1991, J. Appl. Phys.,
63, 4303. to be published.
Jergel, M., G. Vlasfik and P. Duhaj, 1989, Phys. Klahn, S., M. Hartmann, K. Witter and H. Heit-
Status Solidi A 111, 597. mann, 1987, in: Magnetic Properties of Amor-
Jones, R.C., and G.J. Yates, 1975, J. Phys. C 8, phous Metals, eds A. Hernando, V. Madurga,
1705. M.C. Samchez-Trujilo and M. Verzquez
Jouve, H., J.P. Rebouillat and R. Meyer, 1976, (North-Holland, Amsterdam) p. 176.
AIP Conf. Proc. 29, 97. Klahn, S., H. Heitmann, M. Rosenkranz and H.J.
Kahn, F.J., P.S. Pershan and J.P. Remeika, 1969, Tolle, 1988, J. Physique 49, Suppl. C8-1711.
Phys. Rev. 186, 891. Klahn, S., H. Bentin, B. Kn6rr, H. Heitmann and
Kakehashi, Y., 1990a, Phys. Rev. B 40, 11059. H. Wilting, 1990a, J. Appl. Phys. 67, 1442.
Kakehashi, Y., 1990b, J. Magn. & Magn. Mater Klahn, S., P. Hansen and F.J.A.M. Greidanus,
90-91, 333. 1990b, Vacuum 41, 1160.
Kaneyoshi, T., 1986, J. Phys. Soc. Jpn. 55, 1430. Kobayashi, H., T. Ono and A. Tsushima, 1983,
Kaneyoshi, T., 1987, IEEE Trans. Magn. MAG- Appl. Phys. Lett. 43, 389.
23, 2987. Kobayashi, M., Y. Maeno, K. Oishi and K.
Kaneyoshi, T., and I. Tamura, 1984, Phys. Status Kawamura, 1988, Appl. Phys. Lett. 52, 510.
Solidi B 123, 525. Kobe, S., and K. Handrich, 1972, Phys. Status
K/istner, J., H.-J. Schink and E.F. Wassermann, Solidi B 54, 663.
1980, Solid State Commun. 33, 527.
Komatsu, H., and K. Fukamichi, 1989, Sci. Rep.
Kasuya, T., 1956, Progr. Theor. Phys. 16, 45.
RITU (Science Report of the Research Insti-
Katayama, T., and K. Hasegawa, 1982, in:
tutes, Tohoku University) A 34, 251.
Rapidly Quenched Metals IV, eds T. Masu-
Komatsu, H., K. Fukamichi, N. Saito and Y.
moto and K. Suzuki (Japan Institute of
Nakagawa, 1989, IEEE Transl. J. Magn. Jpn.
Metals, Sendai) p. 915.
4, 270.
Katayama, T., M. Hirano, Y. Koizumi, K.
Kawanishi and T. Tsushima, 1977, IEEE Kondo, J., 1976, Physica B 84, 207.
Trans. Magn. MAG-13, 1603. Korenman, V., 1985, J. Appl. Phys. 57, 3000.
Kaul, N.S., and T.V.S.M. Mohan Babu, 1989, J. Korenman, V., J.L. Murray and R.E. Prange,
Phys. C: Condens. Matter 1, 8509. 1977a, Phys. Rev. B 16, 4032.
Kaul, S.N., 1981a, IEEE Trans. Magn. MAG-17, Korenman, V., J.L. Murray and R.E. Prange,
1208. 1977b, Phys. Rev. B 16, 4048.
Kaul, S.N., 1981b, Phys. Rev. B 24, 6550. Korenman, V., J.L. Murray and R.E. Prange,
Kaul, S.N., 1981c, Phys. Rev. B 23, 1205. 1977c, Phys. Rev. B 16, 4058.
Kavesh, S., 1978, in: Metallic Glasses, eds H.J. Korn, D., W. Muner and G. Zibold, 1972, Phys.
Leamy and J.J. Gilman (American Society of Lett. A 47, 117.
Metals, Metals Park, OH) p. 36. Kouvel, J.S., and J.B. Comly, 1968, Phys. Rev.
Kazama, N.S., and N. Fujimori, 1982, in: Rapidly Lett. 20, 1237.
Quenched Metals IV, eds T. Masumoto and Kouvel, J.S., and M.E. Fisher, 1964, Phys. Rev.
K. Suzuki (Japan Institute of Metals, Sendal) A 136, 1626.
p. 799. Krause, J.K., T.C. Long, T. Egami and D.G. Onn,
Kazama, N.S., M. Mitera and T. Masumoto, 1980, Phys. Rev. B 21, 2886.
1978a, in: Rapidly Quenched Metals III, ed. Krey, U., S. Krompiewski and U. Krauss, 1990,
B. Cantor (Chameleon, London) Vol. 2, p. 164. J. Magn. & Magn. Mater. 86, 85.
MAGNETIC AMORPHOUS ALLOYS 445

Krinchik, G.S., and V.A. Artem'ev, 1968, Sov. LeGuillou, J.C., and J. Zinn-Justin, 1975, Appl.
Phys.-JETP 26, 1080. Phys. Lett. 39, 95.
Kronmtiller, H., 1970, Z. Angew. Phys. 30, 9. Lewis, B.G., I.W. Donald and H.A. Davies, 1977,
Kronmtiller, H., 1973, AIP Conf. Proc. 10, 1006. Proc. Int. Conf. on Solidification and Casting,
Kronm/iller, H., 1981a, J. Appl. Phys. 52, 1859. Sheffield.
Kronmfiller, H., 1981b, J. Magn. & Magn. Mater. Li, Z.G., and D.J. Smith, 1989, Appl. Phys. Lett.
24, 159. 55, 919.
Kronmiiller, H., 1981c, J. Appl. Phys. 52, 1859. Liebermann, H.H., 1983, in: Amorphous Metallic
Kronmiiller, H., and M. F/ihnle, 1980, Phys. Alloys, ed. F.E. Luborsky (Butterworths, Lon-
Status Solidi B 97, 513. don) p. 244.
Kronmfiller, H., and B. Gr6ger, 1981, J. Physique Liebermann, H.H., and C.D. Graham Jr, 1976,
42, 1285. IEEE Trans. Magn. MAG-12, 921.
Kronm/iller, H., K.-D. Durst and M. Sagawa, Li6nard, A., and J.P. Rebouillat, 1978, J. Appl.
1988, J. Magn. & Magn. Mater. 74, 291. Phys. 49, 1680.
Kryder, M.H., 1985, J. Appl. Phys. 57, 3913. Liniers, M., J. Flores, F.J. Bermejo, J.M. Gon-
Kryder, M.H., H.-P. Shieh and D.K. Hairston, z~tles, J.L. Vincent and J. Tejada, 1989, IEEE
1987, IEEE Trans. Magn. MAG-23, 165. Trans. Magn. MAG-25, 3363.
Kudo, T., T. Mizoguchi, N. Watanabe, N. Liou, S.H., and C.L. Chien, 1984, J. Appl. Phys.
Nimura, M. Misawa and K. Suzuki, 1978, J. 55, 1820.
Phys. Soc. Jpn. Lett. 45, 1773. Litterst, F.J., 1975, J. Physique 36, L197.
Kursumovic, A., R.W. Cahn and M.G. Scott, Logan, J., 1975, Phys. Status Solidi A 32, 361.
1980a, Scr. Metall. 14, 1245. Lubitz, P., J.H. Schelleng and C. Vittoria, 1976,
Kursumovic, A., M.G. Scott, E. Girt and R.W. AIP Conf. Proc. 29, 178.
Cahn, 1980b, Scr. Metall. 14, 1303. Luborsky, F.E., 1977, in: Amorphous Magnetism
Kusuda, T., S. Honda and M. Ohkoshi, 1982a, II, eds R. Hasegawa and R.A. Levy (Plenum
J. Appl. Phys. 53, 2338. Press, New York) p. 345.
Kusuda, T., S. Honda and M. Ohkoshi, 1982b, Luborsky, F.E., 1978, J. Magn. & Magn. Mater.
in: Rapidly Quenched Metals IV, eds T. 7, 143.
Masumoto and K. Suzuki (Japan Institute of Luborsky, F.E., 1980, in: Handbook of Ferromag-
Metals, Sendal) p. 877.
netic Materials, Vol. 1, ed. E.P. Wohlfarth
Labrune, M., S. Hamzaoui and I.B. Puchalska,
(North-Holland, Amsterdam) p. 451.
1982, J. Magn. & Magn. Mater. 27, 323.
Luborsky, F.E., and L.A. Johnson, 1981, J. Phy-
Labrune, M., S. Andrieu, F. Rio and P. Bernstein,
sique 41, Suppl. C8-820.
1989, J. Magn. & Magn. Mater. 80, 211.
Luborsky, F.E., and J.L. Walter, 1977a, IEEE
Lachowicz, H.K., 1984, J. Magn. & Magn. Mater.
Trans. Magn. MAG-13, 953.
41, 182.
Lachowicz, H.K., and H. Szymczak, 1984, J. Luborsky, F.E., and J.L. Walter, 1977b, IEEE
Magn. & Magn. Mater. 41, 327. Trans. Magn. MAG-13, 1635.
Laurent, D.G., J. Callaway and C.S. Wang, 1979, Luborsky, F.E., P.G. Frischmann and L.A. John-
Phys. Rev. B 20, 1134. son, 1978, J. Magn. & Magn. Mater. 8, 318.
Leamy, H.J., and A.G. Dirks, 1977, J. Phys. D Luborsky, F.E., J.T. Furey, R.E. Skoda and B.C.
10, L95. Wagner, 1985, IEEE Trans. Magn. MAG-21,
Leamy, H.J., and A.G. Dirks, 1978, J. Appl. Phys. 1618, 3184.
49, 3430. Lupi, S., 1988, IEEE Trans. Magn. MAG-64,
Leamy, H.J., and A.G. Dirks, 1979, J. Appl. Phys. 6041.
50, 2871. Maczymowicz, L.J., D. Dendorek and R. Zub-
Leamy, H.J., G.H. Gilmer and A.G. Dirks, 1980, erek, 1985, Thin Solid Films 125, 269.
in: Current Topics in Material Sciences, Vol. Maeda, H., H. Terauchi, N. Kamijo, M. Hida
6, ed. E. Kaldis (North-Holland, Amsterdam) and K. Osamura, 1981, in: Proc. 4th Int. Conf.
p. 311. on Rapidly Quenched Metals, eds T. Masu-
Lee, K., and N. Heiman, 1975, AIP Conf. Proc. moto and K. Suzuki (Institute of Metals,
24, 108. Seandai) p. 397.
Lee, S.R., A.E. Miller and H.A. Blackstead, 1986, Maeda, I., H. Yamauchi and H. Watanabe, 1976,
J. Appl. Phys. 60, 3982. J. Phys. Soc. Jpn. 40, 1559.
446 P. HANSEN

Maehata, Y., S. Tsunashima and S. Uchiyama, McGuire, T.R., and R.J. Gambino, 1987a, J.
1986, IEEE Trans. Magn. MAG-22, 1107. Magn. Soc. Jpn. 11, Suppl. 261.
Majumdar, A.K., and L. Berger, 1973, Phys. Rev. McGuire, T.R., and R.J. Gambino, 1987b, IEEE
B 7, 4203. Trans. Magn. MAG-23, 3211.
Malmh/ill, R., 1983, J. Appl. Phys. 54, 5128. McGuire, T.R., and M. Hartmann, 1985, IEEE
Malozemoff, A.P., 1983, IEEE Trans. Magn. Trans. Magn. MAG-21, 1644.
MAG-19, 1520. McGuire, T.R., and M. Hartmann, 1986, IEEE
Malozemoff, A.P., A.R. Williams, K. Terakura Trans. Magn. MAG-22, 1224.
and V.L. Moruzzi, 1983, J. Magn: & Magn. McGuire, T.R., and R.C. Taylor, 1979, J. Appl.
Mater. 35, 192. Phys. 50, 1605.
Malozemoff, A.P., A.R. Williams and V.L. Mor- McGuire, T.R., R.J. Gambino and R.C. Taylor,
uzzi, 1984, Phys. Rev. B 29, 1620. 1977, IEEE Trans. Magn. MAG-13, 1598.
Malterre, D., J. Durand, A. Siari and G. Marchal, McGuire, T.R., T. Mizoguchi, R.J. Gambino and
1988, J. Physique 49, Suppl. C8-1359. S. Kirkpatrick, 1978, J. Appl. Phys. 49, 1689.
Mansuripur, M., 1982, J. Appl. Phys. 53, 1660. McGuire, T.R., J.A. Aboaf and E. Klokholm,
Mansuripur, M., 1987, J. Appl. Phys. 61, 1580. 1980a, IEEE Trans. Magn. MAG-16, 905.
Mansuripur, M., and G.A.N. Connell, 1983, J. McGuire, T.R., R.J. Gambino and R.C.
Appl. Phys. 54, 4794. O'Handley, 1980b, in: The Hall Effect and its
Mansuripur, M., and G.A.N. Connell, 1984, J. Applications, eds F.L. Chien and C.R. Wast-
Appl. Phys. 55, 3049. gate (Plenum Press, New York) p. 137.
Mansuripur, M., and M.F. Ruane, 1986, IEEE McGuire, T.R., R.J. Gambino, A.E. Bell and G.J.
Trans. Magu. MAG-22, 33. Spr0kel, 1986, J. Magn. & Magn. Mater.
Mansuripur, M., G.A.N. Connell and J.W. Good- 54-57, 1387.
man, 1982, Appl. Opt. 21, 1106. McGuire, T.R., R.J. Gambino, T.S. Plaskett and
Marchal, G., D. Teirlinck, Ph. Mangin, Chr. Janot W. Reim, 1987, J. Appl. Phys. 61, 3352.
and J. Hfibsch, 1980, J. de Phys. (France) 41, Meiklejohn, W.H., 1986, Proc. IEEE 74, 1570.
C8-662. Meiklejohn, W.H., F.E. Luborsky and P.G.
Maringer, R.E., and C.E. Mobley, 1974, J. Vac. Frischmann, 1987, IEEE Trans. Magn. MAG-
Sci. Technol. 11, 1067. 23, 2272.
Maringer, R.E., and C.E. Mobley, 1978, in: Meydan, T., and K.J. Overshott, 1982, J. Appl.
Rapidly Quenched Metals III, Vol. 1, ed. B. Phys. 53, 8383.
Cantor (Chameleon, London) pp. 49. Meyer, R., and H. Kronmiiller, 1982, Phys. Status
Marzwell, N.I., 1977, J. Magn. & Magn. Mater. Solidi B 109, 693.
5, 67, Miedema, A.R., 1979, Z. Metallkd. 70, 345.
Maskiewicz, H., 1982, J. Appl. Phys. 53, 7765. Miedema, A.R., P.F. de Chfitel and F.R. de Boer,
Masui, S., T. Kobayashi, S. Tsunashima and S. 1980, Physica B 100, 1.
Uchiyama, 1984, IEEE Trans. Magn. MAG- Mimura, Y., N. Imamura and T. Kobayashi,
20, 1036. 1976a, IEEE Trans. Magn. MAG-12, 779.
Masumoto, T., S. Ohnuma, K. Shirakawa, M. Mimura, Y., N. Imamura and Y. Kushiro, 1976b,
Nose and K. Kobayashi, 1980, J. Physique J. Appl. Phys. 47, 3371.
41, C8-686. Mimura, Y., N. Imamura, T. Kobayashi, A.
Mathon, J., and E.P. Wohlfarth, 1968, Proc. R. Okada and Y. Kushiro, 1978, J. Appl. Phys.
Soc. A 302, 409. 49, 1208.
Matsuura, M., T. Fukunaga, K. Fukamichi and Misemer, D.K., 1988, J. Magn. & Magn. Mater.
K. Suzuki, 1988, Solid State Commun. 66, 72, 267.
333. Mitera, M., M. Naka, T. Masumoto, M. Kazama
Maurer, M., and J.M. Friedt, 1983, J. Phys. F 13, and K. Watanabe, 1978, Phys. Status Solidi
2175. A 49, K163.
McDaniel, T.W., and M. Mansuripur, 1987, IEEE Miyamoto, H., T. Niihara, H. Sukeda, M. Taka-
Trans. Magn. MAG-23, 2943. hashi, T. Nakao, M. Ojima and N. Ohta,
McGuire, T.R., and R.J. Gambino, 1978, IEEE 1989, J. Appl. Phys. 66, 6138.
Trans. Magn. MAG-14, 838. Miyazaki, T., and M. Takahashi, 1978, Jpn. J.
McGuire, T.R., and R.J. Gambino, 1979, J. Appl. Appl. Phys. 17, 1755.
Phys. 50, 7653. Miyazaki, T., T.X. Yang, K. Takakura and M.
MAGNETIC AMORPHOUS ALLOYS 447

Takahashi, 1986, J. Magn. & Magn. Mater. Mfiller, H.-R., E. J~iger and M. Bitterlich, 1977,
60, 211. Phys. Status Solidi B 80, K85.
Miyazaki, T., K. Hayashi, T. Okaki, M. Taka- Mfiller, K.-H., 1987, Phys. Rev. B 35, 7906.
hashi and T. Shimamori, 1987a, J. Magn. & Munakata, M., K. Goto and Y. Shimada, 1984,
Magn. Mater. 71, 83. Jpn. J. Appl. Phys. 23, L749.
Miyazaki, %, T. Izumo, X.-B. Yang and M. Murata, K.K., and S. Doniach, 1972, Phys. Rev.
Takahashi, 1987b, Proc. of ISPMM 1987, 403. Lett. 29, 285.
Miyazaki, T., K. Hayashi, S. Yamaguchi, M. Nagato, K., A. Kawamoto, T. Sato and T. Yorozu,
Takahashi, A. Yoshihara, T. Shimamori and 1988, J. Appl. Phys. 63, 3856.
T. Wakiyama, 1988, J. Magn. & Magn. Mater. Nagel, R.S., 1977, Phys. Rev. B 16, 1649.
75, 243. Nagel, R.S., and J. Taut, 1975, Phys. Rev. Lett.
Mizoguchi, T., 1981, in: Diffraction Studies on 35, 380.
Non-Crystalline Substances, eds I. Hargittai Nagy, I., T. Tarnoczi and Z. Frait, 1977, Physica
and W.J. Orville-Thomas (Elsevier, Amster- B 86-88, 764.
dam) p. 811. Naka, M., S. Tomizawa, T. Watanabe and T.
Mizoguchi, T., and G.S. Cargill III, 1979, J. Appl. Masumoto, 1976, in: Rapidly Quenched
Phys. 50, 3570. Metals, Sect. l, eds N.J. Grant and B.C.
Mizoguchi, T., T. Yamauchi and H. Miyajima, Giessen (MIT Press, Cambridge, MA) p. 273.
1973a, in: Amorphous Magnetism, eds H.O. Nakao, T., M. Ojima, Y. Miyamura, S. Okamine,
Hooper and A.M. de Graaf (Plenum Press, H. Sukeda, N. Ohta and Y. Takeuchi, 1987,
New York) p. 325. Jpn. J. Appl. Phys. 26, Suppl. 26-4, 149.
Mizoguehi, T., N. Uedo, K. Yamaguchi and H. Narita, K., J. Yamasaki and H. Fukunaga, 1979,
Miyajima, 1973b, J. Phys. Soc. Jpn. 34, 1691. J. Appl. Phys. 50, 7591.
Mizoguchi, T., T.R. McGuire, R.J. Gambino and Nieuwenhuizen, J.M., and H.B. Haanstra, 1966,
S. Kirkpatrick, 1977a, Physica B 86-88, 783. Philips. Tech. Rev. 27, 87.
Mizoguchi, T., T.R. McGuire, S. Kirkpatrick and Niihara, T., S. Takayama and Y. Sugita, 1985,
R.J. Gambino, 1977b, Phys. Rev. Lett. 38, 89. IEEE Trans. Magn. MAG-21, 1638.
Mizoguchi, T., T. Kudo, T. Irisawa, N. Watanabe, Niihara, T., F. Kirino, N. Ogihara, E. Koyama,
N. Nimura, M. Misawa and K. Suzuki, 1978, R. Sudou, K. Shigematsu and N. Ohta, 1988,
in: Rapidly Quenched Metals III, Vol. 2, ed IEEE Trans. Magn. MAG-24, 2437.
B. Cantor (Metals Society, London) p. 384. Nold, E., S. Steeb, P. Lamparter and G. Rainer-
Mohri, K., 1983, IEEE Trans. Magn. MAG-19, Harbach, 1980, J. Physique 41, C8-186.
2142. Oelhafen, P., E. Hauser, H.J. G/intherodt and K.
Mohri, K., 1984, IEEE Trans. Magn. MAG-20, Bennemann, 1979, Phys. Rev. Lett. 43, 1134.
942. Oelhafen, P., E. Hauser and H.J. Gfintherodt,
Mohri, K., and E. Sudoh, 1981, IEEE Trans. 1980, Solid State Commun. 35, 1017.
Magn. MAG-17, 1371. Ogawa, A., T. Katayama, M. Hirano and T.
Moorjani, K., and J.M.D. Coey, 1984, Magnetic Tsunashima, 1975, AIP Conf. Proc. 24, 575.
Glasses (Elsevier, Amsterdam). Oguchi, T., 1971, Statistical Theory of Magne-
Moorjani, K., and S.K. Ghatak, 1977, J. Phys. C tism, Syokabo, Tokyo.
10, 1027. Oguchi, T., K. Terakura and N. Hamada, 1983,
Moriya, T., and A. Kawabata, 1973a, J. Phys. J. Phys. F 13, 145.
Soc. Jpn. 34, 639. O'Handley, R.C., 1975, J. Appl. Phys. 46, 4996.
Moriya, T., and A. Kawabata, 1973b, J. Phys. O'Handley, R.C., 1976, AIP Conf. Proc. 29, 206.
Soc. Jpn. 35, 669. O'Handley, R.C., 1977a, Solid State Commun.
Moruzzi, V.L., and P.M. Marcus, 1988, Phys. 22, 485.
Rev. B 38, 1613. O'Handley, R.C., 1977b, in: Amorphous Magne-
Moruzzi, V.L., P.M. Marcus, K. Schwarz and P. tism II, eds R.A. Levy and R. Hasegawa
Mohn, 1986, Phys. Rev. B 34, 1784. (Plenum Press, New York) p. 379.
Mott, N.F., 1964, Adv. Phys. 13, 325. O'Handley, R.C., 1977c, Solid State Commun. 21,
Movchan, B.A., and A.V. Demchisin, 1969, Phys. 1119.
Met. Metallogr. 28, 83. O'Handley, R.C., 1978a, Phys. Rev. B 18, 930.
Mueller, R., K. Agyeman and C.C. Tsuei, 1980, O'Handley, R.C., 1978b, Phys. Rev. B 18, 2577.
Phys. Rev. B 22, 2665. O'Handley, R.C.,~1987a, J. Appl. Phys. 62, R15.
448 P. HANSEN

O'Handley, R.C., 1987b, J. Appl. Phys. 61, 3225. Pfeifer, F., and W. Kunz, 1982, IEEE Trans.
O'Handley, R.C., and L. Berger, 1978, in: Physics Magn. MAG-18, 1406.
of Transition Metals, eds M.J.G. Lee, J.M. Piecuch, M., C. Canot, G. Marchal and M. Verg-
Perz and E. Fawcett (Institute of Physics, nat, 1983, Phys. Rev. B 28, 1481.
Bristol) p. 477. Polk, D., 1972, Acta Metallogr. 20, 485.
O'Handley, R.C., and N.J. Grant, 1985, Proceed- Pond, R.B., R.E. Maringer and C.E. Mobley,
ings on Rapidly Quenched Metals, eds S. 1974, in: New Trends in Materials Fabrica-
Steeb and H. Warlimont (North-Holland, tion, ed. B.H. Kear, (Am. Soc. Metals, Metals
Amsterdam) p. 1125. Park, OH).
O'Handley, R.C., and M.O. Sullivan, 1981, J. Pooh, S.J., and J. Durand, 1977a, Phys. Rev. B
Appl. Phys. 52, 1841. 16, 316.
O'Handley, R.C., R. Hasegawa, R. Ray and Poon, S.J., and J. Durand, 1977b, in: Amorphous
C.-P. Chou, 1976a, Appl. Phys. Lett. 29, 330. Magnetism II, eds R.A. Levy and R. Haseg-
O'Handley, R.C., R. Hasegawa, R. Ray and awa (Plenum Press, New York) p. 245.
C.-P. Chou, 1976b, J. Appl. Phys. 48, 2095. Poon, S.J., and J. Durand, 1978, Phys. Rev. B 18,
O'Handley, R.C., L.I. Mendelsohn and E.A. Nes- 6253.
bitt, 1976c, IEEE Trans. Magn. MAG-12, 942. Prange, R.E., and V. Korenman, 1975, AIP Conf.
O'Handley, R.C., R. Hasegawa, R. Ray and Proc. 24, 325.
C.-P. Chou, 1977, J. Appl. Phys. 48, 2095. Prange, R.E., and V. Korenman, 1976, AIP Conf.
O'Handley, R.C., M.C. Narasimhan and M.O. Proc. 29, 321.
Sullivan, 1979, J. Appl. Phys. 50, 1633. Prange, R.E., and V. Korenman, 1979a, Phys.
Ohnuma, S., A. Kumimoto and T. Masumoto, Rev. B 19, 4691.
1988, IEEE Trans. Magn. MAG-24, 2551. Prange, R.E., and V. Korenman, 1979b, Phys.
Okamine, S., N. Ohta and Y. Sugita, 1985, IEEE Rev. B 19, 4698.
Trans. Magn. MAG-21, 1641. Prasad, B.B., A.K. Bhatnagar and R. Jaganna-
Okuno, H., S. Matsushita and Y. Sakurai, 1981, than, 1980, Solid State Commun. 36, 661.
IEEE Trans. Magn. MAG-17, 2831. Primak, W., 1955, Phys. Rev. 100, 1677.
Orehotsky, J., and K. Schr6der, 1972, J. Appl. Prinz, G.A., W.G. Maisch, P. Lubitz, D.W.
Phys. 43, 2413. Forester and J.J. Krebs, 1981, IEEE Trans.
Osamura, K., R. Suzuki, T. Murakami and S. Magn. MAG-17, 3232.
Takayama, 1981, J. Mater. Sci. 16, 957. Ramesh, R., and K. Srikrishna, 1988, J. Appl.
Osato, Y., H. Kawade, E. Fujii, N. Kasama and Phys. 64, 6406.
T. Kobayashi, 1987, European Patent Appli- Rao, K.V., H. Gudmundsson, H.U. Astr6n and
cation, EPO258978 AZ. H.S. Chert, 1979, J. Appl. Phys. 50, 1592.
Ounadjela, K., G. Suran and F. Machizaud, 1989, Rapp, O., J.E. Grindberg and K.V. Rao, 1978, J.
Phys. Rev. B 40, 578. Appl. Phys. 49, 1733.
Pacala, T.J., I.S. McDermid and J.B. Lauden- Raskine, D., and C.H. Smith, 1983, in: Amor-
slager, 1984, IEEE Conf. Rec. 16th Power phous Metallic Alloy, ed. F.E. Luborsky (But-
Modulator Symp., Arlington, VA (IEEE, New terworths, London) p. 381.
York) p. 250. Ray, R., 1978, U.S. Patent No. 4, 067, 732 (Jan.
Pan, D., and D. Turnbull, 1974, J. Appl. Phys. 10, 1978).
45, 1406. Read, D.A., T. Moyo, S. Jassim, R.A. Dunlap and
Pappa, C., 1979, T6se de Troisi6me Cycle, Paris. G.C. Hallam, 1989, J. Magn. & Magn. Mater.
Patterson, J.D., G.R. Gruzalski and D.J. Sellmyer, 82, 87.
1978, Phys. Rev. B 18, 1377. Rebouillat, J.P., A. Li6nard, J.M.D. Coey, R.
Paul, D.I., 1976, AIP Conf. Proc. 29, 545. Arrese-Boggiano and J. Chappert, 1977, Phys-
Paul, D.I., 1982, J. Appl. Phys. 53, 1649. ica B 86-88, 773.
Pauthenet, R., 1982a, J. Appl. Phys. 53, 2029. Reim, W., and J. Schoenes, 1990, in: Handbook
Pauthenet, R., 1982b, J. Appl. Phys. 53, 8187. of Ferromagnetic Materials, Vol. 5, eds K.H.J.
Pawellek, R., J. Furthmfiller and M. F/ihnle, 1988, Buschow and E.P. Wohlfarth (North-Holland,
J. Magn. & Magn. Mater. 75, 225. Amsterdam), p. 133.
Peierls, R.E., 1936, Proc. Cambridge Philos. Soc. Reim, W., O.E. Hfisser, J. Schoenes, E. Kaldis, P.
32, 477. Wachter and K. Seiler, 1984, J. Appl. Phys.
Pfeffer, K.-H., 1967, Phys. Status Solidi 21, 857. 55, 2155.
MAGNETIC AMORPHOUS ALLOYS 449

Reisser, R., M. Ffihnle and H. Kronmiiller, 1988, Schultz, M.D., and M.H. Kryder, 1990, J. Appl.
J. Magn. & Magn. Mater. 75, 45. Phys. 68, 5293.
Rhyne, J.J., 1974, Proc. Conf. on Cryst. Electr. Seeger, M., and H. Kronmtiller, 1989, J. Magn.
Field Effects, in: Metals and Alloys, Montreal, & Magn. Mater. 78, 393.
p. 15. Shen, D.H., Y. Mizokawa, H. Iwasaki, D.F. Shen,
Rhyne, J.J., 1976, AIP Conf. Proc. 29, 182. T. Numata and S. Nakamura, 1981, Jpn. J.
Rhyne, J.J., and G.E. Fish, 1985, J. Appl. Phys. Appl. Phys. 20, L757.
57, 3407. Sherwood, R.C., E.M. Gyorgy, H.S. Chen, S.D.
Rhyne, J.J., S.J. Pickart and H.A. Alperin, 1974a, Ferris, G. Norman and H.J. Leamy, 1975,
AIP Conf. Proc. 18, 563. AIP Conf. Proc. 24, 745.
Rhyne, J.J., J.H. Schelleng and N.C. Koon, 1974b, Shiba, K., S. Tsunashima, S. Uchiyama and S.
Phys. Rev. B 10, 4672. Yoshino, 1986, IEEE Trans. Magn. MAG-22,
Rhyne, J.J., G.E. Fish and J.W. Lynn, 1982, J. 1104.
Appl. Phys. 53, 2316. Shieh, H.D., and M.H. Kryder, 1985, IEEE Trans.
Rio, F., P. Bernstein and M. Labrune, 1987, IEEE Magn. MAG-21, 1632.
Trans. Magn. MAG-23, 2266. Shimada, Y., and H. Kojima, 1978, J. Appl. Phys.
Roberts, G.E., W.L. Wilson Jr and C.H. Brown 49, 932.
Jr, 1977, IEEE Trans. Magn. MAG-13, 1535. Shirakawa, K., K. Fukamichi, T. Kaneko and T.
Rodmarcq, B., M. Piecuch, Chr. Janot, G. Masumoto, 1983, Sci. Rep. RITU A 31, 54.
Marchal and Ph. Mangin, 1980, Phys. Rev. B Shirakawa, T., Y. Nakajima, K. Okamoto, S.
21, 1911. Matsushita and Y. Sakurai, 1976, AIP Conf.
Rossnagel, S.M., and J.J. Cuomo, 1989, Thin Proc. 34, 349.
Solid Films 171, 143. Shull, W.H., D.G. Naugle, SJ. Pooh and W.L.
Rudermann, M.A., and C. Kittel, 1954, Phys. Rev. Johnson, 1978, Phys. Rev. B 18, 3263.
96, 99. Simonnin, P., R. Tourbot, B. Boucher, M. Perrin
Rugar, D., C.-J. Lin and R.H. Geiss, 1987, IEEE and J. Vanhaute, 1986, Phys. Status Solidi A
Trans. Magn. MAG-23, 2263. 98, 551.
Rugar, D., H.J. Mamin, P. Guethner, S.E. Lam- Simpson, W.A., 1974, Wiss. Z. Techn. Univ. Dres-
bert, J.E. Stern, I. McFadyen and T. Yogi, den 23, 1027.
1990, J. Appl. Phys. 68, 1169. Simpson, W.A., and D.R. Brambley, 1971, Phys.
Ryan, D.H., J.M.D. Coey and J.O. Str6m-Olsen, Status Solidi B 43, 291.
1987a, J. Magn. & Magn. Mater. 67, 148. Singh, M., C.S. Wang and J. Callaway, 1975,
Ryan, D.H., J.M.D. Coey, Z. Altounian and J.O. Phys. Rev. B 11, 287.
Str6m-Olsen, 1987b, Phys. Rev. B 35, 8630. Siratori, K., K. Nagayama, H. Ino, N. Saito and
Ryan, D.H., L.X. Liao and Z. Altounian, 1988, Y. Nakagawa, 1990, J. Magn. & Magn. Mater.
Solid State Commun. 66, 3. 83, 341.
Saito, N., Y. Togami and T. Morishita, 1986, Jpn. Smart, J.S., 1966, Effective Field Theories of
J. Appl. Phys. 25, L580. Magnetism (Saunders, Philadelphia).
Sanchez, J.M., J.L. Morgn-Lop6z, C. Leroux and Smit, J., 1955, Physica 21, 877.
M.C. Cadeville, 1989, J. Phys. C: Condens. Smit, J., 1958, Physica 24, 39.
Matter I, 491. Smith, C.H., 1988, J. Appl. Phys. 64, 6032.
Sato, K., and Y. Togami, 1983, J. Magn. & Magn. Somekh, R.E., 1984, J. Vac. Sci. Technol. A 2,
Mater. 35, 181. 1285.
Sato, R., N. Saito and Y. Togami, 1985, Jpn. J. Sostarich, M., S. Dey, M. Rosenberg, H.J.V. Niel-
Appl. Phys. 24, L266. sen and U. Gorres, 1982, in: Proc. 4th Int.
Sato, T., K. Nagato, A. Kawamoto, Y,. Yoneyama Conf. on Rapidly Quenched Metals, Vol. 2,
and T. Yorozu, 1988, IEEE Trans. Magn. eds T. Masumoto and K. Suzuki (Jpn Inst. of
MAG-24, 2305. Metals, Sendai) p. 1109.
Sawatzky, E., and G.B. Street, 1971, IEEE Trans. Stein, F., and G. Dietz, 1989, J. Magn. & Magn.
Magn. MAG-7, 374. Mater. 81, 294.
Schelleug, J.H., D.W. Forester, P. Lubitz and C. Stevens, K.W.H., 1952, Proc. Phys. Soc. A 65,
Vittoria, 1984, J. Appl. Phys. 55, 1805. 209.
Schultz, M., and M.H. Kryder, 1989, Appl. Phys. Steward, A.M., and W.A. Phillips, 1978, Philos.
Lett. 54, 1371. Mag. B 37, 561.
450 P. HANSEN

Stobiecki, T., 1982, IEEE Trans. Magn. MAG- and M.S. Green (Academic Press, New York)
18, 780. p. 259.
Stobiecki, T., and K. Kowalski, 1984, IEEE Tajima, K., 1971, J. Phys. Soc. Jpn. 31, 441.
Trans. Magn. MAG-20, 1344. Takagi, H., S. Tsunashima, S. Uchiyama and T,
Stobiecki, T., and F. Stobiecki, 1983, J. Magn. & Fujii, 1979, J. Appl. Phys. 50, 1642.
Magn. Mater. 35, 217. Takahashi, M., H. Fujimori and T. Miyazaki,
Stoner, E.C., and E.P. Wohlfarth, 1948, Philos. 1983, in: Recent Magnetics for Electronics, ed.
Trans. R. Soc. (London) A 240, 599. Y. Sakurai (Ohmsha and North-Holland,
Sugawara, E., T. Katayama and T. Masumoto, Amsterdam) p. 137.
1989, IEEE Transl. J. Magn. Jpn. 4, 204. Takahashi, M., T. Shimamori, T. Miyazaki, K.
Suits, J.C., R.H. Geiss, C.-J. Lin, D. Rugar and Hayashi and T. Wakiyama, 1987, IEEE Trans.
A.E. Bell, 1986, Appl. Phys. Lett. 49, 419. Magn. MAG-23, 3314.
Suits, J.C., R.H. Geiss, C.-J. Lin, D. Rugar and Takahashi, M., T. Niihara and N. Ohta, 1988a,
A.E. Bell, 1987, J. Appl. Phys. 61, 3509. J. Appl. Phys. 64, 262.
Suits, J.C., D. Rugar and C.-J. Lin, 1988, J. Appl. Takahashi, M., A. Yoshihara, T. Shimamori, T.
Phys. 64, 252. Wakiyama, T. Miyazaki, K. Hayashi and S.
Sumiyama, K., Y. Hashimoto, T. Yoshitake and Yamaguchi, 1988b, J. Magn. & Magn. Mater.
Y. Kanamura, 1983, J. Magn. & Magn. Mater. 75, 252.
31-34, 1495. Takayama, S., 1976, J. Mater. Sci. 11, 164.
Sumiyama, K., H. Yasuda and Y. Nakamura, Takayama, S., T. Niihara, K. Kaneko, Y. Sugita
1990, J. Phys. C: Condens. Matter 2, 3595. and M. Ojima, 1987, J. Appl. Phys. 61, 2610.
Suran, G., H. Daver and J.C. Bruyere, 1976, AIP Takeno, Y., M. Suwabe, T. Sakurai and K. Goto,
Conf. Proc. 29, 162. 1986, Jpn. J. Appl. Phys. 25, L657.
Suran, G., J. Sztern, J.A. Aboaf and T.R. Tanaka, F., S. Tanaka and N. Imamura, 1987,
McGuire, 1981, IEEE Trans. Magn. MAG- Jpn. J. Appl. Phys. 26, 231.
17, 3065. Tao, L.-J., R.J. Gambino, S. Kirkpatrick, J.J.
Suzuki, T., 1983, J. Magn. & Magn. Mater. 35, Cuomo and H. Lilienthal, 1974, AIP Conf.
232. Proc. 18, 641.
Suzuki, T., 1985, J. Magn. & Magn. Mater. 50, Taub, A.I., and F. Schaepen, 1980, Acta Metall.
265. 28, 1781.
Taylor, R.C., 1976, J. Appl. Phys. 47, 1164.
Suzuki, T., and T. Katayama, 1986, IEEE Trans.
Taylor, R.C., and A. Gangulee, 1976, J. Appl.
Magn. MAG-22, 1230.
Phys. 47, 4666.
Suzuki, T., A. Murakami and T. Katayama, 1987,
Taylor, R.C., and A. Gangulee, 1977, J. Appl.
IEEE Trans. Magn. MAG-23, 2958.
Phys. 48, 358.
Suzuki, T., C.-J. Lin and A.E. Bell, 1988, IEEE
Taylor, R.C., and A. Gangulee, 1980, Phys. Rev.
Trans. Magn. MAG-24, 2452. B 22, 1320.
Suzuki, Y., and T. Egami, 1983, J. Magn. & Taylor, R.C., T.R. McGuire, J.M.D. Coey and A.
Magn. Mater. 31-34, 1549. Gangulee, 1978, J. Appl. Phys. 49, 2885.
Suzuki, Y., and N. Ohta, 1988, J. Appl. Phys. 63, Teirlinck, D., 1981, Th6se d'Etat, Nancy, France.
3633. Teirlinck, D., M. Piecuch, J.F. Geny, G. Marchal,
Suzuki, Y., S. Takayama, F. Kirino and N. Ohta, Ph. Mangin and Chr. Janot, 1981, IEEE
1987, IEEE Trans. Magn. MAG-23, 2275. Trans. Magn. MAG-17, 3079.
Swierczek, J., and S. Szymura, 1988, Phys. Status Terakura, K., 1976, J. Phys. F 6, 1385.
Solidi A 109, 559. Terakura, K., 1977, Physica B 91, 162.
Szymczak, H., 1987, J. Magn. & Magn. Mater. Terakura, K., and J. Kanamori, 1971, Prog.
67, 227. Theor. Phys. 46, 1007.
Tago, A., C. Nishimura and K. Yanagisawa, 1985, Thornton, J.A., 1974, J. Vac. Sci. Technol. 11, 666.
IEEE Trans. Magn. MAG-21, 2032. Togami, Y., 1981, U.S. Patent, No. 4, 293, 621.
Tahir-Kheli, R.A., 1972a, Phys. Rev. B 6, 2808. Togami, Y., T. Tamaki, K. Sato and A. Yoshihara,
Tahir-Kheli, R.A., 1972b, Phys. Rev. B 6, 2826. 1983, J. Magn. & Magn. Mater. 31-34, 1497.
Tahir-Kheli, R.A., 1972c, Phys. Rev. B 6, 2838. Tsuei, C.C., 1978, Solid State Commun. 27, 691.
Tahir-Kheli, R.A., 1976, in: Phase Transitions and Tsuei, C.C., G. Longworth and S.C.H. Lin, 1968,
Critical Phenomena, Vol. 5B, eds C. Domb Phys. Rev. 170, 603.
MAGNETIC AMORPHOUS ALLOYS 451

Tsujimoto, H., M. Shouji, A. Saito, S. Matsushita Wagner, C.N.J., 1972, in: Liquid Metals, Chemis-
and Y. Sakurai, 1983, J. Magn. & Magn. try and Physics, ed. S.Z. Beer (Marcel Dekker,
Mater. 35, 199. New York) p. 257.
Tsujimoto, H., K. Saiki, S. Inokuchi and Y. Wagner, C.N.J., 1978, J. Non-Crystal Solids 31,
Sakurai, 1984, IEEE Trans. Magn. MAG-20, 1.
1027. Wakabayashi, H., K. Fukamichi, H. Komatsu, T.
Tsunashima, S., H. Takagi, K. Kamegaki, T. Fujii Goto, T. Sakakibara and K. Kuroda, 1987,
and S. Uchiyama, 1978, IEEE Trans. Magn. Proc. Int. Syrup. on Physics of Magnetic
MAG-14, 844. Materials (World Scientific, Singapore) p. 342.
Tsunashima, S., T. Shinoda, H. Miyatake and S. Wakabayasbi, H., T. Goto, K. Fukamichi and H.
Uchiyama, 1980, J. Appl. Phys. 51, 5901. Komatsu, 1990, J. Phys. C: Condens. Matter
Tsunashima, S., Y.J. Choe and S. Uchiyama, 2, 417.
1989, Technical Digest of the International Wan, H., C.P. Peng, R.Y. Fang and D.S. Dai,
Symposium on Optical Memory 1989 1990, Phys. Rev. B 42, 5067.
(ISOM89) Kobe, p. 27D-1. Wang, C.S., and J. Callaway, 1973, Phys. Rev. B
Tsutsumi, K., Y. Fujii, M. Komori, T. Numata 9, 4897.
and Y. Sakurai, 1983, IEEE Trans. Magn. Wang, Y.J., and Q.W. Leng, 1990, Phys. Rev. B
MAG-19, 1760. 41, 651.
Tsuya, N., K.I. Arai, Y. Shiraga, M. Yamada and Warlimont, H., and R. Boll, 1982, J. Magn. &
T. Masumoto, 1975, Phys. Status Solidi A 31, Magn. Mater. 26, 97.
557. Waseda, Y., 1980, The Structure of Non-Crystal-
Twarowski, K., and H.K. Lachowicz, 1979, Phys. line Materials (McGraw-Hill, New York).
Status Solidi A 53, 599. Waseda, Y., and K.T. Aust, 1981, J. Mater. Sci.
Uhlmann, D.R., 1972, J. Non-Cryst. Solids 7, 337. 16, 2337.
Urner-Wille, M., 1981, IEEE Trans. Magn. Waseda, Y., and S. Tamaki, 1976, Z. Phys. B 23,
MAG-17, 3214. 315.
Urner-Wille, M., P. Hansen and K. Witter, 1980, Watanabe, H., H. Morita and H. Yamaguchi,
IEEE Trans. Magn. MAG-16, 1188. 1978, IEEE Trans. Magn. MAG-14, 944.
van Dover, R.B., M. Hong, E.M. Gyorgy, J.F. Weaire, D., M.F. Ashby, J. Logan and M.J. Weins,
Dillon Jr and S.D. Albiston, 1985, J. Appl. 1971, Acta Mater. 19, 779.
Phys. 57, 3897. Weiss, P.R., 1948, Phys. Rev. 74, 1493.
van Dover, R.B., E.M. Gyorgy, R.P. Frankenthal, Weller, D., and W. Reim, 1988, Siemens Forsch.
M. Hong and DJ. Siconolfi, 1986, J. Appl. u. Entwickl. Ber. 17, 111.
Phys. 59, 1291. Weller, D., and W. Reim, 1989a, Mater. Res. Soc.
van Engelen, P.P.J., and K.H.J. Buschow, 1986, Symp. Proc. 150, 33.
J. Magn. & Magn. Mater. 66, 291. Weller, D., and W. Reim, 1989b, Appl. Phys. A
van Engen, P.G., K.H.J. Buschow, R. Jongebreur 49, 599.
and M. Erman, 1983, Appl. Phys. Lett. 42, Weller, D., W. Reim and P. Schrijner, 1988a,
202. IEEE Trans. Magn. MAG-24, 2554.
Velicky, B., S. Kirkpatrick and H. Ehrenreich, Weller, D., W. Reim, H. Ebert, D.D. Johnson and
1968, Phys. Rev. 175, 747. F.J. Pinski, 1988b, J. Physique 49, Suppl., C8-
Vincze, I., D.S. Noudreaux and M. Tegze, 1979, 41.
Phys. Rev. B 19, 4896. Widom, B., 1965, J. Chem. Phys. 43, 3898.
Vittoria, C., P. Lubitz and V. Ritz, 1978, J. Appl. Williams, A., 1981, J. Non-Cryst. Solids 45, 183.
Phys. 49, 4908. Williams, A., 1982, Solid State Commun. 44, 983.
Voloshinskaya, N.M., and G.A. Bolotin, 1974, Williams, A.R., V.L. Moruzzi, C.D. Gelatt Jr, J.
Fiz. Met. & Metalloed. 38, 68. Kfibler and K. Schwarz, 1982, J. Appl. Phys.
Voloshinskaya, N.M., and G.V. Fedorov, 1973, 53, 2019.
Fiz. Met. & Metalloed. 36, 946. Williams, A.R., V.L. Moruzzi, C.D. Gelatt Jr and
yon Heimendahl, L., 1975, J. Phys. F 5, L141. J. Kfibler, 1983a, J. Magn. & Magn. Mater.
yon Molnar, S., R.J. Gambino and J.M.D. Coey, 31-34, 88.
1981, J. Appl. Phys. 52, 2193. Williams, A.R., V.L. Moruzzi, A.P. Malozemoff
Vossen, J.L., and W. Kern, 1978, Thin Film and K. Terakura, 1983b, IEEE Trans. Magn.
Processes (Academic Press, New York). MAG-19, 1983.
452 P. HANSEN

Willich, P., and D. Obertop, 1982, X-Ray Spec- Yamauchi, H., H. Onodera and H. Yamamoto,
trometry 11, 32. 1984, J. Phys. Soc. Jpn. 53, 747.
Wilson, K.G., 1974a, Phys. Rev. B 4, 3174. Yang, X., and T. Miyazaki, 1988, J. Appl. Phys.
Wilson, K.G., 1974b, Phys. Rev. B 4, 3184. 64, 5489.
Woldt, E., and H. Neuh~iuser, 1980, J. Non-Cryst. Yang, X., K. Yamada and T. Miyazaki, 1988, J.
Solids 37, 241. Magn. & Magn. Mater. 71, 172.
Wright, J.G., 1976, IEEE Trans. Magn. MAG- Yang, X.-B., S. Ishio and T. Miyazaki, 1990, IEEE
12, 95. Transl. J. Magn. Jpn. 5, 431.
Xianyu, Z., Y. Ishikawa and S. Onodera, 1982, J. Yasugi, S., S. Honda, M. Ohkoshi and T. Kusuda,
Phys. Soc. Jpn. 51, 1799. 1981, J. Appl. Phys. 52, 2298.
Xiao, G., and C.L. Chien, 1985, J. Magn. & Magn. Yoshino, S., and M. Masuda, 1988, J. Appl. Phys.
Mater. 54-57, 241. 64, 5498.
Xiao, G., and C.L. Chien, 1987, J. Appl. Phys. Yosida, K., 1957, Phys. Rev. 106, 893.
61, 3246. Yu, S.C., J.W. Lynn, J.J. Rhyne and G.E. Fish,
Yagi, M., T. Sato, Y. Sakaki, T. Sawa and K. 1988, J. Appl. Phys. 63, 4083.
Inomata, 1988, J. Appl. Phys. 64, 6050. Zeper, W.B., F.J.A.M. Greidanus and P.F. Carcia,
Yamada, K., Y. Ishikawa, Y. Endoh and T. Masu- 1989a, IEEE Trans. Magn. MAG-25, 3764.
moto, 1975, Solid State Commun. 16, 1335. Zeper, W.B., F.J.A.M. Greidanus, P.F. Carcia and
Yamamoto, M., and I. Yamada, 1988, IEEE C.R. Fincher, 1989b, J. Appl. Phys. 65, 4971.
Transl. J. Magn. Jpn. 3, 683. Ziman, J.M., 1961, Philos. Mag. 6, 1013.

You might also like