You are on page 1of 69

chapter 2

PERMANENT MAGNETS; THEORY

H. ZlJLSTRA
Philips Research Laboratories
Eindhoven
The Netherlands

Ferromagnetic Materials, Vol. 3


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

37
CONTENTS

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

38
1. Introduction

1.1. General properties and applications

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


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

39
40 H. ZIJLSTRA

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

Function Application

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

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

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


compasses, some ammeters

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


cathode-ray tubes, Hall plates

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


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

1.2. The hysteresis loop

Permanent magnet materials are characterized by high coercivities and high


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

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

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

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

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

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

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

2. Suitability criteria for applications

2.1. The energy product

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

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

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


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

curl H = O,

and for the flux density B we always have

div B = 0.

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


1962a)

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

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

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

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

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

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

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

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

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

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

50 j zt ~

20

10
//
/
5 /
/
//

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

1
Ig

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

and, since no currents are present,

Hmlm - G i g = O,

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

H m = Hglg/Im,

and
B m = - tzoHgSg/ Sm

B m / H m = - tZoSglm / Smlg .

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

HgLg = otHmLm,

BgSg = - /3BmSm .

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

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

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


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

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


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

TdS + dA = dU.

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

F=U-TS.

It follows from these two relations that

dF = dA- S dT.

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

dF = dA.

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

dA ½f_ (H. dS - S . d H ) d V,
magn

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

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

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

dFsystem = E dVp+½E dFi,

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

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

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


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

B/
P Br

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

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

½B~BH~ = (1/2/Xo)J 2 •

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


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

3. Magnetic anisotropy

3.1. Anisotropy field and coercivity associated with magnetic anisotropy

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


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

Wk = K sin 2 q~,

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

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

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

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

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


50 H. ZIJLSTRA

easy
axis

Ms

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

H > -2K/IxoMs.

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


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

Hc = 0.48HA

if coherent rotation is assumed.


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

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

0 0,10,90
1

M/M s

T
¢' 0

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

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

1.0

f Y
Sf
0.5 M/Ms

-0.5

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

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

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

3.2. Shape anisotropy

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

I
I
!
I
I
I

Fig. 9. Magnetized prolate spheroid.


PERMANENT MAGNETS; THEORY 53

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

I.l"ondi = -- NiJi ,

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

f (H.B) dV=O,

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

Em = II.~OIR n 2 d V ,

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

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

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

HA = (N±- Nfl/~0 •

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

3.3. Magnetocrystalline anisotropy

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

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

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

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


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

4. Fine particles

4.1. Critical radius for single-domain particles

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


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

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

HA jnc

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

Ticonal 900 370 4.6 100 1.3


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

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


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

We = A(V~p) 2 .

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


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

W r = K sin 2 ~ ,

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


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

/ I.l,o A \ 1/2
Rco = 5.099~-~s2 ) •

If we assume that for iron (Kittel 1949)

A ~ 2 x 10 -11Jm -1 ,

we find that

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

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

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

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


this condition is defined by

K<O.1781J~/~o
or

K < 1.0686Wm, where

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

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


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

R > RCl ,

where

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

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

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


this condition is defined by

K > 0.1781J~/~0,

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

R > Rc2,
where

/x0A ~/: /x0K 1/2


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

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

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


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

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


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

102

two doma ~s

10

curling

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

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

#0A 1/2 /z0K 1/2


R<9tx°2Y
Js =36(js)~ (Js)

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

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

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

4.2. Brown's paradox

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

H = HA + Hw+ Ha+ He,

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

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

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

- H e > HA

the magnetization reverses by uniform rotation.


The coercivity of a spherical crystal is thus always

H~=Ha,

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

5. Coercivity associated with shape anisotropy

5.1. Prolate spheroid

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

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

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


/Xo

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

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

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


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

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

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

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

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

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

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

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

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


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

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

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

Uniform r o t a t i o n

Buckling or
twisting

0.2 HC/HA

o, l Curling

0.05

0.02

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

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

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


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

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

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

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

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

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


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

5.2. Chain of spheres

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


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

a b c

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

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

b
0

I I I

~- Elongation

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

The magnetization reversal in elongated particles has been thoroughly analyzed


theoretically by Aharoni (1966).

6. Coercivity associated with magnetocrystalline anisotropy

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

The fundamental requirements for magnetization reversal by domain wall pro-


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

Hn ~H

IHnl>lHpl

b
J
Hp Hn ~H

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

,~ oo-o.--=

B C ':"

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

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

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


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

6.2. The 180" domain wall

6.2.1. Energy and width of a 180 ° domain wall


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

WK = K sin 2 q~, (6.1)

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


68 H. Z I J L S T R A

easy
axis
I
I

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

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

We=A/d ] 2 (6.2)

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


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

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

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


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

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

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

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

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

tan ½~o= e ~'/~ ,


PERMANENT MAGNETS; THEORY 69

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

Y
0
6 Ii-

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

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


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

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

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

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

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

6.2.2. The exchange energy coefficient A

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

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

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

which for A~ ~ 1 approaches

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

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


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

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

This expression will be applied to various crystal lattices:


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

13

Fig. 21. Body-centred cubic crystal lattice.

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

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

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

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

Cl

Fig. 22. Face-centred cubic crystal lattice.

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

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

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

9i
i
i

i i

i i PF

O-- • - :g-'- - -O


Q,,
6

Fig. 23. Simple cubic crystal lattice.

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


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

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

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

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

0 / ~\
¢_~J \
~X

Fig. 24. Hexagonal close-packed crystal lattice.

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

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

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


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

NM2.2{d~ \2
we = .

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

TABLE 3
Interactions in an hcp cell.

Location in cell Number cos2A p Contribution

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

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

Total contribution 4
PERMANENT MAGNETS; THEORY 73

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


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

Ew = ½1~oNM2

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

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

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


sc: A = la2Ew = ~2Ew,

hcp: A = {a2Ew = {sC2Ew.

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


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

w~j = -2J$~ • ~ ,

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

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

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

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

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

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


74 H. ZIJLSTRA

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

A~105~ZTc, (SI),

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

A ~ 106~2Tc, (cgs).

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

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


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

radius p

0
0 © ©
0 )
d

0 )
t
Domain Wall

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

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

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

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

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


\4~r/ p2 "

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

/4,B-'~ 1/3

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


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

F = r~
.~2/3g
Jm •

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


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

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

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

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

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

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

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

3 )2/3 ~-y
He=
_

4~r
_ _ _

pJs 'x
^ 2/3

" (6.8)

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

1 2
m = g/toMs V,

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

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


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

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

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

AW m 9171"[.ZoIVl
. # 2s P 3•

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

f = d(Wm+ W~)
dx

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

maximum value of its derivative with respect to x,

d Wm') = 0.600 j2p2,


dx ,/max /~0

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

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


• \ " /xopMs]"

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

pe ~ ~o~/Y~ •

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

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

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

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


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

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

6.3.2. Nucleation of reverse domains at large inclusions


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

Hc = 1.23y/tzopMs. (6.10)

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

6.3.3. Domain-wall pinning at small inclusions


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

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


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

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


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

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


calculated by N6el (1944b) to be

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

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

1 2
Wd = gtx0Ms V.

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

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

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

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

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


second term and find

A W = -2KV sin 2 q~.

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

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

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

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

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

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

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


80 H. ZIJLSTRA

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

6.4. Domain-wall nucleation at surface defects

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


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

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

Fig. 27. Surface defect.


PERMANENT MAGNETS; THEORY 81

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

Ha ~ 1.8Ms.

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

6.5. Interaction of domain walls with the crystal lattice

6.5.1. Wall pinning at regions with deviating K and A


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

r~ e..

"" O
E

,- ©

o~
~E

r.~
=2
o

r~

O
PERMANENT MAGNETS; THEORY 83

/
a

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

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


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

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

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


where

Ja=Jc¢Jb,
A a = A~ ¢ A b ,

g . = K ~ gu.
84 H. ZIJLSTRA

I x

I ,/

c ,¸

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

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

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

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

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

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


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

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


Ax

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


Ax
PERMANENT MAGNETS; THEORY 85

i
i

[ i
I I
i 1
I I
I I

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

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

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

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

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


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

2HAaJa - O. (6.14)
AaKa - AbKb

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

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

cos %

J
cos "P2
-p

\
\\

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

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


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

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


cos go - AbJ (6.16)
dX 1

with

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


Cb - AbKb Ab

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


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

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

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

0 . . . . . I
XI X2 ~ X

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

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

(Abl Kb) 112


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

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


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

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


/?A ~ +T "

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

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

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

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

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

(6.19)
88 H. ZIJLSTRA

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

(KaAa - KbAb) sin 2 q?2


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

which in the small-deviations approximation becomes

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


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

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

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

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

6.5.2. Pinning of a domain wall by an antiphase boundary

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

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

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

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


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

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

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

0 -×

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

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

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

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

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

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

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

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

and

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


PERMANENT MAGNETS; THEORY 91

-5

I
1'0

"-5

--10

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

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


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

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


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

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

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


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

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

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


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

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

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

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


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

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

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

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

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

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

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

0.5

\
0.2

0.1
\ \

0.05
\ \

\
\
0.02

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

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


PERMANENT MAGNETS; THEORY 93

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

6.5.3. Nucleation of a domain wall at an antiphase boundary


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

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

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

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

q~

q01 -~ x r

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

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

cos ¢~ = 1 for 0 > 7 > -1


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

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

he = cos q~l •

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

6.5.4. Thin-wall coercivity in a perfect crystal


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

0.5

02 \

ho
l 0.1
"..
0.05

0.02

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

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


PERMANENT MAGNETS; THEORY 95

9_ b

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

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

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

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


i=1 a

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

I I I I I

i i 1 x\
I t

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

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

10-1

aK
10-2

10.3

10-~

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

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

TC
3
&,.p

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

f
0.5

hc

0.2

O.OE

0.02

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

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

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

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

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

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


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

6.5.5. Partial wall pinning at discrete sites


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

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

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


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

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

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


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

AV=/3x (per unit S),

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

AE = yAS - 2HJ A V

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

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

K > 210 J2 ,
100 H. ZIJLSTRA

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

x = HYa2/y, (6.29)

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

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

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

n c o( /,/1/3

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

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

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


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

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

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

AE = Kv.

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

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

10 101° 0.8 x 10 -19


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

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


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

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

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

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

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

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

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

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

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


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

w ~6x 10 ~°s -~,


or

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

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

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

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

A E = 23/6 2 ~ 20A6,

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


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

p= 10 -207S 1,

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


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

\,
UFeU
S

pinning site

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

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


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

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

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


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

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

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

A E = a3"oS,

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


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

S = 10-27 m 2 ,

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


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

References

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

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

You might also like