7.2 Origin and Consequences of Magnetic Order |
407 |
|
|
In a common representation, the Hubbard Hamiltonian is
|
H = ∑ |
k,σ |
εk a† |
akσ |
+ |
I |
∑ |
nασ nα,−σ , |
(7.199) |
|
2 |
|
|
kσ |
|
|
α,σ |
|
|
where σ labels the spin (up or down), k labels the band energies, and α labels the lattice sites (we have assumed only one band—say an s-band—with εk being the band energy for wave vector k). The a†kσ and akσ are creation and annihilation operators and I defines the interaction between electrons on the same site.
It is important to notice that the Hubbard Hamiltonian (as written above) assumes the electron–electron interactions are only large when the electrons are on the same site. A narrow band corresponds to localization of electrons. Thus, the Hubbard Hamiltonian is often said to be a narrow s-band model. The nασ are Wannier site-occupation numbers. The relation between band and Wannier (site localized) wave functions is given by the use of Fourier relations:
ψk = |
1 |
∑R |
|
exp(−ik Rα )W (r − Rα ) , |
(7.200a) |
|
N |
α |
|
|
|
W (r − Rα ) = |
1 |
∑k exp(ik Rα )ψk (r) . |
(7.200b) |
|
|
|
|
N |
|
|
Since the Bloch (or band) wave functions ψk are orthogonal, it is straightforward to show that the Wannier functions W(r − Rα) are also orthogonal. The Wannier functions W(r − Rα) are localized about site α and, at least for narrow bands, are well approximated by atomic wave functions.
Just as a†kσ creates an electron in the state ψk [with spin σ either + or ↑ (up) or − ↓ (down)], so c†ασ (the site creation operator) creates an electron in the state W(r − Rα), again with the spin either up or down. Thus, occupation number operators for the localized Wannier states are n†ασ = c†ασcασ and consistent with (7.200a) the two sets of annihilation operators are related by the Fourier transform
|
akσ = |
|
1 |
∑R exp(ik Rα )cασ . |
(7.201) |
|
|
|
|
N |
α |
|
|
|
|
Substituting this into the Hubbard Hamiltonian and defining |
|
|
Tαβ = |
1 |
|
∑k εk exp[ik (Rα − Rβ )] , |
(7.202) |
|
N |
|
|
|
|
|
|
|
|
|
we find |
|
|
|
|
|
|
|
|
H = ∑ |
Tαβcβσ+ cασ + |
I |
∑ |
nασ+ nα−σ . |
(7.203) |
|
|
|
α,β,σ |
|
|
2 |
α,σ |
|
|
This is the most common form for the Hubbard Hamiltonian. It is often further assumed that Tαβ is only nonzero when α and β are nearest neighbors. The first term then represents nearest-neighbor hopping.
408 7 Magnetism, Magnons, and Magnetic Resonance
Since the Hamiltonian is a many-electron Hamiltonian, it is not exactly solvable for a general lattice. We solve it in the mean-field approximation and thus replace
2I ∑α,σ nασ nα,−σ ,
with
where nα,−σ is the thermal average of pendent of site and so write it down as n−σ
nα,−σ. We also assume nα,−σ is indein (7.204).
Itinerant Ferromagnetism and the Stoner Model (B)
The mean-field approximation has been criticized on the basis that it builds in the possibility of an ordered ferromagnetic ground state regardless of whether the Hubbard Hamiltonian exact solution for a given lattice would predict this. Nevertheless, we continue, as we are more interested in the model we will eventually reach (the Stoner model) than in whether the theoretical underpinnings from the Hubbard model are physical. The mean-field approximation to the Hubbard model gives
H = ∑ |
Tαβ c† |
cασ + I ∑ |
α,σ |
n−σ nασ . |
(7.204) |
α,β,σ |
βσ |
|
|
|
Actually, in the mean-field approximation, the band picture is more convenient to use. Since we can show
∑ nασ = ∑ |
k |
nkσ , |
|
α |
|
|
the Hubbard model in the mean field can then be written as |
|
H = ∑k,σ (εk + In−σ )nkσ . |
(7.205) |
The single-particle energies are given by |
|
|
|
Ek,σ = εk + In−σ . |
(7.206) |
The average number of electrons per site n is less than or equal to 2 and n = n+ + n−, while the magnetization per site n is M = (n+ − n−)μB, where μB is the Bohr magneton.
Note: In order not to introduce another “−” sign, we will say “spin up” for now. This really means “moment up” or spin down, since the electron has a negative charge.
7.2 Origin and Consequences of Magnetic Order |
409 |
|
|
Note n + (M/μB) = 2n+ and n − (M/μB) = 2n−. Thus, up to an additive constant
|
|
|
M |
|
|
|
Ek ± = εk + I |
|
. |
(7.207) |
|
|
|
|
|
2μB |
|
Note (7.207) is consistent with (7.197b). write the following basic equations for the
If we then define Heff = IM/2μB2, we Stoner model:
|
|
|
M = μB (n↑ − n↓) , |
|
(7.208) |
|
|
|
Ek,σ = εk μB Heff , |
|
(7.209) |
|
|
|
|
Heff |
= |
IM |
|
, |
|
|
(7.210) |
|
|
|
|
2μB2 |
|
|
|
|
|
|
|
|
|
|
|
|
nσ = |
1 |
∑k |
|
1 |
|
|
, |
(7.211) |
|
N |
exp[(Ekσ − Mμ) / kT ] +1 |
|
|
|
|
|
|
|
|
|
n↑ + n↓ = n . |
|
|
|
(7.212) |
Although these equations are easy to write down, it is not easy to obtain simple convenient solutions from them. As already noted, the Stoner model contains two basic assumptions: (1) The electronic energy band in the metal is described by a known εk. By standard means, one can then derive a density of states. For free electrons, N(E) (E)1/2. (2) A molecular field approximately describes the effects of the interactions and we assume Fermi–Dirac statistics can be used for the spinup and spin-down states. Much of the detail and even standard notation has been presented by Wohlfarth [7.69]. See also references to Stoner’s work in the works by Wohlfarth.
The only consistent way to determine εk and, hence, N(E) is to derive it from the Hubbard Hamiltonian. However, following the usual Stoner model we will just use an N(E) for free electrons.
The maximum saturation magnetization (moment per site) is M0 = μBn and the actual magnetization is M = μB(n↑ − n↓). For the Stoner model, a relative magnetization is defined below:
ξ = |
M |
= |
n↑ − n↓ |
. |
(7.213) |
M0 |
|
|
|
n |
|
Using (7.212) and (7.213), we have
n |
= n |
= (1+ξ) |
|
n |
, |
(7.214a) |
|
|
+ |
↑ |
2 |
|
|
|
|
|
|
|
n |
= n |
= (1−ξ) |
n |
. |
(7.214b) |
|
− |
↓ |
2 |
|
|
|
|
|
|
|
410 7 Magnetism, Magnons, and Magnetic Resonance
It is also convenient to define a temperature θ′, which measures the strength of the exchange interaction
We now suppose that the exchange energy is strong enough to cause an imbalance in the number of spin-up and spin-down electrons. We can picture the situation with constant Fermi energy μ = EF (at T = 0) and a rigid shifting of the up N+ and the down N− density states as shown in Fig. 7.15.
The ↑ represents the “spin-up” (moment up actually) band and the ↓ the “spindown” band. The shading represents states filled with electrons. The exchange energy causes the splitting of the two bands. We have pictured the density of states by a curve that goes to zero at the top and bottom of the band unlike a freeelectron density of states that goes to zero only at the bottom.
Fig. 7.15. Density states imbalanced by exchange energy
At T = 0, we have
n+ = (1+ξ) |
|
n |
|
= ∫ |
N+(E)dE , |
(7.216a) |
|
|
2 |
|
occ.states |
|
|
n− = (1 −ξ) |
n |
|
= ∫ |
N−(E)dE . |
(7.216b) |
|
|
2 |
|
occ.states |
|
|
This can be easily worked out for free electrons if E = 0 at the bottom of both bands,
N±(E) = |
1 |
Ntotal (E) = |
1 |
|
2m 3/ 2 |
E ≡ N (E) . |
(7.217) |
2 |
4π 2 |
|
2 |
|
|
|
|
|
|
|
7.2 Origin and Consequences of Magnetic Order |
411 |
|
|
We now derive conditions for which the magnetized state is stable at T = 0. If we just use a single-electron picture and add up the single-electron energies, we find, with the (–) band shifted up by and the (+) band shifted down by , for the energy per site
E = n− |
+ ∫EF− |
EN(E)dE − n+ |
+ ∫EF+ |
EN(E)dE . |
|
0 |
|
0 |
|
The terms involving |
are the exchange energy. We can rewrite it from (7.208), |
(7.213), and (7.215) as |
|
|
|
|
However, just as in the Hartree–Fock analysis, this exchange term has double counted the interaction energies (once as a source of the field and once as interaction with the field). Putting in a factor of 1/2, we finally have for the total energy
E = ∫EF+ |
EN (E)dE + ∫EF− |
EN (E)dE − |
1 nkθ′ξ2 . |
(7.218) |
0 |
0 |
|
2 |
|
Differentiating (d/dξ) (7.216) and (7.218) and combining the results, we can show
|
1 dE |
= |
1 |
(EF+ |
− EF− ) − kθ′ξ . |
|
|
|
|
|
n dξ |
2 |
|
|
|
|
Differentiating (7.219) a second time and again using (7.216), we have
1 d |
2 |
E |
|
n |
|
1 |
|
1 |
|
|
|
= |
|
+ |
|
− kθ′ . |
|
|
|
|
|
|
|
n dξ 2 |
|
|
|
|
4 N (EF+ ) |
|
N (EF− ) |
|
Setting dE/dξ = 0, just gives the result that we already know
2kθ′ξ = (EF+ − EF− ) = 2μB Heff = 2 .
Note if ξ = 0 (paramagnetism) and dE/dξ = 0, while d2E/dξ2 < 0 the paramagnetism is unstable with respect to ferromagnetism. ξ = 0, dE/dξ = 0 implies EF+ = EF- and N(EF-) = N(EF+) = N(EF). So by (7.220) with d2E/dξ2 ≤ 0 we have
kθ′ ≥ |
n |
|
2N (EF ) . |
(7.221) |
For a parabolic band with N(E) E1/2, this implies
−1/3
412 7 Magnetism, Magnons, and Magnetic Resonance
We now calculate the relative magnetization (ξ0) at absolute zero for a parabolic band where N(E) = K(E)1/2 where K is a constant. From (7.216)
(1+ξ0 ) |
n |
= |
|
2 |
K (EF+ )3/ 2 , |
|
|
3 |
2 |
|
|
|
|
|
(1−ξ0 ) |
n |
|
= |
|
2 |
K (EF− )3/ 2 . |
|
3 |
2 |
|
|
|
|
|
Also |
|
|
|
|
|
n = |
|
4 |
|
KEF3/ 2 . |
3 |
|
|
|
|
|
|
|
|
Eliminating K and using EF+ − EF− = 2kθ′ξ0, we have
kθ′ |
= |
1 |
|
[(1+ξ0 )2 / 3 − (1−ξ0 )2 / 3] , |
(7.223) |
|
2ξ |
0 |
EF |
|
|
which is valid for 0 ≤ ξ0 ≤ 1. The maximum ξ0 can be is 1 for which kθ′/EF = 2 , and at the threshold for ferromagnetism ξ0 is 0. So, kθ′/EF = 2/3 as already predicted by the Stoner criterion.
Summary of Results at Absolute Zero
We have three ranges:
|
|
kθ′ |
< |
|
2 |
|
= 0.667 and |
ξ0 |
= |
M |
= 0 , |
|
|
|
|
|
3 |
|
|
|
|
EF |
|
|
|
|
|
|
|
|
nμB |
2 |
|
< |
kθ′ |
< |
|
1 |
|
= 0.794 , |
0 < |
ξ0 = |
M |
<1, |
3 |
|
|
|
|
21 3 |
|
|
|
|
EF |
|
|
|
|
|
|
|
|
|
|
nμB |
|
|
|
|
kθ′ |
> |
|
1 |
|
|
and ξ0 = |
M |
|
=1 . |
|
|
|
|
|
|
21 3 |
|
nμB |
|
|
|
|
|
EF |
|
|
|
|
|
|
|
|
|
The middle range, where 0 < ξ0 < 1 is special to Stoner ferromagnetism and not to be found in the Weiss theory. This middle range is called “unstructured” or “weak” ferromagnetism. It corresponds to having electrons in both ↑ and ↓ bands. For very low, but not zero, temperatures, one can show for weak ferromagnetism that
where C is a constant. This is particularly easy to show for very weak ferromagnetism, where ξ0 << 1 and is left as an exercise for the reader.
7.2 Origin and Consequences of Magnetic Order |
413 |
|
|
We now discuss the case of strong ferromagnetism where kθ′/EF > 2−1/3. For this case, ξ0 = 1, and n↑ = n, n↓ = 0. There is now a gap Eg between EF+ and the bottom of the spin-down band. For this case, by considering thermal excitations to the n↓ band, one can show at low temperature that
M = M0 − K′′T 3/ 2 exp(−Eg / kT ) , |
(7.225) |
where K″ is a constant. However, spin-wave theory says M = M0 − C′T3/2, where C′ is a constant, which agrees with low-temperature experiments. So, at best, (7.225) is part of a correction to low-temperature spin-wave theory.
Within the context of the Stoner model, we also need to talk about exchange enhancement of the paramagnetic susceptibility χP (gaussian units with μ0 = 1)
M = χ |
P |
BTotal , |
(7.226) |
|
eff |
|
where M is the magnetization and χP the Pauli susceptibility, which for low temperatures, has a very small αT2 term. It can be written
χP = 2μB2 N (EF )(1+αT 2 ) , |
(7.227) |
where N(E) is the density of states for one subband. Since |
|
BTotal = H |
eff |
+ B = γB + B, |
|
eff |
|
|
it is easy to show that (gaussian with B = H)
χ = |
M |
= |
|
χP |
, |
(7.228) |
B |
|
1 −γχP |
where 1/(1 − γχP) is the exchange enhancement factor.
We can recover the Stoner criteria from this at T = 0 by noting that paramagnetism is unstable if
By using γ = kθ′/nμB2 and χP0 = 2μB2N(EF), (7.229) just gives the Stoner criteria. At finite, but low temperatures where (α = –|a|)
χP = χP0 (1− | a | T 2 ) ,
if we define
θ 2 = γχP0 −1 , γχP0 | a |
and suppose |a|T2 << 1, it is easy to show
|
χ = |
1 |
|
1 |
. |
|
γ | a | T 2 −θ 2 |
|
|
|
414 7 Magnetism, Magnons, and Magnetic Resonance
Thus, as long as T θ, we have a Curie–Weiss-like law:
χ = |
1 |
1 |
. |
(7.230) |
2θγ | a | |
|
T −θ |
At very high temperatures, one can also show that an ordinary Curie–Weiss-like law is obtained:
χ = |
nμB2 |
1 |
. |
(7.231) |
k |
|
T −θ |
|
|
|
|
Summary Comments About the Stoner Model
1.The low-temperature results need to be augmented with spin waves. Although in this book we only derive the results of spin waves for the localized model, it turns out that spin waves can also be derived within the context of the itinerant electron model.
2.Results near the Curie temperature are never qualitatively good in a mean-field approximation because the mean-field approximation does not properly treat fluctuations.
3.The Stoner model gives a simple explanation of why one can have a fractional number of electrons contributing to the magnetization (the case of weak ferromagnetism where ξ0 = MT=0/nμB is between 0 and 1).
4.To apply these results to real materials, one usually needs to consider that there are overlapping bands (e.g. both s and d bands), and not all bands necessarily split into subbands. However, the Stoner model does seem to work for ZrZn2.
7.2.5 Magnetic Phase Transitions (A)
Simple ideas about spin waves break down as Tc is approached. We indicate here one way of viewing magnetic phenomena near the T = Tc region. In this Section we will discuss magnetic phase transitions in which the magnetization (for ferromagnets with H = 0) goes continuously to zero as the critical temperature is approached from below. Thus at the critical temperature (Curie temperature for a ferromagnet) the ordered (ferromagnetic) phase goes over to the disordered (paramagnetic) phase. This “smooth” transition from one phase (or more than one phase in more general cases) to another is characteristic of the behavior of many substances near their critical temperature. In such continuous phase transitions there is no latent heat and these phase transitions are called second-order phase transitions. All second-order phase transitions show many similarities. We shall consider only phase transitions in which there is no latent heat.
No complete explanation of the equilibrium properties of ferromagnets near the magnetic critical temperature (Tc) has yet been given, although the renormalization technique, referred to later, comes close. At temperatures well below Tc we know that the method of spin waves often yields good results for describing the
7.2 Origin and Consequences of Magnetic Order |
415 |
|
|
magnetic behavior of the system. We know that high-temperature expansions of the partition function yield good results. The Green function method provides results for interesting physical quantities at all temperatures. However, the Green function results (in a usable approximation) are not valid near Tc. Two methods (which are not as straightforward as one might like) have been used. These are the use of scaling laws19 and the use of the Padé approximant.20 These methods often appear to give good quantitative results without offering much in the way of qualitative insight. Therefore we will not discuss them here. The renormalization group, referenced later, in some ways is a generalization of scaling laws. It seems to offer the most in the way of understanding.
Since the region of lack of knowledge (around the phase transition) is only near τ = 1 (τ = T/Tc, where Tc is the critical temperature) we could forget about the region entirely (perhaps) if it were not for the fact that very unusual and surprising results happen here. These results have to do with the behavior of the various
quantities as a function of temperature. For example, the Weiss theory predicts for the (zero field) magnetization that M (Tc − T)+1/2 as T → Tc− (the minus sign
means that we approach Tc from below), but experiment often seems to agree better with M (Tc − T)+1/3. Similarly, the Weiss theory predicts for T > Tc that the
zero-field susceptibility behaves as χ (T − Tc)−1, whereas experiment for many materials agrees with χ (T − Tc)−4/3 as T → Tc+. In fact, the Weiss theory fails very seriously above Tc because it leaves out the short-range ordering of the spins. Thus it predicts that the (magnetic contribution to the) specific heat should vanish above Tc, whereas the zero-field magnetic specific heat does not so vanish. Using an improved theory that puts in some short-range order above Tc modifies the specific heat somewhat, but even these improved theories [92] do not fit experiment well near Tc. Experiment appears to suggest (although this is not settled yet) that for many materials C ln |(T − Tc)| as T → Tc+ (the exact solution of the specific heat of the two-dimensional Ising ferromagnet shows this type of divergence), and the concept of short-range order is just not enough to account for this logarithmic or near logarithmic divergence. Something must be missing. It appears that the missing concept that is needed to correctly predict the “critical exponents” and/or “critical divergences” is the concept of (anomalous) fluctuations. [The exponents 1/3 and 4/3 above are critical exponents, and it is possible to set up the formalism in such a way that the logarithmic divergence is consistent with a certain critical exponent being zero.] Fluctuations away from the thermodynamic equilibrium appear to play a very dominant role in the behavior of thermodynamic functions near the phase transition. Critical-point behavior is discussed in more detail in the next section.
Additional insight into this behavior is given by the Landau theory.19 The Landau theory appears to be qualitatively correct but it does not predict correctly the critical exponents.
19See Kadanoff et al [7.35].
20See Patterson et al [7.54] and references cited therein.
416 7 Magnetism, Magnons, and Magnetic Resonance
Critical Exponents and Failures of Mean-Field Theory (B)
Although mean-field theory has been extraordinarily useful and in fact, is still the “workhorse” of theories of magnetism (as well as theories of the thermodynamics behavior of other types of systems that show phase transitions), it does suffer from several problems. Some of these problems have become better understood in recent years through studies of critical phenomena, particularly in magnetic materials, although the studies of “critical exponents” relates to a much broader set of materials than just magnets as referred to above. It is helpful now to define some quantities and to introduce some concepts.
A sensitive test of mean-field theory is in predicting critical exponents, which define the nature of the singularities of thermodynamic variables at critical points of second-order phase transitions. For example,
|
T |
|
−T |
|
β |
|
T |
|
−T |
|
|
−ν |
|
|
|
|
|
|
φ ~ |
c |
|
|
|
and ξ = |
c |
|
|
|
, |
|
|
|
|
|
|
|
|
T |
|
|
|
|
T |
|
|
|
|
|
|
c |
|
|
|
|
|
c |
|
|
|
for T < Tc, where β, ν are critical exponents, φ is the order parameter, which for ferromagnets is the average magnetization M and ξ is the correlation length. In magnetic systems, the correlation length measures the characteristic length over which the spins are ordered, and we note that it diverges as the Curie temperature Tc is approached. In general, the order parameter φ is just some quantity whose value changes from disordered phases (where it may be zero) to ordered phases (where it is nonzero). Note for ferromagnets that φ is zero in the disordered paramagnetic phase and nonzero in the ordered ferromagnetic situation.
Mean-field theory can be quite good above an upper critical (spatial) dimension where by definition it gives the correct value of the critical exponents. Below the upper critical dimension (UCD), thermodynamic fluctuations become very important, and mean-field theory has problems. In particular, it gives incorrect critical exponents. There also exists a lower critical dimension (LCD) for which these fluctuations become so important that the system does not even order (by definition of the LCD). Here, mean-field theory can give qualitatively incorrect results by predicting the existence of an ordered phase. The lower critical dimension is the largest dimension for which long-range order is not possible. In connection with these ideas, the notion of a universality class has also been recognized. Systems with the same spatial dimension d and the same dimension of the order parameter D are usually in the same universality class. Range and symmetry of the interaction potential can also play a role in determining the universality class. Quite dissimilar systems in the same universality class will, by definition, exhibit the same critical exponents. Of course, the order parameter itself as well as the critical temperature Tc, may be quite different for systems in the same universality class. In this connection, one also needs to discuss concepts like the renormalization group, but this would take us too far afield. Reference can be made to excellent statistical mechanics books like the one by Huang.21
21 See Huang [7.32, p441ff].