8 magnetism - tu berlin · 2017-08-20 · 8 magnetism 8.1 introduction the topic of this part of...

52
8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi- als are generally considered to be “non-magnetic”, which is a loose way of saying that they become magnetized only in the presence of an applied magnetic field (dia- and paramag- netism). We will see that in most cases these eects are very weak and the magnetization is lost as soon as the external field is removed. Much more interesting (also from a tech- nological point of view) are those materials which not only have a large magnetization, but also retain it even after the removal of the external field. Such materials are called permanent magnets (ferromagnetism). The property that like (unlike) poles of permanent magnets repel (attract) each other was already known to the Ancient Greeks and Chinese over 2000 years ago. They used this knowledge for instance in compasses. Since then, the importance of magnets has risen steadily. They play an important role in many modern technologies: Recording media (e.g. hard disks) : Data is recorded on a thin magnetic coating. The revolution in information technology owes as much to magnetic storage as to information processing with computer chips (i. e., the ubiquitous silicon chip). Credit, debit, and ATM cards : All of these cards have a magnetic strip on one side. TVs and computer monitors : Monitors contain a cathode ray tube that employs an electromagnet to guide electrons to the screen. Plasma screens and LCDs use dierent technologies. Speakers and microphones : Most speakers employ a permanent magnet and a current- carrying coil to convert electric energy (the signal) into mechanical energy (move- ment that creates the sound). Electric motors and generators : Most electric motors rely upon a combination of an electromagnet and a permanent magnet, and, much like loudspeakers, they con- vert electric energy into mechanical energy. A generator is the reverse: it converts mechanical energy into electric energy by moving a conductor through a magnetic field. Medicine : Hospitals use (nuclear) magnetic resonance tomography to spot problems in a patient’s organs without invasive surgery. Figure 8.1 illustrates the type of magnetism in the elementary materials of the periodic table. Only Fe, Ni, Co and Gd exhibit ferromagnetism. Most magnetic materials are there- fore alloys or oxides of these elements or contain them in another form. There is, however, 214

Upload: others

Post on 18-Jun-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

8 Magnetism

8.1 Introduction

The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally considered to be “non-magnetic”, which is a loose way of saying that theybecome magnetized only in the presence of an applied magnetic field (dia- and paramag-netism). We will see that in most cases these effects are very weak and the magnetizationis lost as soon as the external field is removed. Much more interesting (also from a tech-nological point of view) are those materials which not only have a large magnetization,but also retain it even after the removal of the external field. Such materials are calledpermanent magnets (ferromagnetism).The property that like (unlike) poles of permanent magnets repel (attract) each otherwas already known to the Ancient Greeks and Chinese over 2000 years ago. They usedthis knowledge for instance in compasses. Since then, the importance of magnets has risensteadily. They play an important role in many modern technologies:

• Recording media (e.g. hard disks): Data is recorded on a thin magnetic coating.The revolution in information technology owes as much to magnetic storage as toinformation processing with computer chips (i. e., the ubiquitous silicon chip).

• Credit, debit, and ATM cards: All of these cards have a magnetic strip on one side.

• TVs and computer monitors: Monitors contain a cathode ray tube that employsan electromagnet to guide electrons to the screen. Plasma screens and LCDs usedifferent technologies.

• Speakers and microphones: Most speakers employ a permanent magnet and a current-carrying coil to convert electric energy (the signal) into mechanical energy (move-ment that creates the sound).

• Electric motors and generators : Most electric motors rely upon a combination ofan electromagnet and a permanent magnet, and, much like loudspeakers, they con-vert electric energy into mechanical energy. A generator is the reverse: it convertsmechanical energy into electric energy by moving a conductor through a magneticfield.

• Medicine: Hospitals use (nuclear) magnetic resonance tomography to spot problemsin a patient’s organs without invasive surgery.

Figure 8.1 illustrates the type of magnetism in the elementary materials of the periodictable. Only Fe, Ni, Co and Gd exhibit ferromagnetism. Most magnetic materials are there-fore alloys or oxides of these elements or contain them in another form. There is, however,

214

Page 2: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.1: Magnetic type of the elements in the periodic table. For elements without acolor designation, magnetism is smaller or not explored.

recent and increased research into new magnetic materials such as plastic magnets (or-ganic polymers), molecular magnets (often based on transition metals) or molecule basedmagnets (in which magnetism arises from strongly localized s and p electrons).Electrodynamics gives us a first impression of magnetism. Magnetic fields act on movingelectric charges or, in other words, currents. Very roughly, one may understand the be-havior of materials in magnetic fields as arising from the presence of “moving charges”.Bound core and quasi-free valence electrons possess a spin, which in a very simplified clas-sical picture can be viewed as a rotating charge, i. e., a current. Bound electrons have anadditional orbital momentum which adds another source of current (again in a simplisticclassical picture). These “microscopic currents” react in two different ways to an appliedmagnetic field: First, according to Lenz’ law, a current is induced that creates a mag-netic field opposing the external magnetic field (diamagnetism). Second, the “individualmagnets” represented by the electron currents align with the external field and enhanceit (paramagnetism). If these effects happen for each “individual magnet” independently,they remain small. The corresponding magnetic properties of the material may then beunderstood from the individual behavior of the constituents, i. e., from the individualatoms (insulators, semiconductors) or from the ions and free electrons (conductors), cf.Sec. 8.3. The much stronger ferromagnetism, however, arises from a collective behaviorof the “individual magnets” in the material. In Sec. 8.4 we will first discuss the source forsuch an interaction before we move on to simple models treating either a possible couplingof localized moments or itinerant ferromagnetism.

8.2 Macroscopic Electrodynamics

Before we plunge into the microscopic sources of magnetism in solids, let us first recapa few definitions from macroscopic electrodynamics. In vacuum we have E and B as the

215

Page 3: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

electric field (unit: V/m) and the magnetic flux density (unit: Tesla = Vs/m2), respectively.Note that both are vector quantities, and in principle they are also functions of space andtime. Since we will only deal with constant, uniform fields in this lecture, this dependencewill be dropped throughout. Inside macroscopic media, both fields can be affected by thecharges and currents present in the material, yielding the two new net fields, D (electricdisplacement, unit: As/m2) and H (magnetic field, unit: A/m). For the formulation ofmacroscopic electrodynamics (the Maxwell equations in particular) we therefore need so-called constitutive relations between the external (applied) and internal (effective) fields:

D = D(E,B), H = H(E,B). (8.1)

In general, this functional dependence can be written in form of a multipole expansion. Inmost materials, however, already the first (dipole) contribution is sufficient. These dipoleterms are called P (el. polarization) and M (magnetization), and we can write

H =B

µ0−M+ . . . (8.2)

D = ϵ0 (E+P+ . . .) , (8.3)

where ϵ0 = 8.85 · 10−12 As/Vm and µ0 = 4π · 10−7 Vs/Am are the dielectric constant andpermeability of vacuum. P and M depend on the applied field, which we can formallywrite as a Taylor expansion in E and B. If the applied fields are not too strong, onecan truncate this series after the first term (linear response), i. e., the induced polariza-tion/magnetization is then simply proportional to the applied field. We will see below thatexternal magnetic fields we can generate at present in the laboratory are indeed weak com-pared to microscopic magnetic fields, i. e., the assumption of a magnetic linear responseis often well justified. As a side note, current high-intensity lasers may, however, bringus easily out of the linear response regime for the electric polarization. Correspondingphenomena are treated in the field of non-linear optics.For the magnetization, it is actually more convenient to expand in the internal field H

instead of in B. In the linear response regime, we thus obtain for the induced magnetizationand the electric polarization

M = χmag H ⇒ M = χmag H (8.4)

P = χel E ⇒ P = χel E . (8.5)

Here, χel/mag is the dimensionless electric/magnetic susceptibility tensor. In simple ma-terials (on which we will focus in this lecture), the linear response is often isotropic inspace and parallel to the applied field. The susceptibility tensors then reduce to a scalarform χmag/el.These equations are in fact often directly written as defining equations for the dimension-less susceptibility constants in solid state textbooks. It is important to remember, however,that we made the dipole approximation and restricted ourselves to isotropic media. Forthis case, the constitutive relations between external and internal field reduce to

H =1

µ0µrB (8.6)

D = ϵ0ϵr E, (8.7)

216

Page 4: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

material susceptibility χmag

vacuum 0H20 -8·10−6 diamagneticCu ∼ -10−5 diamagneticAl 2·10−5 paramagneticiron (depends on purity) ∼ 100 - 1000 ferromagneticµ-metal (nickle-iron alloy) ∼ 80,000 - 100,000 basically screens everything (used

in magnetic shielding)

Table 8.1: Magnetic susceptibility of some materials.

where ϵr = 1+χel is the relative dielectric constant, and µr = 1+χmag the relative perme-ability of the medium. For χmag < 0, the magnetization is anti-parallel to H. Accordingly,we have µr < 1 and |H| < | 1

µ0B|. The response of the medium reduces (or screens) the

external field. Such a system is called diamagnetic. For the opposite case (χmag > 0), themagnetization is aligned with H. Hence, we have µr > 1 and therefore |H| > | 1

µ0B|. The

external field is enhanced by the material, and we talk about paramagnetism.To connect to microscopic theories, we consider the energy. The process of screening orenhancing the external field will require work, i. e., the energy of the system is changed(magnetic energy). Assume therefore that we have a material of volume V which we bringinto a magnetic field B (which for simplicity we take to be along one dimension only).From electrodynamics, we know that the energy of this magnetic field is

Emag =1

2BHV ⇒

1

VdEmag =

1

2(B dH + H dB) . (8.8)

From Eq. (8.6), we find dB = µ0µrdH and therefore H dB = B dH, which leads to

1

VdEmag = B dH = (µ0H︸︷︷︸

B0

+µ0M) dH (8.9)

The first term µ0H = B0 describes the field in the vacuum, i. e., without the material,and thus just gives the field induced energy change if no material was present. Only thesecond term is material dependent and describes the energy change of the field in reactionto the material. Due to energy conservation, the energy change in the material itself istherefore

dEmagmaterial = −µ0MV dH. (8.10)

Recalling that the energy of a dipole with magnetic moment m in a magnetic field isE = −mB, we see that the approximations leading to Eq. (8.6) are equivalent to assumingthat the homogeneous solid is build up of a constant density of “molecular dipoles”, i. e.,the magnetization M is the (average) dipole moment density.By rearranging Eq. (8.10), we finally arrive at an expression that is a special case of afrequently employed alternative definition of the magnetization

M(H) = −1

µ0V

∂E(H)

∂H

∣∣∣∣S,V

, (8.11)

and in turn of the susceptibility

χmag(H) =∂M(H)

∂H

∣∣∣∣S,V

= −1

µ0V

∂2E(H)

∂H2

∣∣∣∣S,V

. (8.12)

217

Page 5: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

At finite temperatures, it is straightforward to generalize these definitions to

M(H,T ) = −1

µ0V

∂F (H,T )

∂H

∣∣∣∣S,V

χmag(H,T ) = −1

µ0V

∂2F (H,T )

∂H2

∣∣∣∣S,V

. (8.13)

While the derivation via macroscopic electrodynamics is most illustrative, we will see thatthese last two equations will be much more useful for the actual quantum-mechanicalcomputation of the magnetization and susceptibility of real systems. All we have to do isto derive an expression for the energy of the system as a function of the applied externalfield. The first and second derivatives with respect to the field then yield M and χmag,respectively.

8.3 Magnetism of Atoms and Free Electrons

We had already discussed that magnetism arises from the “microscopic currents” con-nected to the orbital and spin moments of electrons. Each electron therefore representsa “microscopic magnet”. But how do they couple in materials with a large number ofelectrons? Since any material is composed of atoms, it is reasonable to first reduce theproblem to that of the electronic coupling inside an atom before attempting to describethe coupling of “atomic magnets”. Since both orbital and spin momentum of all boundelectrons of an atom will contribute to its magnetic behavior, it will be useful to firstrecall how the total electronic angular momentum of an atom is determined (Sec. 8.3.1)before we turn to the effect of a magnetic field on the atomic Hamiltonian (Sec. 8.3.2).In metals, we have the additional presence of free conduction electrons. Their magneticproperties will be addressed in Sec. 8.3.3. Finally, we will discuss in Sec. 8.3.4 which as-pects of magnetism we can already understand just on the basis of these results on atomicmagnetism, i. e., in the limit of vanishing coupling between the “atomic magnets”.

8.3.1 Total Angular Momentum of Atoms

In the atomic shell model, the possible quantum states of electrons are labeled as nlmlms,where n is the principle quantum number, l the orbital quantum number, ml the orbitalmagnetic quantum number, and ms the spin magnetic quantum number. For any given nthat defines a so-called “shell”, the orbital quantum number l can take only integer valuesbetween 0 and (n− 1). For l = 0, 1, 2, 3 . . ., we generally use the letters s, p, d, f and soon. Within such an nl “subshell”, the orbital magnetic number ml may have the (2l + 1)integer values between −l and +l. The last quantum number, the spin magnetic quantumnumber ms, takes the values −1/2 and +1/2. For the example of the first two shells, thisleads to two 1s, two 2s, and six 2p states.When an atom has more than one electron, the Pauli exclusion principle dictates thateach quantum state specified by the set of four quantum numbers nlmlms can only beoccupied by one electron. In all but the heaviest ions, spin-orbit coupling1 is negligible. Inthis case, the Hamiltonian does not depend on spin or the orbital moment and thereforecommutes with these quantum numbers (Russel-Saunders coupling). The total orbital

1Spin-orbit (or LS) coupling is a relativistic effect that requires the solution of the Dirac equation.

218

Page 6: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

and spin angular momentum for a given subshell are then good quantum numbers andare given by

L =∑

ml and S =∑

ms. (8.14)

If a subshell is completely filled, it is easy to verify that L=S=0. The total electronicangular momentum of an atom, J = L+ S, is therefore determined by the partially filledsubshells. The occupation of quantum states in a partially filled shell is given by Hund’srules :

1. The states are occupied so that as many electrons as possible (within the limitationsof the Pauli exclusion principle) have their spins aligned parallel to one another, i. e.,so that the value of S is as large as possible.

2. When determined how the spins are assigned, then the electrons occupy states suchthat the value of L is a maximum.

3. The total angular momentum J is obtained by combining L and S as follows:

• if the subshell is less than half filled (i. e., if the number of electrons is < 2l+1),then J = L− S;

• if the subshell is more than half filled (i. e., if the number of electrons is > 2l+1),then J = L+ S;

• if the subshell is exactly half filled (i. e., if the number of electrons is = 2l+1),then L = 0, J = S.

The first two Hund’s rules are determined purely from electrostatic energy considera-tions, e. g., electrons with equal spins are farther away from each other on account of theexchange-hole. Only the third rule follows from spin-orbit coupling: Each quantum statein this partially filled subshell is called a multiplet. The term comes from the spectroscopycommunity and refers to multiple lines (peaks) that have the same origin (in this casethe same subshell nl). Without any interaction, e. g. electron-electron, spin-orbit, etc., allquantum states in this subshell would be energetically degenerate and the spectrum wouldshow only one peak at this energy. The interaction lifts the degeneracy and, in an ensembleof atoms, different atoms might be in different quantum states. The single peak then splitsinto multiple peaks in the spectrum. The notation for the ground-state multiplet obtainedby Hund’s rules is not simply denoted LSJ as one would have expected. Instead, for histor-ical reasons, the notation (2S+1)LJ is used. To make matters worse, the angular momentumsymbols are used for the total angular momentum (L = 0, 1, 2, 3, . . . = S, P,D, F, . . .). Therules are in fact easier to apply than their description might suggest at first glance. Table8.2 lists the filling and notation for the example of a d-shell.

8.3.2 General derivation of atomic susceptibilities

Having determined the total angular momentum of an atom, we now turn to the effect ofa uniform magnetic field

H =

⎝00H

⎠ (8.15)

219

Page 7: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

el. ml = 2 1 0 −1 −2 S L = |∑

ml| J Symbol1 ↓ 1/2 2 3/2 2D3/2

2 ↓ ↓ 1 3 2 3F2

3 ↓ ↓ ↓ 3/2 3 3/2 4F3/2

4 ↓ ↓ ↓ ↓ 2 2 0 5D0

5 ↓ ↓ ↓ ↓ ↓ 5/2 0 5/2 6S5/2

6 ↓↑ ↓ ↓ ↓ ↓ 2 2 4 5D4

7 ↓↑ ↓↑ ↓ ↓ ↓ 3/2 3 9/2 4F9/2

8 ↓↑ ↓↑ ↓↑ ↓ ↓ 1 3 4 3F4

9 ↓↑ ↓↑ ↓↑ ↓↑ ↓ 1/2 2 5/2 2D5/2

10 ↓↑ ↓↑ ↓↑ ↓↑ ↓↑ 0 0 0 1S0

Table 8.2: Ground states of ions with partially filled d-shells (l = 2), as constructed fromHund’s rules.

along the z-axis on the electronic Hamiltonian of an atom, He = T e + V e−ion + V e−e.Note that the focus on He means that we neglect the effect of H on the nuclear motionand spin. This is in general justified by the much greater mass of the nuclei rendering thenuclear contribution to the atomic magnetic moment very small, but it would of coursebe crucial if we were to address, e. g., nuclear magnetic resonance (NMR) experiments.Since V e−ion and V e−e are not affected by the magnetic field, we are left with the kineticenergy operator.From classical electrodynamics, we know that in the presence of a magnetic field, we haveto replace all momenta p (= i!∇) by the canonic momenta p → p + eA. For a uniformmagnetic field (along the z-axis), a suitable vector potential A is

A = −1

2(r× µ0H) , (8.16)

for which it is straightforward to verify that it fulfills the conditions H = (∇ ×A) and∇ ·A = 0. The total kinetic energy operator is then written

T e(H) =1

2m

k

[pk + eA]2 =1

2m

k

[pk −

e

2(rk × µ0H)

]2. (8.17)

Expanding the square leads to

T e(B) =∑

k

[p2k2m

+e

2mpk · (µ0H× rk) +

e2

8m(rk × µ0H)2

](8.18)

= T eo +

k

eµ0

2mH · (rk × pk) +

e2µ20H

2

8m

k

(x2k + y2k

),

where we have used the fact that p ·A = A · p if ∇ ·A = 0 and that the first term T eo

describes the kinetic energy of the system without any field. We simplify this equationfurther by introducing the dimensionless (!) total electronic orbital momentum operatorL = 1/!

∑k(rk ×pk) and the Bohr magneton µB = (e!/2m) = 0.579 · 10−4 eV/T, so that

the variation of the kinetic energy operator due to the magnetic field can be expressed as

∆T e(B) = T e(H)− T eo = µBµ0L ·H +

e2µ20

8mH2∑

k

(x2k + y2k

). (8.19)

220

Page 8: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

If this was the only effect of the magnetic field on He, one can, however, show that themagnetization in thermal equilibrium must always vanish (Bohr-van Leeuwen theorem).In simple terms, this is because the vector potential amounts to a simple shift in p whichintegrates out when a summation over all momenta is performed. The free energy is thenindependent of H and the magnetization is zero.In reality, however, the magnetization does not vanish. The problem is that we have usedthe wrong kinetic energy operator. We should have solved the relativistic Dirac equation

(V −mc2 −cσ · (p− eA)

cσ · (p− eA) −V −mc2

)(φχ

)= E

(φχ

). (8.20)

Here, φ =

(φ↑φ↓

)is an electron wave function and χ =

(χ↑χ↓

)a positron one, σ is

a vector of 2 × 2 Pauli spin matrices, c is the speed of light, and V is the potentialenergy. To solve and fully understand this equation, one would have to resort to quantumelectrodynamics (QED), which goes far beyond the scope of this lecture.We therefore heuristically introduce another momentum, the electron spin, that wouldemerge from the fully relativistic treatment mentioned above. The new interaction energyterm has the form

∆Hspin(H) = g0µBµ0S ·H , where S =∑

k

Sk . (8.21)

Here, S is the again dimensionless total spin angular momentum operator andg0 is theso called electronic g-factor (i. e., the g-factor for one electron spin). With the Diracequation one obtains g0 = 2, whereas a full QED approach yields

g0 = 2[1 +

α

2π+O(α2) + . . .

], where α =

e2

!c≈

1

137, (8.22)

= 2.0023 . . . .

For the sake of notational conciseness, we will use a value of g0 = 2 in this lecture.Eventually, the total field-dependent Hamiltonian is therefore

∆He(H) = ∆T e(H)+∆Hspin(H) = µ0µB (L+ g0S)·H +e2µ2

0

8mH2∑

k

(x2k + y2k

). (8.23)

We will see below that the energy shifts produced by Eq. (8.23) are generally quite smallon the scale of atomic excitation energies, even for the highest presently attainable lab-oratory field strengths. Perturbation theory can thus be used to calculate the changes inthe energies connected with this modified Hamiltonian. Remember that magnetizationand susceptibility were the first and second derivative of the field-dependent energy withrespect to the field, which is why we would like to write the energy as a function of H.Since we will require the second derivative later, terms to second order in H must beretained. Recalling second order perturbation theory, we can write the energy of the nthnon-degenerate level of the unperturbed Hamiltonian as

En → En +∆En(H); ∆En = ⟨n|∆He(H)|n⟩+∑

n′ =n

|⟨n|∆He(H)|n′⟩|2

En − E ′n

. (8.24)

221

Page 9: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Substituting Eq. (8.23) into the above and retaining terms to quadratic order in H, wearrive at the basic equation for the theory of the magnetic susceptibility of atoms:

∆En =

∆EPara

︷ ︸︸ ︷µ0µB⟨n| (L+ g0S) ·H|n⟩ +

∆ELL

︷ ︸︸ ︷e2µ2

0

8mH2⟨n|

k

(x2k + y2k

)|n⟩

+ µ20µ

2B

n′ =n

|⟨n|(L+ g0S) ·H|n′⟩|2

En − En′

︸ ︷︷ ︸∆EVV

. (8.25)

Since we are mostly interested in the atomic ground state |0⟩, we will now proceed to eval-uate the magnitude of the three terms contained in Eq. (8.25) and their correspondingsusceptibilities. This will also give us a feeling of the physics behind the different contri-butions. Beforehand, a few remarks –that will be justified during the discussion below–on the order of magnitude of these terms:

• If J = L = S = 0, we are dealing with a closed shell atom. Accordingly, alsoexcitation energies En−En′ will be large and the only non-vanishing term is ∆ELL.

• If J = 0, but L and S are not, we get an additional contribution from ∆EVV. Thecontribution will be the stronger the less energy an electronic excitation En − En′

costs.

• If J = 0, ∆EPara = 0 will completely dominate Eq. (8.25). In this case, however,the ground state is (2J + 1)-degenerate and we have to diagonalize the matrix in∆EPara.

8.3.2.1 ∆ELL: Larmor/Langevin Diamagnetism

To obtain an estimate for the magnitude of this term, we assume spherically symmetricwavefunctions (⟨0|

∑k(x

2k+y2k)|0⟩ ≈ 2/3 ⟨0|

∑k r

2k|0⟩). This allows us to approximate the

energy shift in the atomic ground state due to the second term as

∆Edia0 ≈

e2µ20H

2

12m

k

⟨0|r2k|0⟩ ∼e2µ2

0H2

12mZ r2atom , (8.26)

where Z is the total number of electrons in the atom (resulting from the sum over the kelectrons in the atom), and r2atom is the mean square atomic radius. If we take Z ∼ 30and r2atom of the order of Å2, we find ∆Edia

0 ∼ 10−9 eV even for fields of the order of Tesla.This contribution is therefore rather small. We make a similar observation from an orderof magnitude estimate for the susceptibility,

χmag,dia = −1

µ0V

∂2E0

∂H2= −

µ0e2Zr2atom6mV

∼ −10−4. (8.27)

With χmag,dia < 0, this term represents a diamagnetic contribution (the so called Larmoror Langevin diamagnetism), i. e., the associated magnetic moment screens the externalfield. We have thus identified the term which we qualitatively expected to arise out of theinduction currents initiated by the applied magnetic field.

222

Page 10: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

We will see below that this diamagnetic contribution will only determine the overallmagnetic susceptibility of the atom when the other two terms in Eq. (8.25) vanish. Thisis only the case for atoms with J|0⟩ = L|0⟩ = S|0⟩, i. e., for closed shell atoms or ions(e.g. the noble gas atoms). In this case, the ground state |0⟩ is indeed non-degeneratewhich justifies our use of Eq. (8.25) to calculate the energy level shift. As we recall fromchapter 6 on cohesion, the first excited state is much higher in energy for such closed shellsystems, so that at all but the highest temperatures, there is also a negligible probabilityof the atom being in any but its ground state in thermal equilibrium. This means that thediamagnetic susceptibility will be largely temperature independent (otherwise it wouldhave been necessary to use Eq. (8.13) for the derivation of χmag,dia).

8.3.2.2 ∆EVV: Van Vleck Paramagnetism

We first note that the evaluation of this term for a non-degenerate (J = 0) ground state |n⟩and for a field aligned along the z-direction yields

∆EVV = µ20µ

2BH

2∑

n

|⟨0|(Lz + g0Sz)|n⟩|2

E0 − En, (8.28)

in which Lz and Sz are the total angular projection operators along the z-axis for theorbital and spin angular momenta, respectively. By taking the second derivative, we obtainthe susceptibility

χmag,vleck =2µ0µ2

B

V

n

|⟨0|(Lz + g0Sz)|n⟩|2

En − E0. (8.29)

Note that we have compensated for the minus sign in the definition of the susceptibil-ity (8.12) by reversing the order in the denominator. Since the energy of any excitedstate will necessarily be higher than the ground state energy (En > E0), we thus getχmag,vleck > 0. The third term therefore represents a paramagnetic contribution (the socalled Van Vleck paramagnetism), which is connected to field-induced electronic transi-tions. If the electronic ground state is non-degenerate (and only for this case does theabove formula hold), it is precisely the normally quite large separation between electronicstates which makes this term similarly small as the diamagnetic one. Van Vleck param-agnetism therefore only plays an important role for atoms or ions that fulfill the followingconditions: First, the total angular momentum J has to vanish, otherwise the param-agnetic term ∆EPara discussed in the next section would completely overshadow ∆EVV.Second, L and S must not vanish individually, otherwise the contribution of ∆EVV wouldbecome vanishingly small due to the large separation between ground and excited statesin closed-shell configurations. A closer look at Hund’s rules (see Tab. 8.2 for the exampleof a d-shell) reveals that this is only the case for atoms wich are one electron short ofbeing half filled. The magnetic behavior of such atoms or ions is determined by a balancebetween Larmor diamagnetism and Van Vleck paramagnetism. Both terms are small andtend to cancel each other. The atomic lanthanide series is a notable exception: In this case,the energy spacing between the ground and excited states is small enough so that VanVleck paramagnetism indeed gives a notable contribution that is comparable in strengthto ∆EPara both for Eu (J = 0) and for Sm (J = 5/2)2, see Fig. 8.2. Here the effective

2Obviously, Eq. (8.28) is not directly applicable to Sm with J = 5/2. Rather, the six degenerate statesJz ∈ {−5/2,−3/2, · · · ,+5/2} need to be explicitly considered, as discussed in the next section.

223

Page 11: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.2: Van Vleck paramagnetism for Sm and Eu in the lanthanide atoms. From VanVleck’s nobel lecture.

magneton number µeff is plotted, which is defined in terms of the susceptibility

χ =Nµ2

effµ2B

3kBT.

8.3.2.3 ∆EPara : Paramagnetism

We again evaluate this term for a field aligned with the z-direction and get:

∆EPara = µ0µBH⟨n|Lz + g0Sz|n⟩ . (8.30)

As mentioned before, this first term does not vanish for any atom with J = 0 (which isthe case for most atoms) and will then completely dominate the magnetic behavior, sothat we can neglect the second and third term in Eq. (8.25) in this case. However, atomswith J = 0 have a (2J + 1)-fold degenerate ground state, which implies that the simpleform of Eq. (8.25) cannot be applied. Instead, we have to diagonalize the perturbationmatrix defined in Eq. (8.3.2.3) with respect to the α = 1, . . . , (2J +1) degenerate groundstates:

∆E0,α = µ0µBH

(2J+1)∑

α′=1

⟨0α|Lz + g0Sz|0α′⟩ = µ0µBH

(2J+1)∑

α′=1

Vα,α′ . (8.31)

224

Page 12: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

The respective energy shifts are obtained by diagonalizing the interaction matrix Vα,α′

within the ground-state subspace. This is a standard problem in atomic physics (see e.g.Ashcroft/Mermin) and one finds that the basis diagonalizing this matrix are the statesdefined by J and Jz,

⟨JLS, Jz|Lz + g0Sz|JLS, J ′z⟩ = g(JLS)Jz δJz ,Jz′ . (8.32)

where JLS are the quantum numbers defining the atomic ground state |0⟩ in the shellmodel and g(JLS) is the so-called Landé g-factor. Setting g0 ≈ 2, this factor is obtainedas

g(JLS) =3

2+

1

2

[S(S + 1)− L(L+ 1)

J(J + 1)

]. (8.33)

If we insert Eq. (8.32) into Eq. (8.31), we obtain

∆EJLS,Jz = g(JLS)µ0µBJz H , (8.34)

i. e., the magnetic field splits the degenerate ground state into (2J + 1) equidistant levelsseparated by g(JLS)µ0µBH, i.e.,

g(JLS)µ0µBH Jz with Jz ∈ {−J, J − 1, . . . , J + 1, J} . (8.35)

This is the same effect, as the magnetic field would have on a magnetic dipole withmagnetic moment

matom = −g(JLS)µ0µB J. (8.36)

Therefore, the first term in Eq. (8.25) can be interpreted as the expected paramagneticcontribution due to the alignment of a “microscopic magnet”. This is of course only truewhen we have unpaired electrons in partially filled shells (J = 0).Before we proceed to derive the paramagnetic susceptibility that arises from this contribu-tion, we note that this identification of matom via Eq. (8.25) serves nicely to illustrate thedifference between phenomenological and so-called first-principles theories. We have seenthat the simple Hamiltonian Hspin = −matom ·H would equally well describe the splittingof the (2J + 1) ground state levels. If we had only known this splitting, e.g. from experi-ment (Zeeman effect), we could have simply used Hspin as a phenomenological, so-called“spin Hamiltonian”. By fitting to the experimental splitting, we could even have obtainedthe magnitude of the magnetic moment for individual atoms (which due to their differenttotal angular momenta is of course different for different species). Examining the fittedmagnitudes, we might even have discovered that the splitting is connected to the total an-gular momentum. The virtue of the first-principles derivation used in this lecture, on theother hand, is that this prefactor is directly obtained, without any fitting to experimentaldata. This allows us not only to unambiguously check our theory by comparison with theexperimental data (which is not possible in the phenomenological theory, where the fittingcan often easily cover up for deficiencies in the assumed ansatz). We also directly obtainthe dependence on the total angular momentum and can even predict the properties ofatoms which have not (yet) been measured. Admittedly, in many cases first-principlestheories are harder to derive. Later we will see more examples of “spin Hamiltonians”which describe the observed magnetic behavior of the material very well, but for whichan all-encompassing first-principles theory is still lacking.

225

Page 13: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

0 1 2 3 4x

0

0.2

0.4

0.6

0.8

1B J(x)

J=1/2

1

3/22

5/2

Figure 8.3: Plot of the Brillouin function BJ(x) for various values of the spin J (the totalangular momentum).

8.3.2.4 Paramagnetic Susceptibility: Curie’s Law

The separation of the (2J + 1) ground state levels of a paramagnetic atom in a magneticfield is g(JLS)µ0µBH. Recalling that µB = 0.579 · 10−4 eV/T, we see that the splitting isonly of the order of 10−4 eV, even for fields of the order of Tesla. This is small comparedto kBT for realistic temperatures. Accordingly, more than one level will be populated atfinite temperatures, so that we have to use statistical mechanics to evaluate the suscepti-bility, i.e., the derivatives of the free energy as defined in Eq. (8.13). The Helmholtz freeenergy

F = −kBT lnZ (8.37)

is given in terms of the partition function Z with

Z =∑

n

e− En

kBT . (8.38)

Here, n is Jz = −J, . . . , J and En is our energy spacing g(JLS)µ0µBHJz. By defining

η =g(JLS)µ0µBH

kBT

so that η is the fraction of the magnetic splitting versus thermal energy, it becomes evidentthat the partition function is a geometric sum that evaluates to:

Z =J∑

Jz=−J

e−ηJz = eηJ + eη(J−1) + · · ·+ e−η(J−1) + e−ηJ

=e−η(J+1/2) − eη(J+1/2)

e−η/2 − eη/2=

sinh [(J + 1/2)η]

sinh [η/2]. (8.39)

226

Page 14: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

With this, one finds for the magnetization

M(T ) = −1

µ0V

∂F

∂H= −

1

µ0V

∂(−kBT lnZ)

∂H(8.40)

=kBT

µ0V

∂H[ln (sinh[(J + 1/2)η])− ln (sinh[η/2])] =

g(JLS)µBJ

VBJ(η) .

In the last step, we have defined the so called Brillouin function

BJ(η) =1

J

{(J +

1

2) coth

[(J +

1

2)η/J

]−

1

2coth

[ η2J

]}. (8.41)

As shown in Fig. 8.3, BJ → 1 for η ≫ 1, i.e., in the limit of a strong field and/orlow temperatures. In this case, Eq. (8.40) simply tells us that all atomic magnets withmomentum matom defined in Eq. (8.36) have aligned to the external field and that themagnetization reaches its maximum value of M = matom/V , i. e., the magnetic momentdensity.We had, however, discussed above, that η will be much smaller than unity for normal fieldstrengths. At all but the lowest temperatures, the other limit for the Brillouin function,i. e., BJ(η → 0), will therefore be much more relevant. This small-η expansion yields

BJ(η ≪ 1) ≈J + 1

3η + O(η3) . (8.42)

In this limit, we finally obtain for the paramagnetic susceptibility

χmag,para(T ) =∂M

∂H=

µoµ2Bg(JLS)

2J(J + 1)

3V kB

1

T=

C

T. (8.43)

With χmag,para > 0, we have now confirmed our previous hypothesis that the first term inEq. (8.25) gives a paramagnetic contribution. The inverse dependence of the susceptibilityon temperature is known as Curie’s law, χmag,para = CCurie/T with CCurie being the Curieconstant. Such a dependence is characteristic for a system with “molecular magnets” thatfavor to align with the applied field and are being flipped by thermal fluctuations. Again,we see that the first-principles derivation not only reveals the range of validity for thisempirical law (remember that it only holds in the small η-limit), but also provides theproportionality constant in terms of fundamental properties of the system.To make a quick size estimate, we take the volume of atomic order (V ∼Å3) to findχmag,para ∼ 10−2 at room temperature. The paramagnetic contribution is thus orders ofmagnitude larger than the diamagnetic or the Van Vleck one, even at room temperature.Nevertheless, with χmag,para ≪ 1 it is still small compared to electric susceptibilities, whichare of the order of unity. We will discuss the consequences of this in section 8.3.4, butbefore we have to evaluate a last remaining, additional source of magnetic behavior inmetals, i. e., the free electrons.

8.3.3 Susceptibility of the Free Electron Gas

Having determined the magnetic properties of electrons bound to ions, it is valuable toalso consider the opposite extreme and examine their properties as they move nearlyfreely in a metal. There are again essentially two major terms in the response of free

227

Page 15: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

fermions to an external magnetic field. One results from the “alignment” of spins (whichis a rough way of visualizing the quantum mechanical action of the field on the spins) andit is called Pauli paramagnetism. The other arises from the orbital moments created byinduced circular motions of the electrons. It thus tends to screen the exterior field and isknown as Landau diamagnetism.Let us first try to analyze the origin of Pauli paramagnetism and examine its order ofmagnitude like we have also done for all atomic magnetic effects. For this we consideragain the free electron gas, i. e., N non-interacting electrons in a volume V . In chapter 2,we had already derived the density of states (DOS) per volume

N(ϵ) =V

2π2

(2me

!2

)3/2√ϵ . (8.44)

In the electronic ground state, all states are filled following the Fermi-distribution

N

V=

∫ ∞

0

N(ϵ) f(ϵ, T ) dϵ. (8.45)

The electron density N/V uniquely characterizes our electron gas. For sufficiently lowtemperatures, the Fermi-distribution can be approximated by a step function so that allstates up to the Fermi level are occupied. This Fermi level is conveniently expressed interms of the Wigner-Seitz radius rs that we had already defined in chapter 6

ϵF =50.1 eV(

rsaB

)2 . (8.46)

The DOS at the Fermi level can then also be expressed in terms of rs

N(ϵF) =

(1

20.7 eV

)(rsaB

)−1

. (8.47)

Up to now we have not explicitly considered the spin degree of freedom which leads to thedouble occupation of each electronic state with a spin up and a spin down electron. Suchan explicit consideration becomes necessary, however, when we now include an externalfield. This is accomplished by defining spin up and spin down DOS’s, N↑ and N↓, inanalogy to the spin-unresolved case. In the absence of a magnetic field, the ground stateof the free electron gas has an equal number of spin-up and spin-down electrons, and thedegenerate spin-DOS are therefore simply half-copies of the original total DOS defined inEq. (8.44)

N↑(ϵ) = N↓(ϵ) =1

2N(ϵ) , (8.48)

cf. the graphical representation in Fig. 8.4a.An external field µ0H will now interact differently with spin up and spin down states.Since the electron gas is fully isotropic, we can always choose the spin axis that defines upand down spins to go along the direction of the applied field, and thus reduce the problemto one dimension. If we focus only on the interaction of the spin with the field and neglectthe orbital response for the moment, the effect boils down to an equivalent of Eq. (8.21),i. e., we have

∆Hspin(H) = µ0µBg0H · s = +µ0µBH (for spin up) (8.49)= −µ0µBH (for spin down) , (8.50)

228

Page 16: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.4: Free electron density of states and the effect of an external field: N↑ correspondsto the number of electrons with spins aligned parallel and N↓ antiparallel to the z-axis.(a) No magnetic field, (b) a magnetic field, µ0H0 is applied along the z-direction, (c) asa result of (b) some electrons with spins antiparallel to the field (shaded region on left)change their spin and move into available lower energy states with spins aligned parallelto the field.

229

Page 17: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

where we have again approximated the electronic g-factor g0 by 2. The effect is thereforesimply to shift spin up electrons relative to spin down electrons. This can be expressedvia the spin DOS’s

N↑(ϵ) =1

2N(ϵ− µ0µBH) (8.51)

N↓(ϵ) =1

2N(ϵ+ µ0µBH) , (8.52)

or again graphically by shifting the two parabolas with respect to each other as shownin Fig. 8.4b. If we now fill the electronic states according to the Fermi-distribution, adifferent number of up and down electrons results and our electron gas exhibits a netmagnetization due to the applied field. Since the states aligned parallel to the field arefavored, the magnetization will enhance the exterior field and a paramagnetic contributionresults.For T ≈ 0, both parabolas are essentially filled up to a sharp energy, and the net mag-netization is simply given by the dark grey shaded region in Fig. 8.4c. Even for fields ofthe order of Tesla, the energetic shift of the states is µBµ0H ∼ 10−4 eV, i. e., small onthe scale of electronic energies. The shaded region representing our net magnetization cantherefore be approximated by a simple rectangle of height µ0µBH and width 1

2N(ϵF), i. e.,the Fermi level DOS of the unperturbed system without applied field. We then have

N↑ =N

2+

1

2N(ϵF)µ0µBH (8.53)

up electrons per volume and

N↓ =N

2−

1

2N(ϵF)µ0µBH (8.54)

down electrons per volume. Remember that the magnetization is equal to the dipolemagnetic density and can therefore be written as dipole moment per electron times DOSin such a uniform system. Accordingly, each electron contributes a magnetic moment ofµB to the magnetization and we have

M = (N↑ −N↓)µB = N(ϵF)µ0µ2BH . (8.55)

This then yields the susceptibility

χmag,Pauli =∂M

∂H= N(ϵF)µ0µ

2B . (8.56)

Using Eq. (8.47) for the Fermi level DOS of the free electron gas, this can be rewritten asfunction of the Wigner-Seitz radius rs

χmag,Pauli = 10−6

(2.59

rs/aB

). (8.57)

Recalling that the Wigner-Seitz radius of most metals is of the order of 3-5 aB, we find thatthe Pauli paramagnetic contribution (χmag,Pauli > 0) of a free electron gas is small. In fact,it is as small as the diamagnetic susceptibility of atoms and therefore much smaller thantheir paramagnetic susceptibility. For magnetic atoms in a material, thermal disorder at

230

Page 18: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

finite temperatures prevents their complete alignment to an external field and thereforekeeps the magnetization small. For an electron gas, on the other hand, it is the Pauliexclusion principle that opposes such an alignment by forcing the electrons to occupyenergetically higher lying states when aligned. For the same reason, Pauli paramagnetismdoes not show the linear temperature dependence we observed in Curie’s law in Eq. (8.43).The characteristic temperature for the electron gas is the Fermi temperature. One couldtherefore cast the Pauli susceptibility into Curie form, but with a fixed temperature oforder TF playing the role of T . Since TF ≫ T , the Pauli susceptibility is then hundreds oftimes smaller and almost always completely negligible.Turning to Landau diamagnetism, we note that its calculation would require solving thefull quantum mechanical free electron problem along similar lines as done in the lastsection for atoms. The field-dependent Hamilton operator would then also produce aterm that screens the applied field. Taking the second derivative of the resulting totalenergy with respect to H, one would obtain for the diamagnetic susceptibility

χmag,Landau = −1

3χmag,Pauli , (8.58)

i. e., another term that is vanishingly small, as is the paramagnetic response of the freeelectron gas. Because this term is so small and the derivation does not yield new insightcompared to the one already undertaken for the atomic case, we refer to the book by M.P.Marder, Condensed Matter Physics (Wiley, 2000) for a proper derivation of this result.

8.3.4 Atomic magnetism in solids

So far, we have discussed the magnetic behavior of materials from tight-binding viewpointby assuming that a material is “just” an ensemble of indipendent atoms or ions and elec-trons. That is why we first addressed the magnetic properties of isolated atoms and ions.As an additional source of magnetism in solids, we looked at free (delocalized or itiner-ant) electrons. In both cases we found two major sources of magnetism: a paramagneticone resulting from the alignment of existing “microscopic magnets” (either total angularmomentum in the case of atoms, or spin in case of free electrons) and a diamagnetic onearising from induction currents that try to screen the external field.

• Atoms (bound electrons):

- Paramagnetism χmag,para ≈ 1/T ∼ 10−2 (RT) ∆Epara0 ∼ 10−4 eV

- Van Vleck paramagnetismχmag,VV ≈ const. ∼ −10−4 ∆EVV0 ∼ 10−9 eV

- Larmor diamagnetism χmag,dia ≈ const. ∼ −10−4 ∆Edia0 ∼ 10−9 eV

• Free electrons:

- Pauli paramagnetism χmag,Pauli ≈ const. ∼ 10−6 ∆EPauli0 ∼ 10−4 eV

- Landau diamagnetism χmag,Landau ≈ const. ∼ −10−6 ∆ELandau0 ∼ 10−4 eV

If the coupling between the different sources of magnetism is small, the magnetic behaviorof the material would simply be a superposition of all relevant terms. For insulators thiswould imply, for example, that the magnetic moment of each atom/ion does not change

231

Page 19: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.5: (a) In rare earth metal atoms the incomplete 4f electronic subshell is locatedinside the 5s and 5p subshells, so that the 4f electrons are not strongly affected byneighboring atoms. (b) In transition metal atoms, e.g. the iron group, the 3d electrons arethe outermost electrons and therefore interact strongly with other nearby atoms.

appreciably when transfered to the material and the total susceptibility would be just thesum of all atomic susceptibilities. And this is indeed what one primarily finds when goingthrough the periodic table:Insulators: When J = 0, the paramagnetic contribution vanishes. If in addition L =S = 0 (closed-shell), the response is purely diamagnetic. This is the case for noble gassolids, simple ionic crystals like the alkali halides (recall from the discussion on cohesionthat the latter can be viewed as closed-shell systems!), many molecules (e.g. H2O), andthus also many molecular crystals. Otherwise, i.e., if J = 0 but not necessarily L and S,the response is a balance between Van Vleck paramagnetism and Larmor diamagnetism.In all cases, the effects are small and the results from the theoretical derivation presentedin the previous section are in excellent quantitative agreement with experiment.For J = 0, the response is dominated by the paramagnetic term. This is the case, whenthe material contains rare earth (RE) or transition metal (TM) atoms (with partiallyfilled f and d shells, respectively). These systems indeed obey Curie’s law, i. e., theyexhibit susceptibilities that scale inversely with temperature. For the rare earths, eventhe magnitude of the measured Curie constant corresponds very well to the theoreticalprediction given in Eq. (8.43). For TM atoms in insulating solids, howver, the measuredCurie constants can only be understood if one assumes L = 0, while S is still given byHund’s rules. This effect is referred to as quenching of the orbital angular momentum andis caused by a general phenomenon known as crystal field splitting. As illustrated by Fig.8.5, the d shells of a TM atom belong to the outermost valence shells and are particularlysensitive to the environment. The crystal field lifts the five-fold degeneracy of the TMd-states, as shown for an octahedral environment in Fig. 8.6a). Hund’s rules are then incompetition with the energy scale of the d-state splitting. If the energy separation is large,only the three lowest states can be filled and Hund’s rules give a low spin configuration

232

Page 20: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.6: Effect of octahedral environment on transition metal ion with 5 electrons (d5):a) the degenerate levels split into two groups, b) for large energy separation a low spinand for small energy separation a high spin configuration c) might be favorable.

d-count # unpaired electrons exampleshigh spin low spin

d4 4 2 Cr2+, Mn3+

d5 5 1 Fe3+, Mn2+

d6 4 0 Fe2+, Co3+

d7 3 1 Co2+

Table 8.3: High and low spin octahedral transition metal complexes.

with weak diamagnetism (Fig. 8.6b). If the energy separation is small, Hund’s rules canstill be applied to all five states and a high spin situation with strong paramagnetismemerges (Fig. 8.6c). Please note that the crystal field splitting does not affect RE atoms,since the partially filled f -orbitals of RE atoms are located deep inside the filled s and pshells and are therefore not significantly affected by the crystalline environment. Even inthe case of TM atoms, the important thing to notice is that although the solid perturbsthe magnitude of the “microscopic magnet” connected with the partially filled shell, allindividual moments inside the material are still virtually decoupled from each other (i. e.,the magnetic moment in the material is that of an atom that has simply adapted to anew symmetry situation). This is why Curie’s law is still valid, but the magnetic momentbecomes an effective magnetic moment that depends on the crystalline environment. Ta-ble 8.3 summarizes the configurations for some transition metal ions in an octahedralenvironment.Semiconductors: Covalent materials only have partially filled shells, so they could be ex-pected to have a finite magnetic moment. However, covalent bonds form through a pair of

233

Page 21: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

electrons with opposite spin, and hence the net orbital angular momentum is zero. There-fore, covalent materials like silicon exhibit only a vanishingly small diamagnetic response,as long as they are not doped. Dilute magnetic semiconductors, i. e., semiconductors dopedwith transition metal atoms, hold the promise to combine the powerful world of semicon-ductors with magnetism. However, questions like how pronounced the magnetism is andif room temperature magnetism can be achieved are still heavily debated and subject ofactive research.Metals: In metals, the delocalized electrons add a new contribution to the magnetic be-havior. In simple metals like the alkalis, this new contribution is in fact the only remainingone. As discussed in the chapter on cohesion, such metals can be viewed as closed-shellion cores and a free electron glue formed from the valence electrons. The closed-shell ioncores have J = L = S = 0 and exhibit therefore only a negligible diamagnetic response.The magnetic behavior is then dominated by the Pauli paramagnetism of the conductionelectrons. Measured susceptibilities are indeed in the order of magnitude that can be ex-pected from Eq. (8.57). Quantitative differences arise from exchange-correlation effectsthat were not treated in our free electron model. We will come back to metals below.

8.4 Magnetic Order: Permanent Magnets

The theory we have developed up to now relies on the assumption that the magneticproperties of material can be described entirely by summing up the contributions that stemfrom the individual atoms and free electrons. In all cases, the paramagnetic or diamagneticresponse is very small, at least when compared to electric susceptibilities, which are ofthe order of unity. For this exact reason, typically only electric effects are discussed whenanalysing the interactions of materials with electromagnetic fields. At this stage, we couldtherefore close the chapter on magnetism of solids as something really unimportant, ifit was not for a few exceptions: Some elemental and compound materials (formed of orcontaining transition metal or rare earth atoms) exhibit a completely different magneticbehavior: enormous susceptibilities and a magnetization that does not vanish when theexternal field is removed. These so called ferromagnetic effects are not captured by thehitherto developed theory, since they originate from a strong coupling of the individualmagnetic moments. This is what we will consider next.

8.4.1 Ferro-, Antiferro- and Ferrimagnetism

In our previously discussed, simple theory of paramagnetism the individual magnetic mo-ments (ionic shells of non-zero angular momentum in insulators or the conduction electronsin simple metals) do not interact with one another. In the absence of an external field, theindividual magnetic moments are then thermally disordered at any finite temperature,i. e., they point in random directions yielding a zero net moment for the solid as a whole,cf. Fig. 8.7a. In such cases, an alignment can only be caused by applying an externalfield, which leads to an ordering of all magnetic moments at sufficiently low tempera-tures, i.e., as long as the field is strong enough to overcome the thermodynamic trendtowards disorder. Schematically, this is shown in Fig. 8.7b. However, a similar alignmentcould also be obtained by coupling the different magnetic moments, i. e., by an interac-tion that favors a parallel alignment for example. Already quite short-ranged interactions,

234

Page 22: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.7: Schematic illustration of the distribution of directions of the local magneticmoments connected to “individual magnets” in a material. (a) Random thermal disorderin a paramagnetic solid with insignificant magnetic interactions, (b) complete alignmenteither in a paramagnetic solid due to a strong external field or in a ferromagnetic solidbelow its critical temperature as a result of magnetic interactions, and (c) example ofantiferromagnetic ordering below the critical temperature.

e.g., nearest-neighbor interactions, can potentially lead to an ordered structure as the oneshown in Fig. 8.7b. In literature, interactions that introduce a magnetic ordering are oftencalled magnetic interactions. This does however not imply that the fundamental sourceof the interaction is magnetic (e.g., a magnetic dipole-dipole interaction, which in fact isnot the reason for the ordering as we will see below).Materials that exhibit an ordered magnetic structure in the absence of an applied externalfield are called ferromagnets (or permanent magnets) and their resulting magnetic momentis known as spontaneous magnetization Ms. The complexity of the possible magneticallyordered states exceeds the simple parallel alignment case shown in Fig. 8.7b by far; someadditional examples are shown in Fig. 8.8a. Additionally, Fig. 8.7c and Fig. 8.8b showanother common case: Although, a microscopic ordering is present, the individual localmoments sum to zero, so that no spontaneous magnetization is present. Such magneticallyordered states with Ms = 0 are called antiferromagnetic. If a spontaneous magnetizationMs = 0 occurs from the ordering of magnetic moments of different magnitude that are notcompletely aligned with Ms (not all local moments have a positive component along thedirection of spontaneous magnetization), one talks about ferrimagnets. Fig. 8.8 shows afew examples of possible ordered ferrimagnetic structures. Please note that the complexityof possible magnetic structures is so large that it cannot be covered by a simple sketch.Indeed, some of them do not rigorously fall into any of the three categories discussed sofar. Those structures are well beyond the scope of this lecture and will not be coveredhere.

8.4.2 Interaction versus Thermal Disorder: Curie-Weiss Law

Even in permanent magnets, the magnetic order does not usually prevail at all temper-atures. Above a critical temperature TC, the thermal energy can overcome the orderinginduced by the interactions. Then, the magnetic order vanishes and the material (mostoften) behaves like a simple paramagnet. In ferromagnets, TC is known as the Curietemperature, while in antiferromagnets it is often called as Néel temperature (sometimesdenoted TN). Note that TC may depend strongly on the applied field (both in strength

235

Page 23: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.8: Examples for some typical magnetic ordering that can be found in a simplelinear array of spins: (a) ferromagnetic, (b) antiferromagnetic, and (c) ferrimagnetic.

236

Page 24: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.9: Typical temperature dependence of the magnetization Ms, of the specific heatcV , and of the zero-field susceptibility χo of a ferromagnet. The temperature scale isnormalized to the critical (Curie) temperature TC.

and direction!). If the external field is parallel to the direction of spontaneous magnetiza-tion, TC typically increases with increasing external field since both ordering tendenciessupport each other. For other field directions, a range of complex phenomena can arise asa result of the competition between both ordering tendencies.At first, we are more interested in the ferromagnetic properties of materials in the absenceof external fields, i. e., in the existence of a finite spontaneous magnetization. The gradualloss of order with increasing temperature is reflected in the continuous drop of Ms(T )illustrated in Fig. 8.9. Just below TC, a power law dependence is typically observed

Ms(T ) ∼ (TC − T )β (for T ↗ TC) , (8.59)

with a critical exponent β somewhere around 1/3. Coming from the high temperatureside, the onset of ordering also appears in the zero-field susceptibility, which is found todiverge as T approaches TC

χ0(T ) = χ(T )|B=0 ∼ (T − TC)−γ (for T ↘ TC) , (8.60)

with γ around 4/3. This behavior already indicates that the material experiences dra-matic changes at the critical temperature, which is also reflected by a divergence in otherfundamental quantities like the zero-field specific heat

cV (T )|B=0 ∼ (T − TC)−α (for T → TC) , (8.61)

with α around 0.1. The divergence is connected to the onset of long-range order (align-ment), which sets in more or less suddenly at the critical temperature resulting in a secondorder phase transition. The actual transition is difficult to describe theoretically and the-ories are therefore often judged by how well (or badly) they reproduce the experimentallymeasured critical exponents α, β and γ.Above the critical temperature, the properties of magnetic materials often become “morenormal”. In particular, for T ≫ TC, one usually finds a simple paramagnetic behavior,

237

Page 25: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

ms (in µB) matom (in µB) TC (in K) ΘC (in K)Fe 2.2 6 (4) 1043 1100Co 1.7 6 (3) 1394 1415Ni 0.6 5 (2) 628 650Eu 7.1 7 289 108Gd 8.0 8 302 289Dy 10.6 10 85 157

Table 8.4: Magnetic quantities of some elemental ferromagnets: The saturation magneti-zation at T = 0K is given in form of the average magnetic moment ms per atom and thecritical temperature TC and the Curie-Weiss temperature ΘC are given in K. For compar-ison, also the magnetic moment matom of the corresponding isolated atoms is listed (thevalues in brackets are for the case of orbital angular momentum quenching).

with a susceptibility that obeys the so-called Curie-Weiss law, cf. Fig. 8.9,

χ0(T ) ∼ (T −ΘC)−1 (for T ≫ TC) . (8.62)

ΘC is called the Curie-Weiss temperature. Since γ in Eq. (8.60) is almost never exactlyequal to 1, ΘC does not necessarily coincide with the Curie temperature TC. This can,for example, be seen in Table 8.4, where also the saturated values for the spontaneousmagnetization at T → 0K are listed.In theoretical studies, one typically converts this measured Ms(T → 0K) directly into anaverage magnetic moment ms per atom (in units of µB) by simply dividing by the atomicdensity in the material. In this way, we can compare directly with the corresponding valuesfor the isolated atoms, obtained as matom = g(JLS)µBJ , cf. Eq. (8.36). As apparent fromTable 8.4, the two values agree very well for the rare earth metals, but differ substantiallyfor the transition metals. Recalling the discussion on the quenching of the orbital angularmomentum of TM atoms immersed in insulating materials in section 8.3.4, we could onceagain try to ascribe this difference to the symmetry lowering of interacting d orbitalsin the material. Yet, even using L = 0 (and thus J = S, g(JLS) = 2), the agreementdoes not become much better, cf. Table 8.4. Accordingly, one can summarize that inthe RE ferromagnets the saturation magnetization appears to be a result of the perfectalignment of atomic magnetic moments that are not strongly affected by the presence ofthe surrounding material. Conversely, the TM ferromagnets (Fe, Co, Ni) exhibit a muchmore complicated magnetic structure that is not well described by the magnetic behaviorof the individual atoms.As a first summary of our discussion on ferromagnetism (or ordered magnetic states ingeneral), we are therefore led to conclude that it must arise out of an (yet unspecified)interaction between magnetic moments. This interaction produces an alignment of themagnetic moments and subsequently a large net magnetization that prevails even in theabsence of an external field. This means we are certainly outside the linear response regimethat has been used as fundamental basis for developing the dia- and paramagnetic theoriesin the first sections of this chapter. With respect to the alignment, the interaction seemsto have the same qualitative effect as an applied field acting on a paramagnet, whichexplains the similarity between the Curie-Weiss and the Curie law of paramagnetic solids:Interaction and external field favor alignment, but are opposed by thermal disorder. Far

238

Page 26: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

below the critical temperature the alignment is perfect and far above TC thermal disorder ispredominant. In this case, the competition between alignment with respect to an externalfield and disorder is equivalent to the situation in a paramagnet (just the origin of thetemperature scale has shifted).The interacting magnetic moments in ferromagnets can either stem from delocalized elec-trons (Pauli paramagnetism) or stem from partially filled atomic shells (Paramagnetism).This makes it plausible why ferromagnetism is a phenomenon specific to only some met-als, and in particular metals with a high magnetic moment that arises from partially filledd or f shells (i. e., transition metals, rare earths, as well as their compounds). However,we cannot yet explain why only some and not all TMs and REs exhibit ferromagneticbehavior. The comparison between the measured values of the spontaneous magnetizationand the atomic magnetic moments suggests that ferromagnetism in RE metals can be un-derstood as the coupling of localized magnetic moments (due to the inert partially filledf shells). The situation is more complicated for TM ferromagnets, where the picture of asimple coupling between atomic-like moments fails. We will see below that this so-calleditinerant ferromagnetism arises out of a subtle mixture of delocalized s electrons and themore localized, but nevertheless not inert d orbitals.

8.4.3 Phenomenological Theories of Ferromagnetism

After this first overview of phenomena arising out of magnetic order, let us now try toestablish a proper theoretical understanding. Unfortunately, the theory of ferromagnetismis one of the less well developed of the fundamental theories in solid state physics (a bitbetter for the localized ferromagnetism of the REs, worse for the itinerant ferromagnetismof the TMs). The complex interplay of single particle and many-body effects, as well ascollective effects and strong local coupling makes it difficult to break the topic down intosimple models appropriate for an introductory level lecture. We will therefore not be ableto present and discuss a full-blown ab initio theory like in the preceding chapters. Insteadwe will first consider simple phenomenological theories and refine them later, when needed.This will enable us to address some of the fundamental questions that are not amenableto phenomenological theories (like the source of the magnetic interaction).

8.4.3.1 Molecular (mean) field theory

The simplest phenomenological theory of ferromagnetism is the molecular field theory dueto Weiss (1906). We had seen in the preceding section that the effect of the interactionbetween the discrete magnetic moments in a material is very similar to that of an appliedexternal field (leading to alignment). For each magnetic moment, the net effect of theinteraction with all other moments can therefore be thought of as a kind of internal fieldcreated by all other moments (and the magnetic moment itself). In this way we averageover all interactions and condenses them into one effective field, i. e., molecular field theoryis a typical representative of a mean field theory. Since this effective internal, or so-calledmolecular field corresponds to an average over all interactions, it is reasonable to assumethat it will scale with M , i. e., the overall magnetization (density of magnetic moments).Without specifying the microscopic origin of the magnetic interaction and the resultingfield, the ansatz of Weiss was to simply postulate a linear scaling with M

Hmol = µ0λM , (8.63)

239

Page 27: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.10: Graphical solution of equation 8.66.

where λ is the so-called molecular field constant. This leads to an effective field

Heff = H +Hmol = H + µ0λM , (8.64)

For simplicity, we will consider only the case of an isotropic material, so that the internaland the external point in the same direction, allowing us to write all formulae as scalarrelations.Recalling the magnetization of paramagnetic spins (Eq. 8.40),

M0(T ) =g(JLS)µBJ

VBJ(η) with η ∼

H

T, (8.65)

we obtain

M(T ) = M0

(Heff

T

)→M0

(λM

T

)(8.66)

when the external field is switched off (H=0). We now ask the question if such a field canexist without an external field? Since M appears on the left and on the right hand sideof the equation we perform a graphic solution. We set x = λ

TM(T ), which results in

M0(x) =T

λx . (8.67)

The graphic solution is show in Fig. 8.10 and tells us that the slop of M0(0K) has to belarger than T

λ . If this condition is fulfilled a ferromagnetic solution of our paramagneticmean field model is obtained, although at this point we have no microscopic understandingof λ.

240

Page 28: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

For the mean-field susceptibility we then obtain with H(T ) = M0(Heff

T )

χ =∂M0

∂H=∂M0

∂Heff

∂Heff

∂H=

C

T

(1 +

∂M

∂H

)=

C

T(1 + λχ) (8.68)

where we have used that for our paramagnet ∂M0∂Heff

is just Curie’s law. Solving Eq. 8.68 forχ gives the Curie-Weiss law

χ =C

T − λC∼ (T −ΘC)

−1 . (8.69)

ΘC is the Curie-Weiss temperature. The form of the Curie-Weiss law is compatible with aneffective molecular field that scales linearly with the magnetization. Given the experimen-tal observation of Curie-Weiss like scaling, the mean field ansatz of Weiss seems thereforereasonable. However, we can already see that this simple theory fails, as it predicts an in-verse scaling with temperature for all T > TC, i. e., within molecular field theory ΘC = TC

and γ = 1 at variance with experiment. The exponent of 1, by the way, is characteristicfor any mean field theory.Given the qualitative plausibility of the theory, let us nevertheless use it to derive a firstorder of magnitude estimate for this (yet unspecified) molecular field. This phenomeno-logical consideration therefore follows the reverse logic to the ab initio one we prefer. Inthe latter, we would have analyzed the microscopic origin of the magnetic interaction andderived the molecular field as a suitable average, which in turn would have brought us intoa position to predict materials’ properties like the saturation magnetization and the crit-ical temperature. Now, we do the opposite: We will use the experimentally known valuesof Ms and TC to obtain a first estimate of the size of the molecular field, which might helpus lateron to identify the microscopic origin of the magnetic interactions (about which westill do not know anything). For the estimate, recall that the Curie constant C was givenby Eq. (8.40) as

C =Nµoµ2

Bg(JLS)2J(J + 1)

3kBV≈(N

V

)µom2

atom

3kB

, (8.70)

where we have exploited that J(J + 1) ∼ J2, allowing us to identify the atomic magneticmoment matom = µBg(JLS)J . With this, it is straightforward to arrive at the followingestimate of the molecular field at T = 0K,

Bmol(0K) = µ0λMs(0 K) = µ0TC

C

(N

Vms

)≈

3kBTCms

m2atom

. (8.71)

When discussing the content of Table 8.4 we had already seen that, for most ferromagnets,matom is at least of the same order of ms, so that plugging in the numerical constants wearrive at the following order of magnitude estimate

Bmol ∼[5TC in K]

[matom in µB]Tesla . (8.72)

Looking at the values listed in Table 8.4, one realizes that molecular fields are of theorder of some 103 Tesla, which is at least one order of magnitude more than the strongestmagnetic fields that can currently be produced in the laboratory. As take-home message

241

Page 29: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

we therefore keep in mind that the magnetic interactions must be quite strong to yieldsuch a high internal molecular field.Molecular field theory can also be employed to understand the temperature behavior of thezero-field spontaneous magnetization. As shown in Fig. 8.9 Ms(T ) decays smoothly fromits saturation value at T = 0K to zero at the critical temperature. Since the molecularfield produced by the magnetic interaction is indistinguishable from an applied externalone, the variation of Ms(T ) must be equivalent to the one of a paramagnet, only withthe external field replaced by the molecular one. Using the results from section 8.3.2 foratomic paramagnetism, we therefore obtain directly a behavior equivalent to Eq. (8.40)

Ms(T ) = Ms(0 K) BJ

(g(JLS)µBBmol

kBT

), (8.73)

i. e., the temperature dependence is entirely given by the Brillouin function defined in Eq.(8.41). BJ(η), in fact, exhibits exactly the behavior we just discussed: it decays smoothly tozero as sketched in Fig. 8.9. Similar to the reproduction of the Curie-Weiss law, mean fieldtheory provides us therefore with a simple rationalization of the observed phenomenon,but fails again on a quantitative level (and also does not tell us anything about theunderlying microscopic mechanism). The functional dependence of Ms(T ) in both limitsT → 0K and T ↗ TC is incorrect, with e.g.

Ms(T ↗ TC) ∼ (TC − T )1/2 (8.74)

and thus β = 1/2 instead of the observed values around 1/3. In the low temperaturelimit, the Brillouin function decays exponentially fast, which is also in gross disagreementwith the experimentally observed T 3/2 behavior (known as Bloch T 3/2 law). The failure topredict the correct critical exponents is a general feature mean field theories that cannotaccount for the fluctuations connected with phase transitions. The wrong low temperaturelimit, on the other hand, is due to the non-existence of a particular set of low-energyexcitations called spin-waves (or magnons), a point that we will come back to later.

8.4.3.2 Heisenberg and Ising Hamiltonian

Another class of phenomenological theories is called spin Hamiltonians. They are based onthe hypothesis that magnetic order is due to the coupling between discrete microscopicmagnetic moments. If these microscopic moments are, for example, connected to thepartially filled f -shells of RE atoms, one can replace the material by a discrete latticemodel, in which each site i has a magnetic moment mi. A coupling between two differentsites i and j in the lattice can then be described by a term

Hcouplingi,j = −

Jijµ2

B

mi ·mj , (8.75)

where Jij is known as the exchange coupling constant between the two sites (unit: eV). Inprinciple, the coupling can take any value between −Jijmimj/µ2

B (parallel (ferromagnetic)alignment) and +Jijmimj/µ2

B (antiparallel (antiferromagnetic) alignment) depending onthe relative orientation of the two moment vectors. Note that this simplified interactionhas no explicit spatial dependence anymore. If this is required (e.g. in a non-isotropiccrystal with a preferred magnetization axis), additional terms need to be added to theHamiltonian.

242

Page 30: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.11: Schematic illustrations of possible coupling mechanisms between localizedmagnetic moments: (a) direct exchange between neighboring atoms, (b) superexchangemediated by non-magnetic ions, and (c) indirect exchange mediated by conduction elec-trons.

Applying this coupling to the whole material, we first have to decide which lattice sitesto couple. Since we are now dealing with a phenomenological theory we have no rigorousguidelines that would tell us which interactions are important. Instead, we have to choosethe interactions (e.g. between nearest neighbors or up to next-nearest neighbors) and theirstrengths Jij. Later, after we have solved the model, we can revisit these assumptions andsee if they were justified. We can also use experimental (or first principles) data to fit theJij’s and hope to extract microscopic information. Often our choice is guided by intuitionconcerning the microscopic origin of the coupling. For the localized ferromagnetism ofthe rare earths one typically assumes a short-ranged interaction, which could either bebetween neighboring atoms (direct exchange) or reach across non-magnetic atoms in morecomplex structures (superexchange). As illustrated in Fig. 8.11 this conveys the idea thatthe interaction arises out of an overlap of electronic wavefunctions. Alternatively, onecould also think of an interaction mediated by the glue of conduction electrons (indirectexchange), cf. Fig. 8.11c. However, without a proper microscopic theory it is difficult todistinguish between these different possibilities.Having decided on the coupling, the Hamiltonian for the whole material becomes in itssimplest form of only pair-wise interactions

HHeisenberg =M∑

i,j=1

Hcouplingi,j = −

M∑

i,j=1

Jijµ2

B

mi ·mj . (8.76)

Similar to the discussion in chapter 6 on cohesion, such a pairwise summation is onlyjustified, when the interaction between two sites is not affected strongly by the presenceof magnetic moments at other nearby sites. Since as yet we do not know anything about

243

Page 31: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

the microscopic origin of the interaction, this is quite hard to verify.The particular spin Hamiltonian of Eq. (8.76) is known as Heisenberg Hamiltonian andis the starting point for many theories of ferromagnetism. It looks deceptively simple,but in fact it is difficult to solve. Up to now, no analytic solution has been found. Sinceoften analytic solutions are favored over numerical solutions further simplifications aremade. Examples are one- or two-dimensional Heisenberg Hamiltonians (possibly onlywith nearest neighbor interaction) or the Ising model:

H Ising = −M∑

i,j=1

Jijµ2

B

mi,z ·mj,z︸ ︷︷ ︸±1 only

. (8.77)

These models opened up a whole field of research in the statistical mechanics community.It quickly developed a life of its own and became more and more detached from the originalmagnetic problem. Instead it was and is driven by the curiosity to solve the generic modelsthemselves. In the meantime numerical techniques such as Monte Carlo developed thatcan solve the original Heisenberg Hamiltonian, although they do of course not reach thebeauty of a proper analytical solution.Since the Ising model has played a prominent role historically and because it has analyticsolutions under certain conditions we will illustrate its solution in one dimension. Forthis we consider the Ising model in a magnetic field with only next nearest neighborinteractions

H Ising = −M∑

<ij>

Jijσiσj −∑

j

hjσj (8.78)

where < ij > denotes that the sum only runs over nearest neighbor and hj is the magneticfield µ0H at site j and σi = ±1. Despite this simplified form no exact solution is knownfor the three dimensional case. For two dimensions Onsager found an exact solution in1944 for zero magnetic field. Let us reduce the problem to one dimension and furthersimplify to Ji,i+1 = J and hj = h:

H Ising = −JN∑

i=1

σiσi+1 − h∑

i

σi (8.79)

where we have used the periodic boundary condition σN+1 = σ1. The partition functionfor this Hamiltonian is

Z(N, h, T ) =∑

σ1

. . .∑

σN

eβ[J!N

i σiσi+1+12

!Ni (σi+σi+1)] , (8.80)

with β =1

kBT. In order to carry out the spin summation we define a 2×2 matrix P with

elements:

⟨σ|P |σ′⟩ = eβ[Jσσ′+h

2 (σ+σ′)] (8.81)

⟨1|P |1⟩ = eβ[J+h] (8.82)

⟨−1|P |− 1⟩ = eβ[J−h] (8.83)

⟨1|P |− 1⟩ = ⟨−1|P |1⟩ = e−βJ , (8.84)

244

Page 32: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

P is called the transfer matrix and can be written in more compact form as

P =

(eβ(J+h) e−βJ

e−βJ eβ(J−h)

). (8.85)

With this definition, the partition function becomes

Z(N, h, T ) =∑

σ1

. . .∑

σN

⟨σ1|P |σ2⟩⟨σ2|P |σ3⟩ . . . ⟨σN−1|P |σN⟩⟨σN |P |σ1⟩ (8.86)

=∑

σ1

⟨σ1|PN |σ1⟩ = Tr(PN) . (8.87)

To evaluate the trace, we have to diagonalize P , which yields the following eigenvalues

λ± = eβJ[cosh(βh)±

√sinh2(βh) + e−4βh

]. (8.88)

With this, the trace equates to

Tr(PN) = λN+ + λN− . (8.89)

Since we are interested in the thermodynamic limit N → ∞, λ+ will dominate over λ−,because λ+ > λ− for any h. This finally yields

Z(N, h, T ) = λN+ (8.90)

and yields the free energy per spin,

F (h, T ) = −kBT lnλ+ = −J −1

βln

[cosh(βh)±

√sinh2(βh) + e−4βh

]. (8.91)

The magnetization then follows as usual

M(h, T ) = −∂F

∂h=

sinh(βh) + sinh(βh) cosh(βh)√sinh2(βh)+e−4βh

cosh(βh) +√

sinh2(βh) + e−4βh

. (8.92)

The magnetization is shown in Fig. 8.12 for different ratios J/h. For zero magnetic field(h = 0) coth(βh) = 1 and sinh(βh) = 0 and the magnetization vanishes. The next-nearestneighbor Ising model therefore does not yield a ferromagnetic solution in the absence ofa magnetic field. It also does not produce a ferromagnetic phase transition at finite field,unless the temperature is 0. The critical temperature for this model is thus zero Kelvin.Returning now to the Heisenberg Hamiltonian, let us consider an elemental solid, in whichall atoms are equivalent and exhibit the same magnetic moment m. Inspection of Eq.(8.76) reveals that J > 0 describes ferromagnetic and J < 0 antiferromagnetic coupling.The ferromagnetic ground state is given by a perfect alignment of all moments in onedirection | ↑↑↑ . . . ↑↑⟩. Naively one would expect the lowest energy excitations to be asimple spin flip | ↑↓↑ . . . ↑↑⟩. Surprisingly, a detailed study of Heisenberg Hamiltoniansreveals that these states not eigenstates of the Hamiltonian. Instead, the low energyexcitations have a much more complicated wave-like form, with only slight directional

245

Page 33: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

0 0.1 0.2 0.3 0.4 0.5 0.6β

0

0.2

0.4

0.6

0.8

1

M(β)

h/J=0.1

h/J=0.5

h/J=0.01

h/J=1h/J=2h/J=5

Figure 8.12: The magnetization M(T ) in the 1-dimensional Ising model as a function ofinverse temperature β = (kBT )−1 for different ratios J/h.

Figure 8.13: Schematic representation of the spatial orientation of the magnetic momentsin a one-dimensional spin-wave.

246

Page 34: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

changes between neighboring magnetic moments as illustrated in Fig. 8.13. At secondthought it is then intuitive, that such waves will have only an infinitesimally higher energythan the ground state, if their wave vector is infinitely long. Such spin-waves are calledmagnons, and it is easy to see that they are missing from the mean-field description andalso from the Ising model. As a result the Heisenberg Hamiltonian correctly accounts forthe Bloch T 3/2-scaling of Ms(T ) for T → 0K. Magnons can in fact be expected quitegenerally, whenever there is a direction associated with the local order that can varyspatially in a continuous manner at an energy cost that becomes small as the wavelengthof the variation becomes large. Magnons are quite common in magnetism (see e.g. theground state of Cr) and their existence has been confirmed experimentally by e.g. neutronscattering.The Heisenberg model also does a good job at describing other characteristics of ferro-magnets. The high-temperature susceptibility follows a Curie-Weiss type law. Close tothe critical temperature the solution deviates from the 1/T -scaling. Monte Carlo simula-tions of Heisenberg ferromagnets yield critical exponents that are in very good agreementwith the experimental data. Moreover, experimental susceptibility or spontaneous mag-netization data can be used to fit the exchange constant Jij. The values obtained fornearest neighbor interactions are typically of the order of 10−2 eV for TM ferromagnetsand 10−3 − 10−4 eV for RE ferromagnets. These values are high compared to the energysplittings in dia- and paramagnetism and give us a first impression that the microscopicorigin of ferromagnetic coupling must be comparatively strong.We can also obtain this result by considering the mean-field limit of the HeisenbergHamiltonian:

HHeisenberg = −M∑

i,j=1

Jijµ2

B

mi ·mj

= −M∑

i=1

mi ·

(M∑

j=1

Jijµ2

B

mj

)= −

M∑

i=1

mi · µ0Hi . (8.93)

In the last step, we have realized that the term in brackets has the units of Tesla, and cantherefore be perceived as a magnetic field µ0H arising from the coupling of moment i withall other moments in the solid. However, up to now we have not gained anything by thisrewriting, because the effective magnetic field is different for every state (i. e., directionalorientation of all moments). The situation can only be simplified, by assuming a meanfield point of view and approximating this effective field by its thermal average < H >.This yields

µ0⟨H⟩ =M∑

j=1

Jijµ2

B

⟨mj⟩ =

∑Mj=1 Jij

µ2B

(N

V

)Ms(T ) ∼ λMs(T ) (8.94)

since the average magnetic moment at site i in an elemental crystal is simply given bythe macroscopic magnetization (dipole moment density) divided by the density of sites.As apparent from Eq. (8.94) we therefore arrive at a mean field that is linear in themagnetization as assumed by Weiss in his molecular field theory. In addition, we can noweven relate the field to the microscopic quantity Jij. If we assume for only nearest neighbor

247

Page 35: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

interactions in an fcc metal, it follows from Eq. (8.94) that

J = JNNij =

µ2BB

mol

12ms. (8.95)

Using the saturation magnetizations listed in Table 8.4 and the estimate of molecularfields of the order of 103 Tesla obtained above, we obtain exchange constants of the orderof 10−2 eV for the TM or 10−3− 10−4 eV for the RE ferromagnets in good agreement withthe values mentioned before.

8.4.4 Microscopic origin of magnetic interaction

We have seen that the phenomenological models capture the characteristics of ferromag-netism. However, what is still lacking is microscopic insight into the origin of the stronginteraction J . This must come from an ab initio theory. The natural choice (somehow also(falsely) conveyed by the name “magnetic interaction”) is to consider the direct dipolarinteraction energy of two magnetic dipoles m1 and m2, separated by a distance r. Forthis, macroscopic electrodynamics gives us

Jmag dipole =1

r3[m1 ·m2 − 3(m1 · r)(m2 · r)] , (8.96)

Omitting the angular dependence and assuming the two dipoles to be oriented parallelto each other and perpendicular to the vector between them we obtain the simplifiedexpression

Jmag dipole ≈1

r3m1m2 . (8.97)

Inserting magnetic moments of the order of µB, cf. Table 8.4, and a typical distance of2 Å between atoms in solids, this yields Jmag dipole ∼ 10−4 eV. This value is 1-2 orders ofmagnitude smaller than the exchange constants we derived using the phenomenologicalmodel in the previous section.Having ruled out magnetic dipole interactions the only thing left are electrostatic (Coulomb)interactions. These do in fact give rise to the strong coupling, as we will see in the fol-lowing. To be more precise it is the Coulomb interaction in competition with the Pauliexclusion principle. Loosely speaking, the exclusion principle keeps electrons with parallelspins apart, and thus reduces their Coulomb repulsion. The resulting difference in energybetween the parallel spin configuration and the antiparallel one is the exchange energywe were seeking. This alignment is favorable for ferromagnetism only in exceptional cir-cumstances because the increase in kinetic energy associated with parallel spins usuallyoutweighs the favorable decrease in potential energy. The difficulty in understanding mag-netic ordering lies in the fact that we need to go beyond the independent electron (oneparticle) picture. In the next few sections, we will look at how the Pauli principle andthe Coulomb interaction can lead to magnetic effects. We will do this for two simple yetopposite model cases: two localized magnetic moments and the free electron gas.

8.4.4.1 Exchange interaction between two localized moments

The most simple realization of two localized magnetic moments is the familiar problemof the hydrogen molecule. Here the two magnetic moments are given by the spins of the

248

Page 36: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

two electrons (1) and (2), which belong to the two H ions A and B. If there was only oneion and one electron, we would simply have the Hamiltonian of the hydrogen atom (ho)

ho|φ⟩ = ϵo|φ⟩ , (8.98)

with its ground state solution |φ⟩ = |r, σ⟩. For the molecule we then also have to considerthe interaction between the two electrons and two ions themselves, as well as between theions and electrons. This leads to the new Hamiltonian

H|Φ⟩ = (h0(A) + h0(B) + hint)|Φ⟩ , (8.99)

where the two-electron wave function |Φ⟩ is now a function of the positions of the ions Aand B, the two electrons (1) and (2), and their two spins σ1, σ2. Since H has no explicitspin dependence, all spin operators commute with H and the total two-electron wavefunction must factorize

|Φ⟩ = |Ψorb⟩|χspin⟩ , (8.100)

i. e., into an orbital part and a pure spin part. As a suitable basis for the latter, we canchoose the eigenfunctions of S2 and Sz, so that the spin part of the wave function is givenby

|χspin⟩S = 2−1/2 (| ↑↓⟩ − | ↓↑⟩) S = 0 Sz = 0|χspin⟩T1 = | ↑↑⟩ S = 1 Sz = 1|χspin⟩T2 = 2−1/2 (| ↑↓⟩ + | ↓↑⟩) S = 1 Sz = 0|χspin⟩T3 = | ↓↓⟩ S = 1 Sz = −1 .

The fermionic character of the two electrons (and the Pauli exclusion principle) manifestsitself at this stage in the general postulate that the two-electron wave function mustchange sign when interchanging the two indistinguishable electrons. |χspin⟩S changes signupon interchanging the two spins, while the three triplet wave functions |χspin⟩T do not.Therefore only the following combinations are possible

|Φ⟩singlet = |Ψorb,sym⟩ |χS⟩|Φ⟩triplet = |Ψorb,asym⟩ |χT⟩ , (8.101)

where |Ψorb,sym(1, 2)⟩ = |Ψorb,sym(2, 1)⟩ and |Ψorb,asym(1, 2)⟩ = −|Ψorb,asym(2, 1)⟩. Thistells us that although the Hamiltonian of the system is spin independent a correlationbetween the spatial symmetry of the wave function and the total spin of the systememerges from the Pauli principle.For the hydrogen molecule the full solution can be obtained numerically without too muchdifficulty. However, here we are interested in a solution that gives us insight and that allowsus to connect to the Heisenberg picture. We therefore start in the fully dissociated limitof infinitely separated atoms ( ⟨Φ|hint|Φ⟩ = 0). The total wave function must separateinto a product of wave functions from the isolated H atoms

φA(r) =1√Ne−α(r−RA) . (8.102)

However, the symmetry of the two electron wave function has to be maintained. Omittingthe two ionic choices φ(1A)φ(2A) and φ(1B)φ(2B) that have a higher energy, the two low

249

Page 37: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

energy choices are

|Ψ∞orb,sym⟩ = 2−1/2 [ |φ(1A)⟩|φ(2B)⟩ + |φ(2A)⟩|φ(1B)⟩ ]

|Ψ∞orb,asym⟩ = 2−1/2 [ |φ(1A)⟩|φ(2B)⟩ − |φ(2A)⟩|φ(1B)⟩ ] . (8.103)

Here |φ(1A)⟩ is our notation for electron (1) being in orbit around ion A. These two wavefunctions correspond to the ground state of H2 in either a spin singlet or a spin tripletconfiguration, and it is straightforward to verify that both states are degenerate in energywith 2ϵo.Let us now consider what happens when we bring the two atoms closer together. At acertain distance, the two electron densities will slowly start to overlap and to change.Heitler and London made the following approximation for the two wave functions

φA(r) =1√Ne−αr

"1+β

r·RAr

#

. (8.104)

The two-electron problem is then still of the form given in Eq. (8.103)

|ΨHLorb,sym⟩ = (2 + 2S)−1/2 [ |φ(1A)⟩|φ(2B)⟩ + |φ(2A)⟩|φ(1B)⟩ ] (8.105)

|ΨHLorb,asym⟩ = (2− 2S)−1/2 [ |φ(1A)⟩|φ(2B)⟩ − |φ(2A)⟩|φ(1B)⟩ ] ,

where S is the overlap integral

S = ⟨φ(1A)|φ(1B)⟩⟨φ(2A)|φ(2B)⟩ = |⟨φ(1A)|φ(1B)⟩|2 . (8.106)

Due to this change, the singlet ground state energy ES = ⟨Φ|H|Φ⟩singlet and the tripletground state energy ET = ⟨Φ|H|Φ⟩triplet begin to deviate, and we obtain

ET − ES = 2CS − A

1− S2, (8.107)

where

C = ⟨φ(1A)|⟨φ(2B)| Hint |φ(2B)⟩|φ(1A)⟩ (Coulomb integral) (8.108)A = ⟨φ(1A)|⟨φ(2B)| Hint |φ(2A)⟩|φ(1B)⟩ (Exchange integral) . (8.109)

C is simply the Coulomb interaction of the two electrons with the two ions and themselves,while A arises from the exchange of the indistinguishable particles. If the exchange didnot incur a sign change (as e.g. for two bosons), then A = CS and the singlet-tripletenergy splitting would vanish. This tells us that it is the fermionic character of the twoelectrons that yields A = CS as a consequence of the Pauli exclusion principle and in turna finite energy difference between the singlet ground state (antiparallel spin alignment)and the triplet ground state (parallel spin alignment). For the H2 molecule, the tripletground state is in fact less energetically favorable, because it has an asymmetric orbitalwave function. The latter must have a node exactly midway between the two atoms,meaning that there is no electron density at precisely that point where they could screenthe Coulomb repulsion between the two atoms most efficiently. As we had anticipated,the coupling between the two localized magnetic moments is therefore the result of asubtle interplay of the electrostatic interactions and the Pauli exclusion principle. What

250

Page 38: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

determines the alignment is the availability of a spin state that minimizes the Coulombrepulsion between electrons once Fermi statistics has been taken into account. For thisreason, magnetic interactions are also often referred to as exchange interactions (and thesplitting between the corresponding energy levels as the exchange splitting).By evaluating the two integrals C and A in Eq. (8.109) using the one-electron wave func-tions of the H atom problem, we can determine the energy splitting (ET−ES). Realizingthat the two level problem of the H2 molecule can be also described by the HeisenbergHamiltonian of Eq. (8.76) (as we will show in the next section), we have therefore found away to determine the exchange constant J . The Heitler-London approximation sketchedin this section can be generalized to more than two magnetic moments, and is thenone possible way for computing the exchange constants of Heisenberg Hamiltonians for(anti)ferromagnetic solids. However, this works only in the limit of highly localized mag-netic moments, in which the wave function overlap is such that it provides an exchangeinteraction, but the electrons are not too delocalized (in which case the Heitler-Londonapproximation of Eq. (8.105) is unreasonable). With this understanding of the interac-tion between two localized moments, also the frequently used restriction to only nearest-neighbor interactions in Heisenberg Hamiltonians becomes plausible. As a note aside, inthe extreme limit of a highly screened Coulomb interaction that only operates at eachlattice site itself (and does not even reach the nearest neighbor lattice sites anymore), themagnetic behavior can also be described by a so-called Hubbard model, in which electronsare hopping on a discrete lattice and pay the price of a repulsive interaction U when twoelectrons of opposite spin occupy the same site.

8.4.4.2 From the Schrödinger equation to the Heisenberg Hamiltonian

Having gained a clearer understanding how the exchange interaction between two atomsarises we will now generalize this and derive a magnetic Hamiltonian from the Schrödingerequation. Recapping Chapter 0 the non-relativistic many-body Hamiltonian for the elec-trons is

H = −∑

i

∇2

2−∑

ia

Z

|ri −Ra|+

1

2

i =j

1

|ri − rj|. (8.110)

It is more convenient to switch to 2nd quantization by introducing the field operators

ψ(r) =∑

α,n,λ

φnλ(r− rα)anλ(rα), (8.111)

where anλ(rα) annihilates an excitation from state n and spin state λ at site rα and φ isthe associated orbital. For the interaction part of the Hamiltonian we then obtain in 2ndquantization

1

2

α1α2α3α4n1n2n3n4

λ1λ2

⟨α1n1α2n2|V |α3n3α4n4⟩a†n1λ1(rα1)a

†n2λ2

(rα2)an3λ2(rα3)an4λ1(rα4) . (8.112)

This is still exact, but a bit unwieldy and not very transparent.3 Since the relevant orbitalsare localized we can restrict the terms to those in which α3 = α1 and α4 = α2 or α3 = α2

3 Try to explain why the operators occur in the order a†a†aa instead of a†aa†a. Writing the creatorsto the left and annihilators to the right is called normal ordering in the literature.

251

Page 39: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

and α4 = α1. The remaining terms involve various orbital excitations induced by theCoulomb interaction. These lead to off-diagonal exchange. We shall therefore restrict eachelectron to a definite orbital state. That is, we shall keep only those terms in which n3 = n1

and n4 = n2 or n3 = n2 and n4 = n1. If, for simplicity, we also neglect orbital-transferterms, in which two electrons interchange orbital states, then the interaction part reducesto

1

2

αα′

nn′

λλ′

[⟨αnα′n′|V |αnα′n′⟩a†nλ(rα)a†n′λ′(rα′)an′λ′(rα′)anλ(rα) (8.113)

+⟨αnα′n′|V |α′n′αn⟩ ] a†nλ(rα)a†n′λ′(rα′)anλ′(rα)an′λ(rα′) (8.114)

The first term is the direct and the second the exchange term that we encountered in theprevious section. Using the fermion anticommutation relation

{anλ(rα), a

†n′λ′(rα′)

}= δαα′δnn′δλλ′ , (8.115)

we may write the exchange term as

1

2

αα′

nn′

λλ′

Jnn′(rα, rα′)a†nλ(rα)anλ′(rα)a†n′λ′(rα′)an′λ(rα′) . (8.116)

Expanding the spin sums we can rewrite his

1

2

αα′

nn′

λλ′

Jnn′(rα, rα′)

[1

4+

1

4S(rα) · S(rα′)

], (8.117)

where S is the spin (or magnetic moment) of the atom at site rα. Now mapping over tolattice sites (rα → i and rα′ → j) we finally obtain the Heisenberg Hamiltonian fromEq. (8.75)

Hspin =∑

ij

JijSi · Sj . (8.118)

We have thus shown that magnetism does not arise from a spin-dependent part in theSchrödinger equation, but from the Pauli exclusion principle that the full many-body wavefunction has to satisfy. We could now use the Coulomb integrals derived in this sectionor the microscopic interaction we have derived in the previous section for Jij to fill theHeisenberg Hamiltonian with life.

8.4.4.3 Exchange interaction in the homogeneous electron gas

The conclusion from the localized moment model is that feeding Heitler-London derivedexchange constants into Hubbard or Heisenberg models provides a good semi-quantitativebasis for describing the ferromagnetism of localized moments as is typical for rare earthsolids. However, the other extreme of the more itinerant ferromagnetism found in thetransition metals cannot be described. For this it is more appropriate to return to the

252

Page 40: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

homogeneous electron gas, i. e., treating the ions as a jellium background. We had stud-ied this model already in section 8.3.3 for the independent electron approximation (zeroelectron-electron interaction) and found that the favorable spin alignment in an externalfield is counteracted by an increase in kinetic energy resulting from the occupation ofenergetically higher lying states. The net effect was weak and since there is no drivingforce for a spin alignment without an applied field we obtained a (Pauli) paramagneticbehavior.If we now consider the electron-electron interaction, we find a new source for spin align-ment in the exchange interaction, i. e., an exchange-hole mediated effect to reduce therepulsion of parallel spin states. This new contribution supports the ordering tendency ofan applied field, and could (if strong enough) maybe even favor ordering in the absenceof an applied field. To estimate whether this is possible, we consider the electron gas inthe Hartree-Fock (HF) approximation as the minimal theory accounting for the exchangeinteraction. In the chapter on cohesion, we already derived the HF energy per electron,but without explicitly considering spins

(E/N)jellium = Ts + EXC HF≈

30.1 eV(

rsaB

)2 −12.5 eV(

rsaB

) . (8.119)

Compared to the independent electron model of section 8.3.3, which only yields the firstkinetic energy term, a second (attractive) term arises from exchange. To discuss magneticbehavior we now have to extend this description to the spin-polarized case. This is donein the most straightforward manner by rewriting the last expression as a function ofthe electron density (exploiting the definition of the Wigner-Seitz radius as rs/aB =(

34π (V/N)

)1/3)

Ejellium,HF(N) = N

{78.2 eV

(N

V

)2/3

− 20.1 eV

(N

V

)1/3}

. (8.120)

The obvious generalization to a spin-resolved gas is then

Espinjellium,HF(N

↑, N↓) = Ejellium,HF(N↑) + Ejellium,HF(N

↓) = (8.121)

=

{78.2 eV

[(N↑

V

)2/3

+

(N↓

V

)2/3]− 20.1 eV

[(N↑

V

)1/3

+

(N↓

V

)1/3]}

,

where N↑ is the number of electrons with up spin, and N↓ with down spin (N↑ + N↓ =N). More elegantly, the effects of spin-polarization can be discussed by introducing thepolarization

P =N↑ −N↓

N, (8.122)

which possesses the obvious limits P = ±1 for complete spin alignment and P = 0 for anon-spinpolarized gas. Making use of the relations N↑ = N

2 (1 + P ) and N↓ = N2 (1− P ),

Eq. (8.121) can be cast into

Ejellium,HF(N,P ) = (8.123)

NT

{1

2

[(1 + P )5/3 + (1− P )5/3

]−

5

4α[(1 + P )4/3 + (1− P )4/3

]},

253

Page 41: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.14: Plot of ∆E(P ), cf. Eq. (8.124), as a function of the polarization P and forvarious values of α. A ferromagnetic solution is found for α > 0.905.

with α = 0.10(V/N)1/3. Ferromagnetism, i. e., a non-vanishing spontaneous magnetizationat zero field, can occur when a system configuration with P = 0 leads to a lower energythan the non-polarized case, ∆E(N,P ) = E(N,P )−E(N, 0) < 0. Inserting the expressionfor the HF total energy we thus arrive at

∆E(N,P ) = (8.124)

NT

{1

2

[(1 + P )5/3 + (1− P )5/3 − 2

]−

5

4α[(1 + P )4/3 + (1− P )4/3 − 2

]}.

Figure 8.14 shows ∆E(N,P ) as a function of the polarization P for various values of α.We see that the condition ∆E(N,P ) < 0 can only be fulfilled for

α > αc =2

5(21/3 + 1) ≈ 0.905 , (8.125)

in which case the lowest energy state is always given by a completely polarized gas (P = 1).Using the definition of α, this condition can be converted into a maximum electron densitybelow which the gas possesses a ferromagnetic solution. In units of the Wigner-Seitz radiusthis is (

rsaB

)>∼ 5.45 , (8.126)

i. e., only electron gases with a density that is low enough to yield Wigner-Seitz radiiabove 5.6 aB would be ferromagnetic. Recalling that 1.8 < rs < 5.6 for real metals, onlyCs would be a ferromagnet in the Hartree-Fock approximation.Already this is obviously not in agreement with reality (Cs is paramagnetic and Fe/Co/Niare ferromagnetic). However, the situation becomes even worse, when we start to improvethe theoretical description by considering further electron exchange and correlation effects

254

Page 42: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

beyond Hartree-Fock. The prediction of the paramagnetic-ferromagnetic phase transitionin the homogeneous electron gas has a long history, and the results vary dramatically witheach level of exchange-correlation. The maximum density below which ferromagnetismcan occur is, however, consistently obtained much lower than in Eq. (8.126). The mostquantitative work so far predicts ferromagnetism for electron gases with densities below(rs/aB) ≈ 50±2 [F.H. Zhong, C.Lin, and D.M. Ceperley, Phys. Rev. B 66, 036703 (2002)].From this we conclude that Hartree-Fock theory allows us to qualitatively understand thereason behind itinerant ferromagnetism of delocalized electrons, but it overestimates theeffect of the exchange interaction by one order of magnitude in terms of rs. Yet, evena more refined treatment of electron exchange and correlation does not produce a gooddescription of itinerant magnetism for real metals, none of which (with 1.8 < rs < 5.6)would be ferromagnetic at all in this theory. The reason for this failure cannot be foundin our description of exchange and correlation, but must come from the jellium modelitself. The fact that completely delocalized electrons (as in e.g. simple metals) cannotlead to ferromagnetism is correctly obtained by the jellium model, but not that Fe/Co/Niare ferromagnetic. This results primarily from their partially filled d-electron shell, whichis not well described by a free electron model as we had already seen in the chapter oncohesion. In order to finally understand the itinerant ferromagnetism of the transitionmetals, we therefore need to combine the exchange interaction with something that wehave not yet considered at all: band structure effects.

8.4.5 Band consideration of ferromagnetism

In chapter 3 we had seen that an efficient treatment of electron correlation together withthe explicit consideration of the ionic positions is possible within density-functional theory(DFT). Here, we can write for the total energy of the problem

E/N = Ts[n] + E ion−ion[n] + Eel−ion[n] + EHartree[n] + EXC[n] , (8.127)

where all terms are uniquely determined functions of the electron density n(r). The mostfrequently employed approximation to the exchange-correlation term is the local densityapproximation (LDA), which gives the XC-energy at each point r as the one of the homo-geneous electron gas with equivalent density n(r). Recalling our procedure from the lastsection, it is straightforward to generalize to the so-called spin-polarized DFT, with

E/N = Ts[n↑, n↓] + E ion−ion[n] + Eel−ion[n] + EHartree[n] + EXC[n↑, n↓] . (8.128)

n↑(r) and n↓(r) then are the spin up and down electron densities, respectively. We onlyneed to consider them separately in the kinetic energy term (since there might be adifferent number of up- and down- electrons) and in the exchange-correlation term. Forthe latter, we proceed as before and obtain the local spin density approximation (LSDA)from the XC-energy of the spin-polarized homogeneous electron gas that we discussed inthe previous section. Introducing the magnetization density

m(r) =(n↑(r)− n↓(r)

)µB , (8.129)

we can write the total energy as a function of n(r) and m(r). Since the integral of themagnetization density over the whole unit cell yields the total magnetic moment, self-consistent solutions of the coupled Kohn-Sham equations can then either be obtained

255

Page 43: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.15: Spin-resolved density of states (DOS) for bulk Fe and Ni in DFT-LSDA [fromV.L. Moruzzi, J.F. Janak and A.R. Williams, Calculated electronic properties of metals,Pergamon Press (1978)].

under the constraint of a specified magnetic moment (fixed spin moment (FSM) calcula-tions), or directly produce the magnetic moment that minimizes the total energy of thesystem.

Already in the LSDA this approach is very successful and yields a non-vanishing magneticmoment for Fe, Co, and Ni. Resulting spin-resolved density of states (DOS) for Fe and Niare exemplified in Fig. 8.15. After self-consistency is reached, these three ferromagneticmetals exhibit a larger number of electrons in one spin direction (called majority spin)than in the other (called minority spin). The effective potential seen by up and down elec-trons is therefore different, and correspondingly the Kohn-Sham eigenlevels differ (whichupon filling the states up to a common Fermi-level yields the different number of up anddown electrons that we had started with). Quantitatively one obtains in the LSDA thefollowing average magnetic moments per atom, ms = 2.1 µB (Fe), 1.6 µB (Co), and 0.6 µB

(Ni), which compare excellently with the experimental saturated spontaneous magnetiza-tions at T = 0K listed in Table 8.4. We can therefore state that DFT-LSDA reproduces

256

Page 44: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

the itinerant ferromagnetism of the elemental transition metals qualitatively and quanti-tatively! Although this is quite an achievement, it is still somewhat unsatisfying, becausewe do not yet understand why only these three elements in the periodic table exhibitferromagnetism.

8.4.5.1 Stoner model of itinerant ferromagnetism

To derive at such an understanding, we take a closer look at the DOS of majority andminority spins in Fig. 8.15. Both DOS exhibit a fair amount of fine structure, but the mostobvious observation is that the majority and minority DOS seem to be shifted relativeto each other, but otherwise very similar. This can be rationalized by realizing that themagnetization density (i. e., the difference between the up and down electron densities) issignificantly smaller than the total electron density itself. For the XC-correlation potentialone can therefore write a Taylor expansion to first order in the magnetization density as

V ↑↓XC(r) =

δE↑↓XC[n(r),m(r)]

δn(r)≈ V 0

XC(r)± V (r)m(r) , (8.130)

where V 0XC(r) is the XC-potential in the non-magnetic case. If the electron density is only

slowly varying, one may assume the difference in up and down XC-potential (i. e., theterm V (r)m(r)) to also vary slowly. In the Stoner model, one therefore approximates itwith a constant term given by

V ↑↓XC,Stoner(r) = V 0

XC(r)±1

2IM . (8.131)

The proportionality constant I is called the Stoner parameter (or exchange integral), andM =

∫unit−cell m(r)dr (and thus M = (ms/µB)× number of elements in the unit cell!).

Since the only difference between the XC-potential seen by up and down electrons is aconstant shift (in total: IM), the wave functions in the Kohn-Sham equations are notchanged compared to the non-magnetic case. The only effect of the spin-polarization inthe XC-potential is therefore to shift the eigenvalues ϵ↑↓i by a constant to lower or highervalues

ϵ↑↓i = ϵ0i ± 1/2IM . (8.132)

The band structure and in turn the spin DOS are shifted with respect to the non-magneticcase,

N↑↓(E) =∑

i

BZ

δ(E − ϵ↑↓k,i)dk = N0(E ± 1/2IM) , (8.133)

i. e., the Stoner approximation corresponds exactly to our initial observation of a constantshift between the up and down spin DOS.So what have we gained by? So far nothing, but as we will see, we can employ theStoner model to arrive at a simple condition that tells us by looking at an non-magneticcalculation, whether the material is likely to exhibit a ferromagnetic ground state or not.What we start out with are the results of a non-magnetic calculation, namely the non-magnetic DOS N0(E) and the total number of electrons N . In the magnetic case, Nremains the same, but what enters as a new free variable is the magnetization M . In theStoner model the whole effect of a possible spin-polarization is contained in the parameterI. We now want to relate M with I to arrive at a condition that tells us which I will yield

257

Page 45: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.16: Graphical solution of the relation M = F (M) of Eq. (8.136) for two charac-teristic forms of F (M).

a non-vanishing M (and thereby a ferromagnetic ground state). We therefore proceed bynoticing that N and M are given by integrating over all occupied states of the shiftedspin DOSs up to the Fermi-level

N =

ϵF

[N0(E + 1/2IM) +N0(E − 1/2IM)

]dE (8.134)

M =

ϵF

[N0(E + 1/2IM)−N0(E − 1/2IM)

]dE . (8.135)

Since N0(E) and N are fixed, these two equations determine the remaining free variablesof the magnetic calculation, ϵF and M . However, the two equations are coupled (i. e.,they depend on the same variables), so that one has to solve them self-consistently. Theresulting M is therefore characterized by self-consistently fulfilling the relation

M =

ϵF(M)

[N0(E + 1/2IM)−N0(E − 1/2IM)

]dE ≡ F (M) (8.136)

As expected M is a only function F (M) of the Stoner parameter I, the DOS N0(E) ofthe non-magnetic material, and the filling of the latter (determined by ϵF and in turn byN). To really quantify the resulting M at this point, we would need an explicit knowledgeof N0(E). Yet, even without such a knowledge we can derive some universal propertiesof the functional dependence of F (M) that simply follow from its form and the fact thatN0(E) > 0,

F (M) = F (−M) F (0) = 0F (±∞) = ±M∞ F ′(M) > 0 .

Here M∞ corresponds to the saturation magnetization at full spin polarization, when allmajority spin states are occupied and all minority spin states are empty. The structure ofF (M) must therefore be quite simple: monotonously rising from a constant value at −∞to a constant value at +∞, and in addition exhibiting mirror symmetry. Two possiblegeneric forms of F (M) that comply with these requirements are sketched in Fig. 8.16.In case (A), there is only the trivial solution M = 0 to the relation M = F (M) of Eq.

258

Page 46: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

(8.136), i. e., only a non-magnetic state results. In contrast, case (B) has three possiblesolutions. Apart from the trivial M = 0, two new solutions with a finite spontaneousmagnetization M = Ms = 0 emerge. Therefore, if the slope of F (M = 0) is larger than 1,the monotonous function F (M) must necessarily cross the line M in another place, too.A sufficient criterion for the existence of ferromagnetic solutions with finite magnetizationMs is therefore F ′(0) > 1.

Taking the derivative of Eq. (8.136), we find that F ′(0) = IN0(ϵF), and finally arrive atthe famous Stoner criterion for ferromagnetism

IN0(ϵF) > 1 . (8.137)

A sufficient condition for a material to exhibit ferromagnetism is therefore to have a highdensity of states at the Fermi-level and a large Stoner parameter. The latter is a materialconstant that can be computed within DFT (by taking the functional derivative of V ↑↓

xc (r)with respect to the magnetization; see e.g. J.F. Janak, Phys. Rev. B 16, 255 (1977)).Loosely speaking it measures the localization of the wave functions at the Fermi-level.In turn it represents the strength of the exchange interaction per electron, while N0(ϵF)tells us how many electrons are able to participate in the ordering. Consequently, theStoner criterion has a very intuitive form: strength per electron times number of electronsparticipating. If this exceeds a critical threshold, a ferromagnetic state results.

Figure 8.17 shows the quantities entering the Stoner criterion calculated within DFT-LSDA for all elements. Gratifyingly, only the three elements Fe, Co and Ni that areindeed ferromagnetic in reality fulfill the condition. With the understanding of the Stonermodel we can now, however, also rationalize why it is only these three elements and notthe others: For ferromagnetism we need a high number of states at the Fermi-level, suchthat only transition metals come into question. In particular, N0(ϵF) is always highestfor the very late TMs with an almost filled d-band. In addition, we also require a stronglocalization of the wave functions at the Fermi-level to yield a high Stoner parameter.And here, the compact 3d orbitals outwin the more delocalized ones of the 4d and 5delements. Combining the two demands, only Fe, Co, and Ni remain.

Yet, it is interesting to see that also Ca, Sc and Pd come quite close to fulfilling the Stonercriterion. This is particularly relevant, if we consider other situations than the bulk. TheDOSs of solids (although in general quite structured) scale to first order inversely withthe band width W . In other words, the narrower the band, the higher the number ofstates per energy within the band. Just as much as by a stronger localization, W is alsodecreased by a lower coordination. In the atomic limit, the band width is zero and theStoner criterion is always fulfilled. Correspondingly, all atoms with partially filled shellshave a non-vanishing magnetic moment, given by Hund‘s rules. Somewhere intermediateare surfaces and thin films, where the atoms have a reduced coordination compared tothe bulk. The bands are then narrower and such systems can exhibit a larger magneticmoment or even a reversal of their magnetic properties (i. e., when they are non-magneticin the bulk). The magnetism of surfaces and thin films is therefore substantially richerthan the bulk magnetism discussed in this introductory lecture. Most importantly, it isalso the basis of many technological applications.

259

Page 47: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.17: Trend over all elements of the Stoner parameter I, of the non-magnetic DOSat the Fermi-level (here denoted as D(EF)) and of their product. The Stoner criterionis only fulfilled for the elements Fe, Co, and Ni [from H. Ibach and H. Lüth, Solid StatePhysics, Springer (1990)].

260

Page 48: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.18: Schematic illustration of magnetic domain structure.

8.5 Magnetic domain structure

With the results from the last section we can understand the ferromagnetism of the REmetals as a consequence of the exchange coupling between localized moments due to thepartially filled (and mostly atomic-like) f -shells. The itinerant ferromagnetism of the TMsFe, Co and Ni, on the other hand, comes from the exchange interaction of the still local-ized, but no longer atomic-like d orbitals together with the delocalized s-electron density.We can treat the prior localized magnetism using discrete Heisenberg or Hubbard modelsand exchange constants from Heitler-London type theories (or LDA+U type approaches),while the latter form of magnetism is more appropriately described within band theo-ries like spin-DFT-LSDA. Despite this quite comprehensive understanding, an apparentlycontradictory feature of magnetic materials is that a ferromagnet does not always showa macroscopic net magnetization, even when the temperature is well below the Curietemperature. In addition it is hard to grasp, how the magnetization can be affected byexternal fields, when we know that the molecular internal fields are vastly larger.The explanation for these puzzles lies in the fact that two rather different forces operatebetween the magnetic moments of a ferromagnet. At very short distance (of the orderof atomic spacings), magnetic order is induced by the powerful, but quite short-rangedexchange forces. Simultaneously, at distances large compared with atomic spacings, themagnetic moments still interact as magnetic dipoles. We stressed in the beginning of thelast section that the latter interaction is very weak; in fact between nearest neighbors it isorders of magnitude smaller than the exchange interaction. On the other hand, the dipolarinteraction is rather long-ranged, falling off only as the inverse cube of the separation. Con-sequently, it can still diverge, if one sums over all interactions between a large populationthat is uniformly magnetized. In order to accommodate these two competing interactions(short range alignment, but divergence if ordered ensembles become too large), ferromag-nets organize themselves into domains as illustrated in Fig. 8.18. The dipolar energy isthen substantially reduced, since due to the long-range interaction the energy of everyspin drops. This bulk-like effect is opposed by the unfavorable exchange interactions withthe nearby spins in the neighboring misaligned domains. Because, however, the exchangeinteraction is short-ranged, it is only the spins near the domain boundaries that will havetheir exchange energies raised, i. e., this is a surface effect. Provided that the domains arenot too small, domain formation will be favored in spite of the vastly greater strength of

261

Page 49: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.19: Schematic showing the alignment of magnetic moments at a domain wallwith (a) an abrupt boundary and (b) a gradual boundary. The latter type is less costlyin exchange energy.

the exchange interaction: Every spin can lower its (small) dipolar energy, but only a few(those near the domain boundaries) have their (large) exchange energy raised.Whether or not a ferromagnet exhibits a macroscopic net magnetization, and how thismagnetization is altered by external fields, has therefore less to do with creating alignment(or ripping it apart), but with the domain structure and how its size and orientationaldistribution is changed by the field. In this respect it is important to notice that theboundary between two domains (known as the domain wall or Bloch wall) has often notthe (at first thought intuitive) abrupt structure. Instead a spread out, gradual change asshown in Fig. 8.19 is less costly in exchange energy. All an external field has to do then isto slightly affect the relative alignments (not flip one spin completely), in order to inducea smooth motion of the domain wall, and thereby increase the size of favorably aligneddomains. The way how the comparably weak external fields can "magnetize" a piece ofunmagnetized ferromagnet is therefore to rearrange and reorient the various domains, cf.Fig. 8.20. The corresponding domain wall motion can hereby be reversible, but it may alsowell be that it is hindered by crystalline imperfections (e.g. a grain boundary), throughwhich the wall will only pass when the driving force due to the applied field is large. Whenthe aligning field is removed, these defects may prevent the domain walls from returning totheir original unmagnetized configuration, so that it becomes necessary to apply a ratherstrong field in the opposite direction to restore the original situation. The dynamics of thedomains is thus quite complicated, and their states depend in detail upon the particularhistory that has been applied to them (which fields, how strong, which direction).As a consequence, ferromagnets always exhibit so-called hysteresis effects, when plottingthe magnetization of the sample against the external magnetic field strength as shown inFig. 8.21. Starting out with an initially unmagnetized crystal (at the origin), the appliedexternal field induces a magnetization, which reaches a saturation value Ms when alldipoles are aligned. On removing the field, the irreversible part of the domain boundarydynamics leads to a slower decrease of the magnetization, leaving the so-called remnantmagnetization MR at zero field. One needs a reverse field (opposite to the magnetizationof the sample) to further reduce the magnetization, which finally reaches zero at theso-called coercive field Bc.Depending on the value of this coercive field, one distinguishes between soft and hardmagnets in applications: A soft magnet has a small Bc and exhibits therefore only a small

262

Page 50: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.20: (a) If all of the magnetic moments in a ferromagnet are aligned they producea magnetic field which extends outside the sample. (b) The net macroscopic magnetizationof a crystal can be zero, when the moments are arranged in (randomly oriented) domains.(c) Application of an external magnetic field can move the domain walls so that thedomains which have the moments aligned parallel to the field are increased in size. Intotal this gives rise to a net magnetization.

Figure 8.21: Typical hysteresis loop for a ferromagnet.

263

Page 51: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.22: Hysteresis loops for soft and hard magnets.

hysteresis loop as illustrated in Fig. 8.22. Such materials are employed, whenever oneneeds to reorient the magnetization frequently and without wanting to resort to strongfields to do so. Typical applications are kernels in engines, transformators, or switchesand examples of soft magnetic materials are Fe or Fe-Ni, Fe-Co alloys. The idea of ahard magnet which requires a huge coercive field, cf. Fig. 8.22, is to maximally hinderthe demagnetization of a once magnetized state. A major field of application is thereforemagnetic storage with Fe- or Cr-Oxides, schematically illustrated in Fig. 8.23. The medium(tape or disc) is covered by many small magnetic particles each of which behaves as asingle domain. A magnet – write head – records the information onto this medium byorienting the magnetic particles along particular directions. Reading information is thensimply the reverse: The magnetic field produced by the magnetic particles induces amagnetization in the read head. In magnetoresistive materials, the electrical resistivitychanges in the presence of a magnetic field so that it is possible to directly convert themagnetic field information into an electrical response. In 1988 a structure consisting ofextremely thin, i. e., three atom-thick-alternative layers of Fe and Co was found to have amagnetoresistance about a hundred times larger than that of any natural material. Thiseffect is known as giant magnetoresistance and is in the meanwhile exploited in todaysdisc drives.

264

Page 52: 8 Magnetism - TU Berlin · 2017-08-20 · 8 Magnetism 8.1 Introduction The topic of this part of the lecture are the magnetic properties of materials. Most materi-als are generally

Figure 8.23: Illustration of the process of recording information onto a magnetic medium.The coil converts an electrical signal into a magnetic field in the write head, which in turnaffects the orientation of the magnetic particles on the magnetic tape or disc. Here thereare only two possible directions of magnetization of the particles, so the data is in binaryform and the two states can be identified as 0 and 1.

265