Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Fundamentals of the Physics of Solids / 14-Magnetically Ordered Systems

.pdf
Скачиваний:
23
Добавлен:
08.01.2014
Размер:
943.46 Кб
Скачать

14.3

Simple Models of Magnetism

469

eg

 

 

eg

 

 

 

 

 

 

 

 

 

 

 

 

 

t2g

 

 

 

 

 

 

 

 

 

t2g

 

3+

 

 

 

 

 

S 2

 

4+

 

 

 

S 3/2

Mn

 

 

 

 

 

 

 

Mn

 

 

 

 

 

Fig. 14.10. The d electrons of Mn3+ and Mn4+ ions on the t2g and eg levels

takes place with high probability only when the spins of neighboring ions are aligned ferromagnetically. Since hopping lowers the total energy by reducing the kinetic energy, ferromagnetic alignment of the spins is energetically favored. Thus, in contrast to superexchange, which is often antiferromagnetic, double exchange is always ferromagnetic. The propagation (hopping) of the electrons in the crystal indicates that such systems are metallic.

Such a double exchange plays an important role in magnetite, which contains divalent (Fe2+, 3d6) and trivalent (Fe3+, 3d5) iron ions. The spins of the Fe2+ and Fe3+ ions at octahedral sites are aligned by double exchange interactions. On the other hand, the spins of the Fe3+ ions at tetrahedral sites are antiparallel to the spins at octahedral sites because of antiferromagnetic superexchange interactions.

14.3 Simple Models of Magnetism

In the previous section we saw that, up to an additive constant, the interaction between two localized spins S1 and S2 can always be written as

H = 2JS1 · S2 ,

(14.3.1)

independently of the specific mechanism of the exchange – although the coupling strength J depends on the character of the exchange and on the separation of atomic spins. We now have to generalize this formula to the case when atomic spins are arranged in a regular crystalline structure.

14.3.1 The Isotropic Heisenberg Model

It is not easy to generalize the exchange interaction between two spins to a solid that contains a large number of magnetic atoms. For simplicity we shall assume that a net spin Si is located at each lattice point Ri of the crystal. It is known that the atomic magnetic moment can also come from orbital motion; indeed, one should therefore deal with the total angular momentum J i and the corresponding magnetic moment μi = −gJ μBJ i. However, the microscopic origin of the moments is of little importance for the phenomena

470 14 Magnetically Ordered Systems

discussed here, therefore we shall continue to speak about spins and use the notation Si – although by this we shall invariably mean the total localized angular momentum of the ion. The only notational hint for this will be using g instead of ge or gJ for the g-factor.

The net spin Si can be the resultant of several electronic spins, and electrons may be exchanged with the electrons of several neighboring atoms. The problem is greatly simplified by the assumption that the total interaction can be written as the sum of pairwise interactions. If it is also assumed that for each atom it is su cient to consider only the ground-state configuration as determined by Hund’s rules (or by the splitting due to the crystal field), and higher-lying excited states can be neglected, then using some simple considerations the results obtained for two electrons are fairly straightforwardly generalized without a full quantum mechanical treatment. In the subspace of the Hilbert space that contains the lowest-lying states the magnitude of each atomic spin (and therefore of each atomic magnetic moment) is constant. Interactions may at most change their orientations. If the interaction between the atomic spins at lattice points Ri and Rj is given by a term of the form of the Heisenberg exchange interaction, Si · Sj , and the strength of the interaction can be characterized by the exchange coupling – i.e., the amplitude Jij = J(Ri − Rj ), which depends on the separation of the spins –, then the

system of spins is described by the e ective Hamiltonian

 

 

 

 

 

 

 

H =

 

 

(14.3.2)

 

Jij Si · Sj .

 

 

 

ij

 

 

 

 

 

 

 

Since the indices i and j both run over the entire lattice, each pair of spins occurs with the correct weight factor. It was first pointed out by W. Heisenberg (1928), and independently of him by J. Frenkel that ferromagnetism can be explained in terms of such pairwise interactions between localized spins. For this reason the model described by the above Hamiltonian is called the Heisenberg model of magnetism. According to the foregoing, states in which electrons are transferred from one atom to the other (for no matter how short a period) are not present explicitly in this model. Only the orientation of atomic moments is considered.

At low temperatures the exchange interaction does not usually leave atomic spins independent: they tend to align each other into a preferred direction, giving rise to an ordered structure. The simplest of them is ferromagnetic order.This is observed in materials in which the sign of the exchange interaction is positive, hence it lines up neighboring spins in the same direction. However, the exchange interaction is not necessarily of ferromagnetic character. As it was discussed at the beginning of this chapter, ferromagnets form a relatively small group of magnetically ordered materials. More numerous are antiferromagnets, in which the dominant nearest-neighbor coupling is antiferromagnetic (Jij < 0), which gives rise to an ordered array of alternately

14.3 Simple Models of Magnetism

471

oriented spins – the details of which depend on the coupling between more distant neighbors.

In magnetic materials described with the isotropic Heisenberg model the interaction fixes only the relative orientation of spins; it does not determine their exact spatial orientation, as the Hamiltonian itself possesses full spherical symmetry (O(3) or SU(2) symmetry). The ordered ferromagnetic or antiferromagnetic state does not show this continuous spherical symmetry. The continuous symmetry of the Hamiltonian is therefore spontaneously broken in the actual state. This has important consequences on the low-energy excitation spectrum of the system. Namely, according to Goldstone’s theorem presented in Chapter 6, when the ground state of the system breaks a continuous symmetry of the Hamiltonian, there must be a gapless bosonic branch in the excitation spectrum.

Turning on a symmetry-breaking magnetic field facilitates the mathematical determination of the symmetry-breaking solutions of the Hamiltonian. Since magnetization will point in this direction, it is practical to choose this direction as the spin quantization axis. At the end of the calculation we may take the limit of vanishing magnetic field. In this approach the Hamiltonian is customarily written as

H = i,j

Jij Si · Sj − gμBB ·

i

Si ,

(14.3.3)

 

 

 

 

 

where g is the g-factor of the localized moment. If the moment comes entirely from the spin then g is the g-factor of the electron, which is negative. If the total atomic moment has a part that comes from the orbital angular momentum then the Landé factor gJ must be used, however comparison with (3.2.70) shows that g = −gJ .

The field B can be replaced by μ0H. If we were to use the full magnetic induction μrμ0H for B then the e ects of other moments would be counted twice in magnetically ordered materials as the exchange part of the Heisenberg model contains precisely the e ects of the neighbors.

14.3.2 Anisotropic Models

We have seen that the e ective Hamiltonian of the exchange interaction between two spins is spherically symmetric, therefore it is invariant under arbitrary rotations of the two spins through a common angle. For spins in a crystal lattice the interaction is expected to depend on the orientation of the spins with respect to the crystalline axes, since the symmetries of the crystal also have to appear in the form of spin–spin interactions. When only terms bilinear in the spin are allowed, the magnetic behavior for systems with cubic symmetry remains to be described by the Hamiltonian

H = − Jij Si · Sj , (14.3.4)

i,j

472 14 Magnetically Ordered Systems

since the only bilinear expression in the spin components that is invariant under the symmetries of a cubic crystal is the isotropic combination

 

Jij SixSjx + SiySjy + Siz Sjz

!

 

H =

.

(14.3.5)

i,j

In tetragonal crystals one axis has a privileged status over the two others. This uniaxial anisotropy has to be reflected in the magnetic Hamiltonian, too. Choosing the z-axis along this direction, the Hamiltonian that possesses the symmetries of the crystal is

 

Jij SixSjx + SiySjy + ΔSiz Sjz

!

 

H =

.

(14.3.6)

i,j

When > 1, the z components of the spins are coupled most strongly, so a nonvanishing magnetization appears preferentially along this direction. In this case the magnetic material has an easy axis of magnetization.When < 1, the spin components in the (x, y) plane are coupled strongly, and so magnetization appears in this plane – the easy plane of magnetization. In an extreme case of uniaxial anisotropy only the z components of the spins are coupled:

 

 

HIsing = − Jij Siz Sjz .

(14.3.7)

i,j

This is the Ising model,6 a fundamental model of statistical physics, since the partition function can be calculated as a sum over all possible classical configurations of the spin projections. On the other hand, when = 0, the XY model is recovered:

 

Jij SixSjx + SiySjy

!

 

HXY =

.

(14.3.8)

i,j

In tetragonal crystals, in addition to the contribution of ion pairs to uniaxial anisotropy, a term that corresponds to single-ion anisotropy,

 

 

−D (Siz )2

(14.3.9)

i

may also appear in the Hamiltonian. In cubic crystals the first anisotropic contribution contains the fourth-order product of spins. The single-ion anisotropy term is

Haniso = K

i

(Six)4 + (Siy )4 + (Siz )4 .

(14.3.10)

 

 

 

 

We shall choose the z-axis as the quantization direction, and instead of the operators Sx and Sy we shall repeatedly use the spin-projection-lowering and -raising operators

6E. Ising, 1925. It would be more appropriate to call it the Lenz–Ising model, since it was proposed by Ising’s supervisor, W. Lenz, in 1920.

14.4

The Mean-Field Approximation 473

Si+ = Six + iSiy ,

Si= Six iSiy .

(14.3.11)

In terms of these the Hamiltonian of the isotropic and uniaxially anisotropic systems are

H =

 

 

21

 

Si+Sj+ SiSj+

+ Siz Sjz

!

(14.3.12)

Jij

 

 

 

 

 

 

 

i,j

 

 

 

 

 

 

 

 

 

 

and

=

 

1

 

 

+

+ S

S+

+ ΔSz Sz

!

(14.3.13)

 

 

H

Jij

 

S

.

 

2

 

i Sj

 

i j

i j

 

i,j

These forms show even more clearly that the Heisenberg model – unlike the Ising model – is truly quantum mechanical. And although the classical model is more easily treated, we shall deal with the quantum mechanical model, as it corresponds more closely to physical reality.

14.4 The Mean-Field Approximation

In the introductory section magnetically ordered states were characterized by the relative orientation of a classical vector, the expectation value of the spin or magnetic moment. At low temperatures the orientation of spins is such that their projection along the preferred direction is maximum with high probability. As temperature is increased, the expectation value of the magnetic moment decreases and finally vanishes. The simplest description of this phenomenon can be given in the framework of the mean-field theory. This approximation is based on the assumption that the alignment of individual spins is not sensitive to the thermal or quantum fluctuations of neighboring spins. Assuming that neighboring moments can be replaced by their average, they create an e ective static magnetic field, and the magnetic moment of the atom lines up with this. On the other hand, when the g-factor in μi = BSi (i.e., the negative of the Landé factor) is negative, spins line up in the opposite direction.

To determine the field generated by the neighbors, we shall write the spin operator Si in the equivalent form

Si = Si + (Si − Si ) ,

(14.4.1)

where denotes the thermal average. Substituting this in the Hamiltonian,

H = i,j

Jij (Si − Si + Si ) · (Sj − Sj + Sj ) − gμ0μBH ·

i

Si

 

 

Jij Sj Si − gμ0μBH ·

 

Si

 

 

= i,j

Jij Si Sj 2 i,j

i

 

 

 

 

 

 

 

 

 

 

− Jij (Si − Si ) · (Sj − Sj ) .

 

 

(14.4.2)

i,j

474 14 Magnetically Ordered Systems

The mean-field approximation consists of neglecting the last term that is of second order in the fluctuations, i.e., in the deviations Si − Si from the mean value. The Hamiltonian obtained in this way is linear in the spin operators, so it can be rewritten as

 

Jij Si Sj − gμBBe ·

 

Si ,

 

Hmean field = i,j

i

(14.4.3)

where

2

 

 

 

 

 

 

 

 

 

 

 

Be = μ0H +

 

 

 

 

(14.4.4)

 

 

Jij Sj .

 

B

 

 

 

 

 

 

j

 

 

 

Using the magnetic moment instead of the spin,

 

 

 

 

2

 

 

 

 

 

Be = μ0H +

 

Jij μj .

 

(14.4.5)

g2μB2

 

 

 

 

 

 

j

 

 

 

Even in the absence of an external magnetic field, the e ective magnetic field

which depends on the average magnetic moment of the neighboring spins

can give rise to magnetic ordering. This e ective internal field is called the mean field or the Weiss field, as it was P. Weiss (1907) who gave the first phenomenological account of ferromagnetism based on the assumption of such an internal field. The term molecular field is also used.

14.4.1 The Mean-Field Theory of Ferromagnetism

In a ferromagnetically ordered state the mean value of the atomic magnetic moment is the same at each lattice point, therefore Be is independent of the choice of the lattice point. Making use of the relation M = N μi /V between magnetization and atomic moment,

Be = μ0H +

2V

 

 

(14.4.6)

 

 

 

Jij M .

N g2μ2

 

 

B

j

 

 

 

 

 

As in the ferromagnetic case (Jij > 0) the internal field is parallel to the external field, scalar quantities can be used:

Be = μ0H +

2V

 

 

(14.4.7)

 

 

 

Jij M .

N g2μ2

 

 

B

j

 

 

 

 

 

In the previous formula the exchange interaction appears in the combination

 

 

J0 = Jij .

(14.4.8)

j

Expressing the e ective field in terms of this,

14.4 The Mean-Field Approximation

475

Be = μ0H + λM , where

λ =

2V

 

J0 .

(14.4.9)

2

2

 

 

N g μ

 

 

 

 

 

B

 

 

The coe cient λ specifies the relation between the internal field and magnetization in the Weiss theory. If exchange occurs only with the z nearest neighbors then J0 = J z. When secondand third-neighbor interactions cannot be neglected, J0 is the weighted average of the various coupling constants. The necessary condition for ferromagnetic ordering is that J0 should be positive. This can happen even if some of the latter couplings are antiferromagnetic.

It was derived on page 57 in Section 3.2.6 on paramagnetism that the magnetization due to the magnetic moment of an atom of spin S in a magnetic field B is

M =

N

|g|μBSBS (β|g|μBSB) ,

(14.4.10)

V

where BS (x) is the Brillouin function. In the mean-field theory this expression has to be modified by replacing the magnetic induction B by the e ective field

Be , leading to

N

 

 

M =

|g|μBSBS (β|g|μBSBe ) .

(14.4.11)

V

In contrast to the paramagnetic case, the e ective field itself also depends on magnetization. Therefore the self-consistent equation

M =

N

|g|μBSBS β|g|μBS(μ0H + λM )

(14.4.12)

V

has to be solved now.

In the absence of an external magnetic field the argument of the Brillouin function is

x = β|g|μBSλM =

2V SJ0M

(14.4.13)

N |g|μBkBT .

Expressing condition (14.4.12) of self-consistency in terms of this variable,

kBT

 

2J0S2 x = BS (x) .

(14.4.14)

The solutions to this equation are easily obtained graphically by plotting the leftand right-hand sides separately, as in Fig. 14.11. The thermal average of the magnetic moment can be read o from the intersection points.

The slope of the expression on the left-hand side is proportional to temperature. Therefore above a certain critical temperature TC only the trivial solution x = 0 – and consequently M = 0 – exists. The same solution exists also below TC, however it becomes unstable. In this regime there exists another solution with finite magnetization whose free energy is lower.

According to the expansion (3.2.84), the slope of the Brillouin function at x = 0 is (S +1)/3S, thus the equation that determines the critical temperature of the ferromagnetic state, the Curie temperature is

476 14 Magnetically Ordered Systems

Fig. 14.11. Graphical determination of the mean-field magnetization from the intersection points of the Brillouin function and straight lines of temperature-dependent slope

 

 

kBTC

=

S + 1

,

 

(14.4.15)

 

 

2J0S2

3S

 

 

 

 

 

 

from which

 

 

 

 

 

 

TC =

2J0S(S + 1)

.

(14.4.16)

 

 

 

 

 

 

3kB

 

 

 

Spontaneous magnetization vanishes above TC. However, an external magnetic field aligns the magnetic moments, at least partially. The spin system then behaves paramagnetically. To determine the magnetic susceptibility, we shall take (14.4.11) and (14.4.12), but this time in the presence of a magnetic field. For weak fields it is su cient to keep the leading, linear term in the expansion of the right-hand side, obtained via (3.2.84). In this approximation

M =

N

g μ

S

S + 1

 

|g|μB

SB

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

V

| |

B

3S kBT

 

e

 

N g2

μB2

 

 

 

 

V

| |

 

 

3S kBT

 

 

 

 

=

N

g μBS

S + 1

 

|g|μB

S μ0H +

2V

J0 M .

 

 

 

 

 

 

 

 

 

 

 

Substituting TC from the boxed formula,

 

 

 

 

 

 

 

 

 

 

 

 

N g2μ2 S(S + 1)

 

k T

C

 

 

 

M =

 

 

 

 

 

B

 

 

 

 

μ0H +

 

B

 

M .

 

 

 

V k T

3

 

 

k T

 

 

 

 

 

 

 

 

 

B

 

 

 

 

 

B

 

 

 

Solving this equation for magnetization,

 

N g2μ2

S(S + 1)

 

M =

 

 

B

 

 

μ0H .

V

 

3kB(T − TC)

 

(14.4.17)

(14.4.18)

(14.4.19)

Thus in the paramagnetic region the susceptibility satisfies the Curie–Weiss law :

14.4 The Mean-Field Approximation

477

 

N g2μ2

μ0S(S + 1)

 

 

χ =

 

 

B

 

 

.

(14.4.20)

V

 

3kB(T − TC)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

It is immediately seen that the susceptibility diverges as the critical temperature TC of the ferromagnetic transition is approached.

Below but close to TC an explicit formula can be obtained for the temperature dependence of spontaneous magnetization as well by taking into account the first correction beyond the linear term in the expansion of the Brillouin function. The self-consistent solution of the equation then leads to

M (TC − T )1/2 .

(14.4.21)

In the critical point itself magnetization is a nonlinear function of the magnetic field, as (14.4.18) shows that the term which is linear in magnetization drops out at T = TC. This leads to

M H1/3 .

(14.4.22)

(a)

(b)

Fig. 14.12. Comparison of the measured magnetic properties of nickel [P. Weiss and R. Forrer, Ann. de Phys. 5, 153 (1926)] with the results obtained in the mean-field theory for S = 1/2. (a) Magnetization and (b) inverse susceptibility, as functions of temperature

At low temperatures magnetization tends to the saturation value M0 = (N/V )|g|μBS. Using the expansion coth x ≈ 1 + 2e2x in the form (3.2.79) of the Brillouin function for large values of its argument, the leading-order contribution is

1

 

BS (x) 1 S e−x/S .

(14.4.23)

478 14 Magnetically Ordered Systems

In this region the first correction to the temperature dependence of magnetization is obtained by replacing the magnetization by its saturation value on the right-hand side of (14.4.12). Then

M = M0

1 S exp

S + 1 T .

(14.4.24)

 

1

 

 

3 TC

 

At intermediate temperatures no analytical expression can be derived. Figure 14.12 shows the measured temperature dependence of magnetization and inverse susceptibility for nickel, compared with the results obtained in the mean-field approximation for a spin S = 1/2 using the Brillouin function, and from the Curie–Weiss law.

A similar behavior is observed in other ferromagnetic materials, too. This leads to the conclusion that the mean-field theory provides a qualitatively good description for the finite-temperature behavior of the properties of ferromagnetic materials, however, its quantitative predictions are incorrect at low temperatures as well as close to the critical point – where analytical results can be obtained. To improve the description, one has to go beyond the mean-field approximation. At low temperatures a more accurate treatment of the low-energy excited states is required, while in the vicinity of the critical point critical fluctuations need to be taken into account.

14.4.2 The Mean-Field Theory of Antiferromagnetism

We have seen that even in cubic crystals various antiferromagnetic structures may occur. The reason for this is that besides the exchange interaction between nearest-neighbor moments, secondand third-neighbor couplings also play a role in determining the relative orientation of magnetic moments. This can be demonstrated most easily on the example of structures formed in a simple cubic lattice.

We shall consider the spins Si localized at lattice points as classical vectors, expand the spin density

 

 

S(r) = Siδ(r − Ri)

(14.4.25)

i

into a Fourier series, and assume that the magnetic structure can be represented by a single Fourier component, i.e.,

S(r) = Sq eiq·r ,

(14.4.26)

which is equivalent to saying that localized spins can be given in the form

Si = Sq eiq·Ri .

(14.4.27)

This choice is justified for collinear structures, where the upward or downward direction of the spin vector is determined by the phase factors. Taking the