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

Patterson, Bailey - Solid State Physics Introduction to theory

.pdf
Скачиваний:
1134
Добавлен:
08.01.2014
Размер:
7.07 Mб
Скачать

7.4 Magnetic Resonance and Crystal Field Theory

437

 

 

When we include a relaxation time (or an interaction process), we find that the time rate of change of the magnetization (along the field) is proportional to the deviation of the magnetization from its equilibrium value. This guarantees a relaxation of magnetization along the field. If we add an alternating magnetic field along the x- or y-axes, it is also necessary to add a term (M × H)z that is proportional to the torque. Thus for the component of magnetization along the constant external magnetic field, it is reasonable to write

dM z =

M0 M z

+ (γμ0 )(M × H )z .

(7.265)

 

dt

T1

 

As noted, (7.265) has a built-in relaxation process of Mz to M0, the spin-lattice relaxation time T1. However, as we approach equilibrium in a static magnetic field H0kˆ, we will want both Mx and My to tend to zero. For this purpose, a new term with a relaxation time T2 is often introduced. We write

 

dM x

= γμ0 (M × H )x

M x

,

(7.266)

 

dt

T2

 

 

 

 

 

 

 

and

 

 

 

 

 

 

 

 

dM y

= γμ0

(M × H ) y

 

M y

 

.

(7.267)

 

dt

 

T2

 

 

 

 

 

 

 

 

Equations (7.265), (7.266), and (7.267) are called the Bloch equations. T2 is often called the spin-spin relaxation time. The idea is that the term involving T1 is caused by the interaction of the spin system with the lattice or phonons, while the term involving T2 is caused by something else. The physical origin of T2 is somewhat complicated. Consider, for example, two nuclei precessing in an external static magnetic field. The precession of one nucleus produces a varying magnetic field at the second nucleus and hence tends to “flip” the spin of the second nucleus (and vice versa). Waller27 first pointed out that there are two different types of spin relaxation processes.

Ferromagnetic Resonance (B)

Using a simple quantum picture, for an atomic system, we have already argued (see (7.258))

dμ

= γμ× B ,

(7.268)

dt a

where Ba = μ0H. This implies a precession of μ and M about the constant magnetic field Ba with frequency ω = γBa the Larmor frequency, as already noted. For

27See Waller [7.66]. Discussion of ways to calculate T1 and T2 is contained in White [7.68 p124ff and 135ff].

438 7 Magnetism, Magnons, and Magnetic Resonance

ferromagnetic resonance (FMR) all spins precess together and M = , where N is the number of spins per unit volume. Thus by (7.268)

dM

= γM × Ba .

(7.269)

dt

 

 

Several comments can be made. The above equation is valid also for M = M(r) varying slowly in space. We will also use this equation for spin-wave resonance when the wavelengths of the waves are long compared to the atom to atom spacing that allows the classical approach to be valid. One generalizes the above equation by replacing Ba by B where

B = Ba

(applied)

+ Brf

(due to a radio-frequency applied field)

+ Bdemag

(from demagnetizing fields that depend on geometry)

+Bexchange (as derived from the Heisenberg Hamiltonian)

+Banisotropy (an effective field arising from interactions producing anisotropy).

We should also include dissipative or damping and relaxation effects.

We start with all fields zero or negligible except for the applied field (note here

Bexchange M, which is assumed to be uniform, so M × Bexchange = 0). This gives resonance at the natural precessional frequency of the uniform precessional mode.

With B = B0kˆ we have

 

dM x

= γM

B ,

 

dM y

 

 

= −γM

x

B ,

dM z

= 0.

(7.270)

 

 

 

 

 

 

 

 

 

dt

y 0

 

 

 

dt

 

 

 

 

 

0

dt

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

We look for solutions with

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

M

x

= A eiωt ,

 

 

 

 

 

 

 

 

 

 

 

 

1

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

M

 

y

= A eiωt ,

 

 

 

 

 

(7.271)

 

 

 

 

 

 

 

2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

M z

= constant,

 

 

 

 

 

and so we have a solution provided

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

iω

 

 

γB0

 

= 0 ,

 

 

 

(7.272)

 

 

 

 

 

 

 

 

 

 

 

 

γB

 

 

 

 

iω

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

0

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

or

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

ω

 

=

 

γB0

 

 

,

 

 

 

 

 

(7.273)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

7.4 Magnetic Resonance and Crystal Field Theory

439

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Fig. 7.30. (a) Thin film with magnetic field, (b) “Unpinned” spin waves, (c) “Pinned” spin waves

which as expected is just the Larmor precessional frequency. In actual situations we also need to include demagnetization fields and hence shape effects, which will alter the resonant frequencies. FMR typically occurs at microwave frequencies. Antiferromagnetic resonance (AFMR) has also been studied as a way to determine anisotropy fields.

Spin-Wave Resonance (A)

Spin-wave resonance is a direct way to experimentally prove the existence of spin waves (as is inelastic neutron scattering – see Kittel [7.39 pp456-458]). Consider a thin film with a magnetic field B0 perpendicular to the film (Fig. 7.30a). In the simplest picture, we view the spin waves as “vibrations” in the spin between the surfaces of the film. Plotting the amplitude versus position, Fig. 7.30b is obtained for unpinned spins. Except for the uniform mode, these have no net interaction (absorption) with the electromagnetic field. The pinned case is a little different (Fig. 7.30c). Here only waves with an even number of half-wavelengths will show no net interaction energy with the field while the ones with an odd number of halfwavelengths (n = 1, 3, etc.) will absorb energy. (Otherwise the induced spin flippings will absorb and emit equal amounts of energy).

We get absorption when

n

λ

= T

{n odd, T

thickness of film},

 

2

2π

 

 

nπ

 

 

π

 

k =

=

or

k = (2n +1)

{n = 0,1, 2}.

λ

T

T

 

 

 

 

 

 

440 7 Magnetism, Magnons, and Magnetic Resonance

With applied field normal to film and with demagnetizing field and exchange Dk2, absorption will occur for

ω

0

= γ (B μ

M ) + Dk 2

(SI) ,

 

0

0

 

 

where M is the static magnetization in the direction of B0. The spin-wave frequency is determined by both the FMR frequency (the first term including demagnetization) and the dispersion relation typical for spin waves.

We now analyze spin-wave resonance in a little more detail. First we develop the Heisenberg Hamiltonian in the continuum approximation,

H = −Jij Si S j = −

1

μi Biex

(7.274)

2

defines the effective field Biex acting on the moment at site i, μi = γSi. (γ < 0 for electrons)

Biex =

2

j Jij S j .

(7.275)

γ

 

 

 

Assuming nearest neighbors (nn) at distance a and nn interactions only. We find for a simple cubic (SC) structure after expansion, and using cancellation resulting from symmetry

γBex

=12JS

i

+ 2Ja2 2 S

.

 

i

 

 

 

 

 

 

 

i

 

 

Consistent with the classical continuum approximation

 

 

M

=

 

Si

,

 

 

(7.276)

 

M

 

S

 

 

 

 

 

 

 

 

 

 

 

Bex = λM + K2 (M / M ),

(7.277)

where

 

 

 

 

 

 

 

 

 

 

 

λ =

12JS

 

, K′ =

 

2Ja2S

.

 

(7.278)

γM

 

γ

 

 

 

 

 

 

 

 

 

 

As an aside we note Bex is consistent with results obtained before (Sect. 7.3.1). Since

U = −

1

μi Biex ,

(7.279)

2

neglecting constant terms (resulting from the magnitude of the magnetization being constant) we have

U = −

JS 2

m 2mdV {m = M / M}.

(7.280)

a

 

 

 

 

 

 

 

 

 

7.4 Magnetic Resonance and Crystal Field Theory

441

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Fig. 7.31. Spin wave resonance spectrum for Ni film, room temperature, 17 GHz. After Puszharski H, “Spin Wave Resonance”, Magnetism in Solids Some Current Topics, Scottish Universities Summer School in Physics, 1981, p. 287, by permission of SUSSP. Original data in Mitra DP and Whiting JSS, J Phys F: Metal Physics, 8, 2401 (1978)

Assuming ∫mx mx·dA etc = 0 for a large surface we can also recast the above as

U =

JS 2

[( α1)2 + ( α2 )2 + ( α3)2 ]dV

(7.281)

a

 

 

 

which is the same as we obtained before, with a slightly different analysis. The αi are of course the direction cosines.

The anisotropy energy and effective field can be written in the same way as before, and no further comments need be made about it.

When one generalizes the equation for the time development of M, one has the Landau–Lifshitz equations. Damping causes broadening of the absorption lines. Then

M = γM × Beff +

α

M × (M × Beff ),

(7.282)

M

 

 

 

442 7 Magnetism, Magnons, and Magnetic Resonance

where α is a constant characterizing the damping. Spin-wave resonance has been observed as shown in Fig. 7.31. The integers label the modes of excitation. The figure is complicated by surface spin waves that are labeled 2, 1 and not fully resolved. Reference to the original paper must be made for complete details.

We have discussed Beff in the Section on FMR. Allowing M to vary with r and using the pinned boundary conditions, (7.282) can be used to quantitatively discuss SWR.

7.4.4 Crystal Field Theory and Related Topics (B)

This Section is primarily related to EPR. The general problem is to analyze the effects of neighboring ions on paramagnetic ions in a crystal. This cannot be exactly solved, and so we must seek physically reasonable simplifying assumptions.

Some atoms or ions when placed in a crystal act as if they undergo very little change. When this is so, we can predict the changes by perturbation theory. In order to estimate the perturbing effects of a host crystal on a paramagnetic ion, we ought to be able to treat the host crystal fairly crudely. For example, for an ionic crystal it might be sufficient to treat the ions as point charges. Then it would be fairly simple to estimate the change in the potential at the paramagnetic ion due to the host crystal. This potential energy could serve as a perturbation on the Hamiltonian of the paramagnetic ion.

Another simplification is possible. The crystal potential must have the symmetry of the point group describing the surroundings of the paramagnetic ion. As we will discuss later, group theory is useful in taking this into account.

The effect of the crystal field is to split the energy levels of a paramagnetic ion. In order to show how this comes about, it is useful to know what we mean by the energy levels. The best way to do this is to write down the Hamiltonian (whose eigenvalues are the energy levels) for the electrons. With no external field, the Hamiltonian has a form similar to

 

P2

 

 

Ze2

 

 

H = i

 

i

 

 

 

 

 

+ ai Ji I eφc (ri )

 

2m

4πε

0

r

 

 

 

 

 

 

 

 

 

i

 

(7.283)

 

 

1

 

e2

 

 

 

 

 

 

 

 

 

 

 

+ i, j 2

 

 

 

+ i, j λij Li S j .

 

4πε

0

r

 

 

 

 

 

 

 

ij

 

 

 

 

The origin of the coordinate system for (7.283) is the nucleus of the paramagnetic ion. The sum over i and j is a sum over electronic coordinates. The first term is the kinetic energy. The second term is the potential energy of the electrons in the field of the nucleus. The third term is the hyperfine interaction of the electron (with total angular momentum Ji) with the nucleus that has angular momentum I. The fourth term is the crystal field energy. The fifth term is the potential energy of the electrons interacting with themselves. The last term is the spin (Sj)-orbit (angular momentum Li) interaction (see Appendix F) of the electrons. By the unperturbed energy levels of the paramagnetic ion, one often means the energy eigenstates of the first, second, and fifth terms obtained perhaps by Hartree–Fock calculations. The

7.4 Magnetic Resonance and Crystal Field Theory

443

 

 

rest of the terms are usually thought of as perturbations. In the discussion that follows, the hyperfine interaction will be neglected.

To avoid complicated many-body effects, we will assume that the sources of the crystal field (Ec ≡ − φc) are external to the paramagnetic ion. Thus in the vicinity of the paramagnetic ion, it can be assumed that 2φc = 0.

Weak, Medium, and Strong Crystal Fields (B)

In discussing the effect of the crystal field on the energy levels, which is important to EPR, three cases can be distinguished [47].

Weak crystal fields are by definition those for which the spin-orbit interaction is stronger than the crystal field interaction. This is often realized when the electrons of the paramagnetic shell of the ion lie “fairly deep” within the ion, and hence are shielded from the crystalline field by the outer electrons. This may happen in ionic compounds of the rare earths. Rare earths have atomic numbers (Z) from 58 to 71. Examples are Ce, Pr, and Ne, which have incomplete 4f shells.

By a medium crystal field we mean that the crystal field is stronger than the spin-orbit interaction. This happens when the paramagnetic electrons of the ion are mainly distributed over the outer portions of the ion and hence are not well shielded. In this situation something else may occur. The potential that the paramagnetic ions move in is no longer even approximately spherically symmetric, and hence the orbital angular momentum is not conserved. We say that the orbital angular momentum is (at least partially) “quenched” (this means ψ|L|ψ = 0, ψ|L2|ψ ≠ 0). Paramagnetic crystals that have iron group elements (Z = 21 to 29, e.g., Cr, Mn, and Fe that have an incomplete 3d shell) are typical examples of the medium-field case.

Strong crystal field by definition means covalent bonding. In this situation, the wave functions for the paramagnetic ion electrons overlap considerably with the wave functions of the other electrons of the crystal. Crystal field theory does not work here. This type of situation will not be discussed in this chapter.

As we will see, group theory can be an aid in understanding how energy levels are split by perturbations.

Miscellaneous Theorems and Facts (In Relation to Crystal Field Theory) (B)

The theorems below will not be proved. They are stated because they are useful in carrying out actual crystal field calculations.

The Equivalent Operator Theorem. This theorem is used in calculating needed matrix elements in crystal field calculations. The theorem states that within a manifold of states for which l is constant, there are simple relations between the matrix elements of the crystal-field potential and appropriate angular momentum operators. For constant l, the rule says to replace the x by Lx (operator, in this case Lx is the x operator equivalent) and so forth for other coordinates. If the result is a product in which the order of the factors is important, then we must use all possible different permutations. There is a similar rule for manifolds of constant J

444 7 Magnetism, Magnons, and Magnetic Resonance

(where we include both the orbital angular momentum and the spin angular momentum).

There is a straightforward way of generating operator equivalents (OpEq) by using

[L+,Op Eq YlM ] Op Eq YlM +1 ,

and

[L,Op Eq Y M ] Op Eq Y M 1 .

(7.284)

l

l

 

The constants of proportionality can be computed from a knowledge of the Clebsch–Gordon coefficients.

Table 7.4. Effective magneton number for some representative trivalent lanthanide ions

Ion

Configuration

Ground state

g J (J +1) *

 

 

 

 

 

 

Pr (3+)

…4f 2

5s2

5p6

3H4

3.58

Nd (3+)

…4f 3

5s2

5p6

4I9/2

3.62

Gd (3+)

…4f 7

5s2

5p6

8S7/2

7.94

Dy (3+)

…4f 9

5s2

5p6

6H15/2

10.63

* g = g(Lande) =1 +

J (J +1) + S(S +1) L(L +1)

2J (J +1)

 

Table 7.5. Effective magneton number for some representative iron group ions*

Ion

Configuration

Ground state

2 S(S +1)

 

 

 

 

Fe (3+)

…3d5

6S5/2

5.92

Fe (2+)

…3d6

5D4

4.90

Co (2+)

…3d7

4F9/2

3.87

Ni (2+)

…3d8

3F4

2.83

* Quenching with J = S, L = 0 (so g = 2) is assumed for better agreement with experiment

Kramers’ Theorem. This theorem tells us about systems that must have a degeneracy. The theorem says that the systems with an odd number of electrons on which a purely electrostatic field is acting can have no energy levels that are less than two-fold degenerate. If a magnetic field is imposed, this two-fold degeneracy can be lifted.

Jahn–Teller effect. This effect tells us that high degeneracy may be unlikely. The theorem states that a nonlinear molecule that has a (orbitally) degenerate ground state is unstable, and tends to distort itself so as to lift the degeneracy. Because of the Jahn–Teller effect, the symmetry of a given atomic environment in a solid is

7.4 Magnetic Resonance and Crystal Field Theory

445

 

 

frequently slightly different from what one might expect. Of course, the Jahn– Teller effect does not remove the fundamental Kramers’ degeneracy.

Hund’s rules. Assuming Russel–Sanders coupling, these rules tell us what the ground state of an atomic system is. Hund’s rules were originally obtained from spectroscopic evidence, but they have been confirmed by atomic calculations that include the Coulomb interactions between electrons. The rules state that in figuring out how electrons fill a shell in the ground state we should (1) assign a maximum S allowed by the Pauli principle, (2) assign maximum L allowed by S, (3) assign J = L S when the shell is not half-full, and J = L + S when the shell is over half-full. See Problems 7.17 and 7.18. Results from the use of Hund’s rules are shown in Tables 7.9 and 7.10.

Energy-Level Splitting in Crystal Fields by Group Theory (A)

In this Section we introduce enough group theory to be able to discuss the relation between degeneracies (in the energies of atoms) and symmetries (of the environment of the atoms). The fundamental work in the field was done by H. A. Bethe (see, e.g., Von der Lage and Bethe [7.64]). For additional material see Knox and Gold [61, in particular see Table 1-2 pp. 5-8 for definitions].

We have already discussed some of the more elementary ideas related to groups in Chap. 1 (see Sect. 1.2.1). The most important new concept that we will introduce here is the concept of group representations. A group representation starts with a set of nonsingular square matrices. For each group element gi there is a matrix Ri such that gigj = gk implies that RiRj = Rk. Briefly stated, a representation of a group is a set of matrices with the same multiplication table as the original group.

Two representations (R′, R) of g that are related by

 

R(g) = S 1R(g)S

 

 

 

 

(7.285)

are said to be equivalent. In (7.285), S is any nonsingular matrix.

 

We define

 

 

 

 

 

 

 

 

R(g) R(1)

(g) R(2)

 

(1)

(g)

 

0

 

(7.286)

(g) R

 

 

.

 

 

 

0

R

(2)

 

 

 

 

 

 

(g)

 

In (7.286) we say that the representation R(g) is reducible because it can be reduced to the direct sum of at least two representations. If R(g) is of the form (7.286), it is said to be in block diagonal form. If a matrix representation can be brought into block diagonal form by a similarity transformation, then the representation is reducible. If no matrix representation reduces the representation to block diagonal form, then the matrix representation is irreducible. In considering any representation that is reducible, the most interesting information is to find out what irreducible representations are contained in the given reducible representation. We should emphasize that when we say a given representation R(g) is re-

446 7 Magnetism, Magnons, and Magnetic Resonance

ducible, we mean that a single S in (7.285) will put R′(g) in block diagonal form for all g in the group.

In a typical problem in crystal field theory, a reducible representation (with respect to some group) of interest might be the irreducible representation R(l) of the three-dimensional rotation group. That is, we would like to know what irreducible representations of a group of interest is contained in a given irreducible representation of R(l) for some l. As we will see later, this can tell us a good deal about what happens to the electronic energy levels of a spherical atom in a crystal field.

It is worthwhile to give an explicit example of the irreducible representations of a group. Let us consider the group D3 already defined in Chap. 1 (see Table 1.2).

In Table 7.6 note that R(1) and R(2) are unfaithful (many elements of the group correspond to the same matrix) representations while R(3) is a faithful (there is a one-to-one correspondence between group elements and matrices) representation. R(1) is, of course, the trivial representation.

Table 7.6. The irreducible representations of D3

D3

 

g1

 

 

g2

 

 

 

 

g3

 

 

 

g4

 

 

 

g5

 

 

 

g6

R(1)

 

1

 

 

1

 

 

 

 

1

 

 

 

1

 

 

 

1

 

 

 

1

R(2)

 

1

 

 

1

 

 

 

 

1

 

 

 

−1

 

 

 

−1

 

 

 

−1

R(3)

 

1, 0

1

 

1,

+ 3

 

1

 

1, 3

 

1

 

1, 3

 

1

 

1,

3

 

 

1, 0

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

2

 

 

2

 

2

 

2

 

 

 

 

0, 1

3,

1

 

3, 1

3, 1

3,

1

 

0, 1

Since a similarity transformation will induce so many equivalent irreducible representations, a quantity that is invariant to similarity transformation might be (and in fact is) of considerable interest. Such a quantity is the character. The character of a group element is the trace of the matrix representing that group element.

It is elementary to show that the trace is invariant to similarity transformation. A similar argument shows that all group elements in the same class28 have the same character. The argument goes as indicated below:

Tr(R(g)) = Tr(R(g)SS 1) = Tr(S 1R(g)S) = Tr(R(g)) ,

 

if R′(g) is defined by (7.285).

 

 

In summary the characters are defined by

 

 

χ(i) (g) = Rαα(i)

(g) .

(7.287)

α

 

 

Equation (7.287) defines the character of the group element g in the ith representation. The characters still serve to distinguish various representations. As an example, the character table for the irreducible representation of D3 is shown in Table 7.7. In Table 7.7, the top row labels the classes.

28Elements in the same class are conjugate to each other that means if g1 and g2 are in the same class there exists a g G g1 = g–1g2g.

Соседние файлы в предмете Химия