7.2 Origin and Consequences of Magnetic Order |
417 |
|
|
Critical exponents for magnetic systems have been defined in the following way. First, we define a dimensionless temperature that is small when we are near the critical temperature.
t = (T −Tc ) / Tc .
We assume B = 0 and define critical exponents by the behavior of physical quantities such as M:
Magnetization (order parameter): M ~ | t |β .
Magnetic susceptibility: |
χ ~ | t |−γ . |
Specific heat: |
C ~ | t |−α . |
There are other critical exponents, such as the one for correlation length (as noted above), but this is all we wish to consider here. Similar critical exponents are defined for other systems, such as fluid systems. When proper analogies are made, if one stays within the same universality class, the critical exponents have the same value. Under rather general conditions, several inequalities have been derived for critical exponents. For example, the Rushbrooke inequality is
α + 2β + γ ≥ 2 .
It has been proposed that this relation also holds as an equality. For mean-field theory α = 0, β = 1/2, and γ = 1. Thus, the Rushbrooke relation is satisfied as an equality. However, except for α being zero, the critical exponents are wrong. For ferromagnets belonging to the most common universality class, experiment, as well as better calculations than mean field, suggest, as we have mentioned (Sect. 7.2.5), β = 1/3, and γ = 4/3. Note that the Rushbrooke equality is still satisfied with α = 0. The most basic problem mean-field theory has is that it just does not properly treat fluctuations nor does it properly treat a related aspect concerning short-range order. It must include these for agreement with experiment. As already indicated, short-range correlation gives a tail on the specific heat above Tc, while the mean-field approximation gives none.
The mean-field approximation also fails as T → 0 as we have discussed. An elementary calculation from the properties of the Brillouin function shows that (s = 1/2)
M = M 0[1− 2 exp(−2Tc / T )] , whereas for typical ferromagnets, experiment agrees better with
M = M 0 (1 − aT 3/ 2 ) .
As we have discussed, this dependence on temperature can be derived from spin wave theory.
Although considerable calculation progress has been made by high-tem- perature series expansions plus Padé Approximants, by scaling, and renormalization group arguments, most of this is beyond the scope of this book. Again,
418 7 Magnetism, Magnons, and Magnetic Resonance
Huang’s excellent text can be consulted.21 Tables 7.2 and 7.3 summarize some of the results.
Table 7.2. Summary of mean-field theory
|
Failures |
Successes |
|
Neglects spin-wave excitations near ab- |
Often used to predict the type of magnetic |
|
solute zero. |
structure to be expected above the lower criti- |
|
Near the critical temperature, it does not |
cal dimension (ferromagnetism, ferrimagnet- |
|
ism, antiferromagnetism, heliomagnetism, |
|
give proper critical exponents if it is be- |
|
etc.). |
|
low the upper critical dimension. |
|
Predicts a phase transition, which certainly will |
|
May predict a phase transition where |
|
occur if above the lower critical dimension. |
|
there is none if below the lower critical |
|
Gives at least a qualitative estimate of the val- |
|
dimension. For example, a one- |
|
dimension isotropic Heisenberg magnet |
ues of thermodynamic quantities, as well as the |
|
would be predicted to order at a finite |
critical exponents – when used appropriately. |
|
temperature, which it does not. |
Serves as the basis for improved calculations. |
|
Predicts no tail in the specific heat for |
|
The higher the spatial dimension, the better it |
|
typical magnets. |
|
is. |
|
|
Table 7.3. Critical exponents (calculated)
|
α |
β |
γ |
|
|
|
|
Mean field |
0 |
0.5 |
1 |
Ising (3D) |
0.11 |
0.32 |
1.24 |
Heisenberg (3D) |
–0.12 |
0.36 |
1.39 |
Adapted with permission from Chaikin PM and Lubensky TC,
Principles of Condensed Matter Physics, Cambridge University
Press, 1995, p. 231.
Two-Dimensional Structures (A)
Lower-dimensional structures are no longer of purely theoretical interest. One way to realize two dimensions is with thin films. Suppose the thin film is of thickness t and suppose the correlation length of the quantity of interest is c. When the thickness is much less than the correlation length (t << c), the film will behave two dimensionally and when t >> c the film will behave as a bulk threedimensional material. If there is a critical point, since c grows without bound as the critical point is approached, a thin film will behave two-dimensionally near the two-dimensional critical point. Another way to have two-dimensional behavior is in layered magnetic materials in which the coupling between magnetic layers, of spacing d, is weak. Then when c << d, all coupling between the layers can
7.2 Origin and Consequences of Magnetic Order |
419 |
|
|
be neglected and one sees 2D behavior, whereas if c >> d, then interlayer coupling can no longer be neglected. This means with magnetic layers, a twodimensional critical point will be modified by 3D behavior near the critical temperature.
In this chapter we are mainly concerned with materials for which the threedimensional isotropic systems are a fairly good or at least qualitative model. However, it is interesting that two-dimensional isotropic Heisenberg systems can be shown to have no spontaneous (sublattice − for antiferromagnets) magnetization [7.49]. On the other hand, it can be shown [7.26] that the highly anisotropic
two-dimensional Ising ferromagnet (defined by the Hamiltonian H ∑i,j(nn.)σizσjz, where the σs refer to Pauli spin matrices, the i and j refer to lattice sites) must show spontaneous magnetization.
We have just mentioned the two-dimensional Heisenberg model in connection with the Mermin–Wagner theorem. The planar Heisenberg model is in some ways even more interesting. It serves as a model for superfluid helium films and predicts the long-range order is destroyed by formation of vortices [7.40].
Another common way to produce two-dimensional behavior is in an electronic inversion layer in a semiconductor. This is important in semiconductor devices.
Spontaneously Broken Symmetry (A)
A Heisenberg Hamiltonian is invariant under rotations, so the ensemble average of the magnetization is zero. For every M there is a −M of the same energy. Physically this answer is not correct since magnets do magnetize. The symmetry is spontaneously broken when the ground state does not have the same symmetry as the Hamiltonian, The symmetry is recovered by having degenerate ground states whose totality recovers the rotational symmetry. Once the magnet magnetizes, however, it does not go to another degenerate state because all the magnets would have to rotate spontaneously by the same amount. The probability for this to happen is negligible for a realistic system. Quantum mechanically in the infinite limit, each ground state generates a separate Hilbert space and transitions between them are forbidden—a super selection rule. Because of the symmetry there are excited states that are wave-like in the sense that the local ground state changes slowly over space (as in a wave). These are the Goldstone excitations and they are orthogonal to any ground state. Actually each of the (infinite) number of ground states is orthogonal to each other: The concept of spontaneously broken symmetry is much more general than just for magnets. For ferromagnets the rotational symmetry is broken and spin waves or magnons appear. Other examples include crystals (translation symmetry is broken and phonons appear), and superconductors (local gauge symmetry is broken and a Higgs mode appears—this is related to the Meissner effect – see Chap. 8).22
420 7 Magnetism, Magnons, and Magnetic Resonance
7.3 Magnetic Domains and Magnetic Materials (B)
7.3.1 Origin of Domains and General Comments23 (B)
Because of their great practical importance, a short discussion of domains is merited even though we are primarily interested in what happens in a single domain.
We want to address the following questions: What are the domains? Why do they form? Why are they important? What are domain walls? How can we analyze the structure of domains, and domain walls? Is there more than one kind of domain wall?
Magnetic domains are small regions in which the atomic magnetic moments are lined up. For a given temperature, the magnetization is saturated in a single domain, but ferromagnets are normally divided into regions with different domains magnetized in different directions.
When a ferromagnet splits into domains, it does so in order to lower its free energy. However, the free energy and the internal energy differ by TS and if T is well below the Curie temperature, TS is small since also the entropy S is small because the order is high. Here we will neglect the difference between the internal energy and the free energy. There are several contributions to the internal energy that we will discuss presently.
Magnetic domains can explain why the overall magnetization can vanish even if we are well below the Curie temperature Tc. In a single domain the M vs. T curve looks somewhat like Fig. 7.16.
M
MS
H = 0
Tc
T
Fig. 7.16. M vs. T curve for a single magnetic domain
For reference, the Curie temperature of iron is 1043 K and its saturation magnetization MS is 1707 G. But when there are several domains, they can point in different directions so the overall magnetization can attain any value from zero up to saturation magnetization. In a magnetic field, the domains can change in size (with those that are energetically preferred growing). Thus the phenomena of hysteresis, which we sketch in Fig. 7.17 starting from the ideal demagnetized state, can be understood (see Section Hysteresis, Remanence, and Coercive Force).
23 More details can be found in Morrish [68] and Chikazumi [7.11].
7.3 Magnetic Domains and Magnetic Materials (B) 421
M
T << Tc
Ms = saturation magnetization
H
Fig. 7.17. M vs. H curve showing magnetic hysteresis
Wall
M
M
domains 
Fig. 7.18. Two magnetic regions (domains) separated by a domain wall, where size is exaggerated
In order for some domains to grow at the expense of others, the domain walls separating the two regions must move. Domain walls are transition regions that separate adjacent regions magnetized in different directions. The idea is shown in Fig. 7.18.
We now want to analyze the four types of energy involved in domain formation. We consider (1) exchange energy, (2) magnetostatic energy, (3) anisotropy energy, and (4) magnetostrictive energy. Domain structures with the lower sum of these energies are the most stable.
Exchange Energy (B)
We have seen (see Section The Heisenberg Hamiltonian and its Relationship to the Weiss Mean-Field Theory) that quantum mechanics indicates that there may be an interaction energy between atomic spins Si that is proportional to the scalar product of the spins. From this, one obtains the Heisenberg Hamiltonian describing the interaction energy. Assuming J is the proportionality constant (called the exchange integral) and that only nearest-neighbor (nn) interactions need be considered, the Heisenberg Hamiltonian becomes
H = −J ∑i, j Si S j , |
(7.232) |
(nn) |
|
422 7 Magnetism, Magnons, and Magnetic Resonance
where the spin Si for atom i when averaged over many neighboring spins gives us the local magnetization. We now make a classical continuum approximation. For the interaction energy of two spins we write:
Uij = −2JSi S j . |
(7.233) |
Assuming ui is a unit vector in the direction of Si we have since Si = Sui:
U |
ij |
= −2JS 2u u |
j |
. |
(7.234) |
|
i |
|
|
If rji is the vector connecting spins i and j, then
u |
j |
= u + r |
ji |
( u) |
i |
, |
(7.235) |
|
i |
|
|
|
treating u as a continuous function r, u = u(r). Then since
(u |
j |
− u )2 |
= u2 |
+ u2 |
− 2u u |
j |
= 2(1− u u |
j |
) , |
(7.236) |
|
i |
j |
i |
i |
i |
|
|
we have, neglecting an additive constant that is independent of the directions of ui and uj,
Uij = +JS 2 (u j − ui )2 .
So |
|
|
|
|
|
|
|
|
|
|
|
|
|
Uij |
= +JS 2 (rji |
u)2 . |
|
(7.237) |
Thus the total interaction energy is |
|
|
|
|
|
|
|
|
|
U = |
1 |
∑U |
ij |
= |
JS 2 |
∑ |
i, j |
(r |
ji |
u)2 |
, |
(7.238) |
2 |
|
|
2 |
|
|
|
|
|
where we have inserted a 1/2 so as not to count bonds twice. If u =α1i +α2 j +α3k ,
where the αi are the direction cosines, for rji = ai, for example:
∑(rji u)2 = 2a2 |
∂α |
|
|
∂α |
2 |
j + |
∂α |
3 |
|
2 |
|
|
|
|
|
1 i + |
|
|
|
k |
|
|
|
|
±ai |
|
|
∂x |
|
|
∂x |
|
|
∂x |
|
|
|
|
|
(7.239) |
|
|
|
∂α1 |
|
2 |
∂α2 |
|
2 |
|
∂α3 |
|
2 |
|
|
|
|
= 2a |
2 |
|
|
+ |
|
|
. |
|
|
∂x |
|
|
+ |
∂x |
|
|
∂x |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
For a simple cubic lattice where we must also include neighbors at rji = ±aj and
±ak, we have:24
U = |
JS 2 |
∑ [( α1)2 + ( α2 )2 + ( α3)2 ]i a3 , |
(7.240) |
|
a |
i (all spins) |
|
24An alternative derivation is based on writing U ∑μiBi, where μi is the magnetic moment Si and Bi is the effective exchange field ∑j(nn) JijSj, treating the Sj in a continuum spatial approximation and expanding Sj in a Taylor series (Sj = Si + a∂Si /∂x + etc. to 2nd order). See (7.275) and following.
7.3 Magnetic Domains and Magnetic Materials (B) 423
or in the continuum approximation:
|
U = |
JS 2 |
∫[( α1)2 + ( α2 )2 + ( α3)2 ]dV . |
(7.241) |
|
a |
|
|
|
|
For variation of M only in the y direction, and using spherical coordinates r, θ, φ, a little algebra shows that (M = M(r, θ, φ))
Energy |
|
∂θ |
2 |
|
2 |
|
∂ϕ |
2 |
|
|
|
|
|
+ sin |
|
|
, |
(7.242) |
Volume |
= A |
|
|
|
θ |
|
|
|
|
∂y |
|
|
|
|
∂y |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where A = JS2/a and has the following values for other cubic structures (Afcc = 4A, and Abcc = 2A). We have treated the exchange energy first because it is this inter-
action that causes the material to magnetize.
Magnetostatic Energy (B)
We have already discussed magnetostatics in Sect. 7.2.2. Here we want to mention that along with the exchange interaction it is one of the two primary interactions of interest in magnetism. It is the driving mechanism for the formation of domains. Also, at very long wavelengths, as we have mentioned, it can be the causative factor in spin-wave motion (magnetostatic spin waves). A review of magnetostatic fields of relevance for applications is given by Bertram [7.6].
Anisotropy (B)
Because of various energy-coupling mechanisms, certain magnetic directions are favored over others. As discussed in Sect. 7.2.2, the physical origin of crystalline anisotropy is a rather complicated subject. As discussed there, a partial understanding, in some materials, relates it to spin-orbit coupling in which the orbital motion is coupled to the lattice. Anisotropy can also be caused by the shape of the sample or the stress it is subjected to, but these two types are not called crystalline anisotropy. Regardless of the physical origin, a ferromagnetic material will have preferred (least energy) directions of magnetization. For uniaxial symmetry, we can write
Hanis = −Da ∑i (k Si )2 , |
(7.243) |
where k is the unit vector along the axis of symmetry. If we let K1 = DaS2/a3, where a is the atom–atom spacing, then since sin2θ = 1 − cos2θ and neglecting unimportant additive terms, the anisotropy energy per unit volume is
u |
anis |
= K sin2 |
θ . |
(7.244) |
|
1 |
|
|
424 7 Magnetism, Magnons, and Magnetic Resonance
Also, for proper choice of K1, this may describe hexagonal crystals, e.g. cobalt (hcp) where θ is the angle between M and the hexagonal axis. Figure 7.19 shows some data related to anisotropy. Note Fe with a bcc structure has easy directions in 100 and Ni with fcc has easy directions in 111 .
Magnetization, 4πM (gauss)
25000
20000
15000 

10000 |
|
|
Observed data |
|
|
|
|
[100] |
|
|
|
|
[110] |
|
5000 |
|
|
[111] |
|
|
|
|
|
0 |
|
|
|
|
0 |
100 |
200 |
300 |
400 |
H (Oe)
Fig. 7.19. Magnetization curves showing anisotropy for single crystals of iron with 3.85% silicon [Reprinted with permission from Williams HJ, Phys Rev 52, 1 (1937). Copyright 1937 by the American Physical Society.]
Wall Energy (B)
The wall energy is an additive combination of exchange and anisotropy energy, which are independent. Exchange favors parallel moments and a wide wall. Anisotropy prefers moments along an easy direction and a narrow wall. Minimizing the sum of the two determines the width of the wall. Consider a uniaxial ferromagnet with the magnetization varying only in the y direction. If the energy per unit volume is (using spherical coordinates, see, e.g., (7.242) and Fig. 7.25)
|
∂θ |
2 |
|
∂ϕ |
|
2 |
|
|
|
|
|
|
|
|
|
|
+ K sin2 |
θ , |
(7.245) |
w = A |
|
|
|
|
+ sinθ |
|
|
|
|
|
|
|
|
|
∂y |
|
|
|
|
1 |
|
|
|
|
∂y |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
JS |
2 |
|
|
|
|
|
|
D S 2 |
|
|
|
A = α |
|
|
|
and |
K = κ |
|
a |
, |
|
(7.246) |
|
|
|
|
a3 |
|
|
1 a |
|
|
1 |
|
|
|
1 |
|
|
|
and α1, κ1 differ for different crystal structures, but both are approximately unity. For simplicity in what follows we will set α1 and κ1 equal to one.
7.3 Magnetic Domains and Magnetic Materials (B) 425
Using δ∫wdy = 0 we get two Euler–Lagrange equations. Inserting (7.245) in the Euler–Lagrange equations, we get the results indicated by the arrows.
∂w |
− |
d ∂w |
|
|
= 0 → |
d |
|
|
K |
sin2 |
θ = 2A |
d |
∂θ |
, |
|
|
dy |
|
|
∂θ |
|
|
|
|
|
|
∂θ |
|
∂ |
|
|
|
|
|
|
dθ 1 |
|
|
dy ∂y |
|
|
|
|
|
|
|
∂y |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
∂w |
|
|
d |
|
∂w |
|
|
|
|
|
|
|
d |
|
|
∂ϕ |
|
|
|
|
|
|
− |
|
|
|
|
|
|
|
|
|
|
= 0 |
→ |
|
2 |
|
|
sin2 θ |
|
= 0 . |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
∂ϕ |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
∂ϕ |
|
dy |
|
∂ |
|
|
|
|
|
|
|
dy |
|
∂y |
|
|
|
|
|
|
|
|
|
|
|
|
∂y |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
For Bloch walls by definition, φ = 0, which is a possible solution. The first equation (7.247) has a first integral of
A dθ |
= sinθ , |
|
(7.249) |
K dy |
|
|
|
|
1 |
|
|
|
|
which integrates in turn to |
|
|
|
|
θ = 2 arctan(e y |
0 ) , |
0 = |
A . |
(7.250) |
|
|
|
K1 |
|
The effective wall width is obtained by approximating dθ/dy by its value at the midpoint of the wall, where θ = π/2.
dθ |
= |
K1 |
|
1 |
D |
, |
(7.251) |
dy |
|
A |
|
a |
J |
|
|
so the wall width/a is |
|
|
|
|
|
|
|
wall width |
= π |
D . |
|
|
a |
|
|
|
J |
|
|
One can also show the wall width per unit area (perpendicular to the y-axis in Fig. 7.25) is 4(AK1)1/2. For Iron, the wall energy per unit area is of order 1
erg/cm2, and the wall width is of order 500 Å.
Magnetostrictive Energy (B)
Magnetostriction is the variation of size of a magnetic material when its magnetization varies. Magnetostriction implies a coupling between elastic and magnetic effects caused by the interaction of atomic magnetic moments and the lattice. The magnetostrictive coefficient λ is δl/l, where δl is the change in length associated with the magnetization change. In general λ can be either sign and is typically of the order of 10−5 or so. There may also be a change in volume due to changing magnetization. In any case the deformation is caused by a lowering of the energy.
426 7 Magnetism, Magnons, and Magnetic Resonance
Magnetostriction is a very complex matter and a detailed description is really outside the scope of this book. We needed to mention it because it has a bearing on domains. See, e.g., Gibbs [7.24].
Formation of Magnetic Domains (B)
We now give a qualitative account of the formation of domains. Consider a cubic material, originally magnetized along an easy direction as shown in Fig. 7.20. Because the magnetization M and demagnetizing fields have opposite directions (7.136), this configuration has large magnetostatic energy. The magnetostatic energy can be reduced if the material splits into domains as shown in Fig. 7.21
Fig. 7.20. Magnetic domain formation within a material
Since the density of surface poles is +M · an interface the net magnetic charge per unit
(M2 − M1)
n where nM is the outward normal, at area is
nM 2 ,
where nM2 is a unit vector pointing from region 1 to region 2. Thus when M · n is continuous, there are no demagnetizing fields (assuming also M is uniform in the interior). Thus (for typical magnetic materials with cubic symmetry) the magnetostatic energy can be further reduced by forming domains of closure, as shown in Fig. 7.22. The overall magnetostrictive and strain energy can be reduced by the formation of more domains of closure (see Fig. 7.23). That is, this splitting into smaller domains reduces the extra energy caused by the internal strain brought about by the spontaneous strain in the direction of magnetization. This process will not continue forever because of the increase in the wall energy (due to exchange and anisotropy). An actual material will of course have many imperfections as well as other complications that will cause irregularities in the domain structure.