Top Banner
Introduction to Solid State Physics F RANK G ¨ OHMANN Fakult¨ at f ¨ ur Mathematik und Naturwissenschaften Bergische Universit¨ at Wuppertal arXiv:2101.01780v1 [cond-mat.str-el] 5 Jan 2021
116

Introduction to Solid State Physics - arXiv

Oct 04, 2021

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Introduction to Solid State Physics - arXiv

Introduction to Solid State Physics

FRANK GOHMANN

Fakultat fur Mathematik und NaturwissenschaftenBergische Universitat Wuppertal

arX

iv:2

101.

0178

0v1

[co

nd-m

at.s

tr-e

l] 5

Jan

202

1

Page 2: Introduction to Solid State Physics - arXiv

2

Preface

This script is based on lecture notes prepared for the regular Introduction to TheoreticalSolid State Physics at the University of Wuppertal held by the author in the winter semestersof 2003/04, 2004/05, 2010/11, 2011/12, 2013/14, 2014/15 and in the summer semesters of2006 and 2020. Due to the prevailing Covid 19 pandemic all teaching at the University ofWuppertal in the summer of 2020 went online. In order to support my students with theirhome training programme I decided to typeset at least the present part of my lecture notes.In regular semesters I would have delivered 28-30 lectures, 90 minutes each. Since thebeginning of our semester was delayed due to the pandemic, the number of lectures wasrestricted to 23. For this reason I cut out two lectures on the Hartree-Fock approximationand two lectures usually devoted to recall the formalism of the second quantization. Icondensed the remaining material to fit into the available 23 time slots. The missed outmaterial as well as some of the material of the regular lectures on Advanced TheoreticalSolid State Physics may follow on a later occasion, e.g., after the next pandemic. Most ofthe lectures are complemented with a few intermediate level homework exercises which areconsidered an integral part of the course. These exercises were discussed by the participantsin a separate online exercise session on a weekly basis.

At the university of Wuppertal the introduction to theoretical solid state physics is partof the Master course. Students who attend this course are expected to have successfullypassed the basic courses in theoretical physics (Mechanics, Electrodynamics, QuantumMechanics, Statistical Mechanics) and a course on Advanced Quantum Mechanics.

I would like to thank all those participants of my lectures whose attention and whosequestions helped to improve this manuscript. Particular thanks are due to Saskia Faulmannand Siegfried Spruck who read the entire text and pointed out a number of typos andinaccuracies to me. When I started teaching Theoretical Solid State Physics in 2003, Idevised my lecture after lecture notes of my dear colleague Holger Fehske whom I wouldlike to thank at this point. My own first lecture notes then gradually evolved over the yearsand most probably will continue to evolve in future. I prefer to think of the followingtypeset version as of a snapshot taken in the year of the pandemic 2020.

What is solid state physics?

• Application of what we have learned heretofore (QM + StatMech) to the descriptionof ‘condensed matter’, no new fundamental theory.

• Arguably the most important branch of physics as far as the daily life of non-physicists is concerned, since many important technologies owe their existence ourknowledge of solid state physics. Examples are the semiconductor technology, lasers,magnetic and charge based information storage devices.

• Solid state physics is intellectually most challenging with new questions arising fromexperiments year by year. Solids show us the universe in a nutshell. Their behaviourreaches from single-particle to highly collective. All tools of modern theoreticalphysics are needed for at least an approximate understanding, in particular, QFT,Feynman graphs, high-performance computing, and non-perturbative many-bodytechniques.

Page 3: Introduction to Solid State Physics - arXiv

3

What are the goals of this lecture?

• Introduction to the basic concepts, meaning that the emphasis is, in the first instance,on the single-particle aspects.

• Service for Experimental Solid State Physics.

• Emphasis on the explanation of concepts and basic ideas, not always quantitative,justification of the use of simplified ‘model Hamiltonians’.

• Raise some understanding why many-body physics is mostly phenomenology.

• Convey the following main idea: (Collective) elementary excitations are ‘quasi-particles’ characterized by their dispersion relation p 7→ ε(p) and by certain quan-tum numbers like spin and charge. The most important two are ‘the phonon’ (=quantized lattice vibration) and ‘the electron’ (= quantized charge excitation of thesolid, which has as much to do with the electron of elementary particle physics aswater waves have to do with water).

Page 4: Introduction to Solid State Physics - arXiv

4

Contents

L1 The Hamiltonian of the solid and its eigenvalue problem 81.1 Hamiltonian of the solid . . . . . . . . . . . . . . . . . . . . . . . . . 81.2 Natural units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81.3 An eigenvalue problem for the ions . . . . . . . . . . . . . . . . . . . 9

L2 Born-Oppenheimer approximation 112.1 More scaling arguments . . . . . . . . . . . . . . . . . . . . . . . . . 112.2 Exercise 1. The heavy harmonic oscillator . . . . . . . . . . . . . . . . 122.3 Application to the ionic motion and adiabatic decoupling . . . . . . . . 122.4 Summary and interpretation . . . . . . . . . . . . . . . . . . . . . . . 152.5 Exercise 2. A simple application of Born-Oppenheimer . . . . . . . . . 16

L3 Crystal lattices 163.1 The Bravais lattice . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173.2 The reciprocal lattice . . . . . . . . . . . . . . . . . . . . . . . . . . . 183.3 Properties of the reciprocal lattice . . . . . . . . . . . . . . . . . . . . 183.4 Exercise 3. Real space interpretation of the Miller indices . . . . . . . 193.5 Lattice periodic functions . . . . . . . . . . . . . . . . . . . . . . . . 19

L4 Crystal symmetries 214.1 The crystal lattice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214.2 The Euclidean group . . . . . . . . . . . . . . . . . . . . . . . . . . . 214.3 The symmetry group of a crystal . . . . . . . . . . . . . . . . . . . . . 214.4 The point group of a crystal . . . . . . . . . . . . . . . . . . . . . . . 224.5 Remarks on point groups . . . . . . . . . . . . . . . . . . . . . . . . . 224.6 Classification of all point and space groups . . . . . . . . . . . . . . . 234.7 Exercise 4. Lattice planes in the cubic face-centered lattice . . . . . . . 244.8 Exercise 5. Face centered tetragonal structure . . . . . . . . . . . . . . 24

L5 The action of the Bravais lattice on states 245.1 Shift operators, lattice momentum and Bloch’s theorem . . . . . . . . . 245.2 Periodic boundary conditions . . . . . . . . . . . . . . . . . . . . . . 27

L6 Phonons – spectrum and states 286.1 The Hamiltonian of the lattice vibrations in harmonic approximation . . 286.2 Implications of the translational invariance . . . . . . . . . . . . . . . 296.3 Block diagonalization of the force matrix . . . . . . . . . . . . . . . . 306.4 Transformation of the kinetic energy . . . . . . . . . . . . . . . . . . . 316.5 Properties of the matrix κ(q) . . . . . . . . . . . . . . . . . . . . . . 326.6 Exercise 6. Classical harmonic chain with various boundary conditions 33

L7 Phonons – spectrum and states continued 347.1 Reduction of the Hamiltonian to a diagonal quadratic form . . . . . . . 347.2 Zero modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357.3 Diagonalization of the phonon Hamiltonian . . . . . . . . . . . . . . . 367.4 Creation and annihilation operators in terms of the original displacements 377.5 Construction of the eigenstates . . . . . . . . . . . . . . . . . . . . . . 38

Page 5: Introduction to Solid State Physics - arXiv

5

L8 Phonons – examples and general properties 398.1 Example 1 – the harmonic chain . . . . . . . . . . . . . . . . . . . . . 398.2 Example 2 – diatomic chain with alternating forces . . . . . . . . . . . 408.3 General properties of phonons in a 3d lattice . . . . . . . . . . . . . . 428.4 Exercise 7. Linear chain with a mass defect . . . . . . . . . . . . . . . 448.5 Exercise 8. Mass defect in the thermodynamic limit . . . . . . . . . . 458.6 Exercise 9. Harmonic oscillations of a two-dimensional lattice . . . . . 46

L9 Statistical mechanics of the harmonic crystal 469.1 Partition function and free energy . . . . . . . . . . . . . . . . . . . . 469.2 The Bose-Einstein distribution . . . . . . . . . . . . . . . . . . . . . . 469.3 The density of states . . . . . . . . . . . . . . . . . . . . . . . . . . . 479.4 Van-Hove singularities . . . . . . . . . . . . . . . . . . . . . . . . . . 489.5 Exercise 10. Van-Hove singularities . . . . . . . . . . . . . . . . . . . 499.6 Exercise 11. Density of states in linear chains . . . . . . . . . . . . . . 49

L10 Specific heat of the harmonic crystal 4910.1 Low-temperature specific heat . . . . . . . . . . . . . . . . . . . . . . 5010.2 Internal energy and specify heat at high temperatures . . . . . . . . . . 5110.3 Debye interpolation . . . . . . . . . . . . . . . . . . . . . . . . . . . 5210.4 The anomaly of the harmonic crystal . . . . . . . . . . . . . . . . . . 53

L11 Neutron scattering 5511.1 Thermal neutrons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5511.2 Scattering experiments and cross sections . . . . . . . . . . . . . . . . 5611.3 Cross section and transition rate . . . . . . . . . . . . . . . . . . . . . 5711.4 Form factor and structure factor . . . . . . . . . . . . . . . . . . . . . 5811.5 Rewriting the structure factor . . . . . . . . . . . . . . . . . . . . . . 58

L12 Neutron scattering continued 5912.1 Disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5912.2 Summary of the theory of neutron scattering – general case . . . . . . . 6012.3 Theory of neutron scattering – specialization to crystal structures . . . 6112.4 Dynamic structure factor of the harmonic crystal . . . . . . . . . . . . 6112.5 Expansion into multi-phonon processes . . . . . . . . . . . . . . . . . 6212.6 Coherent elastic neutron scattering . . . . . . . . . . . . . . . . . . . . 6212.7 Interpretation of the Bragg condition . . . . . . . . . . . . . . . . . . 6212.8 Exercise 12. A proof of equation (12.13) . . . . . . . . . . . . . . . . 63

L13 Inelastic neutron scattering 6513.1 Debye-Waller factor . . . . . . . . . . . . . . . . . . . . . . . . . . . 6513.2 Exercise 13. Debye-Waller factor of a cubic lattice . . . . . . . . . . . 6613.3 Coherent inelastic scattering – the single-phonon contribution . . . . . 6613.4 Measuring dispersion relations of phonons . . . . . . . . . . . . . . . 69

L14 Electronic excitations in solids 6914.1 Hamiltonian of the electrons in adiabatic approximation . . . . . . . . 6914.2 Reduction to a single-particle problem . . . . . . . . . . . . . . . . . . 7014.3 Particles in a periodic potential . . . . . . . . . . . . . . . . . . . . . . 72

Page 6: Introduction to Solid State Physics - arXiv

6

14.4 Exercise 14. Kronig-Penney model . . . . . . . . . . . . . . . . . . . 7214.5 Almost free electrons . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

L15 Particles in a periodic potential 7615.1 Degenerate levels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7615.2 Example d = 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7715.3 The tight-binding method . . . . . . . . . . . . . . . . . . . . . . . . 78

L16 Electrons in a periodic potential 8016.1 Method of orthogonalized plane waves . . . . . . . . . . . . . . . . . 8016.2 Other methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8116.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8116.4 Exercise 15. Electronic band structure in one-dimensional solids by WKB 8116.5 The Fermi distribution . . . . . . . . . . . . . . . . . . . . . . . . . . 8316.6 Grand canonical potential of the electron gas . . . . . . . . . . . . . . 8416.7 Fermi energy and Fermi surface . . . . . . . . . . . . . . . . . . . . . 84

L17 Low-temperature specific heat of the electron gas 8517.1 Specific heat of metals . . . . . . . . . . . . . . . . . . . . . . . . . . 8617.2 Specific heat of insulators . . . . . . . . . . . . . . . . . . . . . . . . 88

L18 Electrons in solids – the second quantized picture 9018.1 An auxiliary potential . . . . . . . . . . . . . . . . . . . . . . . . . . 9118.2 Bloch basis and Wannier basis . . . . . . . . . . . . . . . . . . . . . . 9118.3 Hamiltonian in second quantization . . . . . . . . . . . . . . . . . . . 9218.4 Exercise 16. Wannier functions in one dimension . . . . . . . . . . . . 9318.5 Exercise 17. Spinless Fermions on the lattice . . . . . . . . . . . . . . 93

L19 The Hubbard model 9419.1 Motivation and definition . . . . . . . . . . . . . . . . . . . . . . . . . 9419.2 Tight-binding approximation . . . . . . . . . . . . . . . . . . . . . . . 9519.3 Interpretation of the Hubbard Hamiltonian . . . . . . . . . . . . . . . 9619.4 Exercise 18. Peierls phases . . . . . . . . . . . . . . . . . . . . . . . . 99

L20 The strong-coupling limit 10020.1 Degenerate perturbation theory – a reminder . . . . . . . . . . . . . . 10020.2 Application to the Hubbard model at strong coupling . . . . . . . . . . 10220.3 Explicit form of the projection operators . . . . . . . . . . . . . . . . . 10220.4 Application toH2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10220.5 Exercise 19. Strong coupling limit of the Hubbard model . . . . . . . . 103

L21 Heisenberg model and Mott transition 10321.1 Heisenberg Hamiltonian in the language of Fermi operators . . . . . . 10321.2 Heisenberg model in the language of spin operators . . . . . . . . . . . 10421.3 Interpretation of the exchange interaction . . . . . . . . . . . . . . . . 10521.4 Mott transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10721.5 Exercise 20: A short XXX chain . . . . . . . . . . . . . . . . . . . . . 107

Page 7: Introduction to Solid State Physics - arXiv

7

L22 Linear response theory 10822.1 Time evolution of the statistical operator . . . . . . . . . . . . . . . . 10822.2 Time evolution of expectation values . . . . . . . . . . . . . . . . . . 10922.3 Absorption of energy . . . . . . . . . . . . . . . . . . . . . . . . . . . 10922.4 Application to quantum spin chains . . . . . . . . . . . . . . . . . . . 110

L23 Microwave absorption by the Heisenberg-Ising chain 11123.1 The isotropic chain . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11123.2 Resonance shift and line width in the anisotropic case . . . . . . . . . 11223.3 The shifted moments of the dynamic susceptibility . . . . . . . . . . . 11323.4 Exercise 21: Shifted moments . . . . . . . . . . . . . . . . . . . . . . 115

References 116

Page 8: Introduction to Solid State Physics - arXiv

8

L1 The Hamiltonian of the solid and its eigenvalue problem

1.1 Hamiltonian of the solid

Solid state physics is quantum mechanics and statistical mechanics of many (∼ 1023)particles at ‘low energies’ (typically ∼ 1 meV). Relativistic effects (retardation, spin-orbitcoupling, . . . ) can often be neglected in the explanation of phenomena in solids (apartfrom the fact the spin is a relativistic effect). The large number of particles Np is best takencare of by considering the thermodynamic limit Np → ∞ which typically brings aboutsimplifications of the theory.

In solid state physics not only the electrons composing the solid but also the ions areregarded as elementary. The only relevant interaction is the Coulomb interaction. Thus, theHamiltonian of the solid is the Hamiltonian of non-relativistic ions and electrons interactingvia Coulomb interaction,

H =L∑j=1

‖Pj‖22Mj

+∑

1≤j<k≤L

ZjZke2

‖Rj −Rk‖← ions, charge Zje, mass Mj

+N∑j=1

‖pj‖22m

+∑

1≤j<k≤N

e2

‖rj − rk‖← electrons, charge −e, mass m

−L∑j=1

N∑k=1

Zje2

‖Rj − rk‖. ← ion-electron interaction (1.1)

↑ kinetic energy ↑ Coulomb interaction

Here L is the number of ions and N is the number of electrons. The Hamiltonian ofthe solid is the same as in atomic and molecular physics. The only difference is in thenumber of constituents. A solid is a very large molecule. The numbers that separate thesub-disciplines are

• L = 1 atomic physics

• L ∼ 100 molecular physics

• L ∼ 1023 solid state physics

There is no sharp distinction between a molecule and a solid. In between molecules andsolids are macro molecules. DNA, for instance, is composed of about 2 · 1011 atoms.

1.2 Natural units

A first understanding of the length and energy scales in solids comes from a dimensionalanalysis that allows us to introduce natural units. Asking, when the kinetic energy ofan electron associated with a given wavelength is of the same order of magnitude as theCoulomb energy of two electrons one wavelength apart from each other,

~2

2m`2∼ E ∼ e2

`, (1.2)

Page 9: Introduction to Solid State Physics - arXiv

9

we recover the typical length scale of atomic physics

` ∼ ~2

me2= a0 = 0,529 · 10−10 m ∼ 1

20nm , (1.3)

the Bohr radius. The Coulomb energy of two electrons at this distance,

E =e2

a0=me4

~2= 2Ry = 27,2 eV , (1.4)

is twice the ionization energy of a hydrogen atom (which is one Rydberg (1 Ry)) and thusa typical atomic binding energy.

Measuring all lengths in units of the Bohr radius (i.e., replacing rj/a0 → rj , Rj/a0 →Rj) and the energy in units of 2 Ry (by replacing H

/(me4/~2) → H) we obtain the

Hamiltonian of the solid in natural units,

H = −1

2

L∑j=1

m

Mj∂2Rαj

+Hel(R) , (1.5)

where

Hel(R) =∑

1≤j<k≤L

ZjZk‖Rj −Rk‖

− 1

2

N∑j=1

∂2rαj

+∑

1≤j<k≤N

1

‖rj − rk‖−

L∑j=1

N∑k=1

Zj‖Rj − rk‖

. (1.6)

Here we have used the summation convention with respect to Greek indices and the furtherconvention that R = (R1, . . . ,RL). Similarly, we shall write r = (r1, . . . , rN ).

Our scale analysis shows us that the ionisation numbers Zj and the mass ratios m/Mj

are the only parameters of the system. The variation of these pure numbers is responsiblefor the rich phenomenology of molecules and solids and, in fact, of the world around us aswe perceive it with our senses.

Note that the mass ratiom

Mj∼ 10−4 (1.7)

in (1.5) is a small parameter. This fact turns out to be of fundamental importance for thetheory of solids and determines much of the structure of the world around us.

1.3 An eigenvalue problem for the ions

If we naively send all masses Mj in (1.5) to infinity, the kinetic energy of the ions goes tozero, the ions stop moving. The corresponding ionic parts of the eigenfunctions separatemultiplicatively and become products of delta functions of the form δ(Rj −R

(0)j ). If the

masses are large but finite, the ions will still move, but slower than the electrons. Theirwavefunctions will not be delta functions, but typically more localized than those of theelectrons.

Our favorite classical example system of interacting point particles of very differentmasses is the planetary system. The ratio of the earth mass m to the sun mass M , forinstance, is about m/M = 1/3× 10−5. Earth and sun exert equal but oppositely directed

Page 10: Introduction to Solid State Physics - arXiv

10

forces ±F onto each other, MX = F = −mx, if X and x are the position vectors ofsun and earth, respectively. This means that sun experiences a much smaller averageacceleration than the earth. Consequentially, as compared to the sun, the earth moves muchfaster and has a much larger orbit around the center of mass of the sun-earth system. Inthis case, as the mass ratio is so small, the center of mass lies inside the sun. Hence, toa very good approximation, the earth moves around the sun and follows it along its waythrough the universe.

In a similar way we expect the electrons in a solid to follow the slower motion of themore massive ions. Translated into the language of quantum mechanics we expect that thejoint motion of electrons and ions can be approximately described by a product of an ionicwave function times an electronic wave function calculated for fixed positions of the ions.The latter would be interpreted as a conditional probability amplitude for the electronsgiven the positions of the ions. The product structure would mimic the fact in probabilitytheory that the joint probability p(A ∩ B) of two events A and B (corresponding to thewave function of electrons and ions) is equal to p(A|B)p(B), where p(B) is the probabilityof B (corresponding to the ions) and p(A|B) is the conditional probability of A given B(corresponding to the electronic wave function for fixed positions of the ions).

We shall try to work out this idea more formally. Let us start with the ‘electroniceigenvalue problem’

Hel(R)ϕ(r|R) = ε(R)ϕ(r|R) (1.8)

which depends parametrically on the positions R of all ions. For every R the eigenstatesϕn(r|R)

∣∣n∈N of Hel(R) corresponding to the eigenvalues εn(R) form a basis of the

electronic Hilbert space. Hence, every solution Ψ(r,R) of the full eigenvalue problemHΨ = EΨ can be expanded in terms of the ϕn,

Ψ(r,R) =∑n∈N

ϕn(r|R)φn(R) . (1.9)

Assuming the ϕn to be known we want to derive an eigenvalue problem for the φnwhich will later be interpreted as the ionic wave functions. For this purpose we insert (1.9)into the full eigenvalue problem and write the result as

(H − E)Ψ(r,R) =∑m∈N

ϕm(r|R)

[−1

2

L∑j=1

m

Mj∂2Rαj

+ εm(R)− E]φm(R)

− 1

2

L∑j=1

m

Mj

[2(∂Rαj ϕm

)(r|R)∂Rαj +

(∂2Rαjϕm)(r|R)

]φm(R)

= 0 . (1.10)

From here we obtain an equation for the φn upon multiplication by ϕ∗n(r|R) and integrationover r, [

−1

2

L∑j=1

m

Mj∂2Rαj

+ εn(R)− E]φn(R) =

∑m∈N

Cnm(R)φm(R) , (1.11)

where

Cnm(R) = Anm(R) +Bnm(R) , (1.12a)

Page 11: Introduction to Solid State Physics - arXiv

11

Anm(R) =L∑j=1

m

Mj

∫d3Nr ϕ∗n(r|R)

(∂Rαj ϕm

)(r|R)∂Rαj , (1.12b)

Bnm(R) =1

2

L∑j=1

m

Mj

∫d3Nr ϕ∗n(r|R)

(∂2Rαjϕm)(r|R) . (1.12c)

If there is no external magnetic field we may assume that the ϕn are real. Then, using that∫d3Nr ϕn(r|R)

(∂Rαj ϕn

)(r|R) =

1

2∂Rαj

∫d3Nr ϕ2

n(r|R) = 0 , (1.13)

we see that Ann(R) = 0. Introducing the notation

T = −1

2

L∑j=1

m

Mj∂2Rαj

(1.14)

we can therefore rewrite (1.11) in the form[T + εn(R)−Bnn(R)− E

]φn(R) =

∑m∈N,m 6=n

Cnm(R)φm(R) . (1.15)

This equation still contains the full information about the ion system. The ions are nowcoupled through their Coulomb interaction (contained in εn(R)) and through the electrons,whose degrees of freedom have been formally integrated out. Equation (1.15) will turn outto be an appropriate starting point for a perturbative analysis of the ion system with m/Mj

taken as a small parameter.

L2 Born-Oppenheimer approximation

2.1 More scaling arguments

Classically we have a general idea, how the time scale, e.g. for the motion of a particle ofmass m with one degree of freedom in a potential V , depends on the mass. The Lagrangianof such a system is

L =M(∂tx)2

2− V (x) =

m

2

(∂ t√

M/m

x)2− V (x) . (2.1)

Denoting by xµ(t) a trajectory of a particle with mass µ, we see that

xM (t) = xm

(t√M/m

)(2.2)

is a trajectory of a particle of mass M . The motion slows down if M > m. Heavierparticles (of the same energy) move slower. In particular, oscillators oscillate with lowerfrequency if their mass is enlarged.

In quantum mechanics we have to consider the stationary Schrodinger equation, a time-independent problem. Hence, we are rather interested in how the spatial behaviour of theeigenfunctions varies with mass. For the bounded motion around an equilibrium position

Page 12: Introduction to Solid State Physics - arXiv

12

described by a quadratic minimum of the potential we may recourse to the harmonicoscillator

H =p2

2M+Kx2

2. (2.3)

Comparing kinetic and potential energy in a similar scaling argument as in (1.2),

~2

2m`2∼ E ∼ K`2

2, (2.4)

we find the intrinsic length scale

` =

(~2

mK

) 14

. (2.5)

Comparing the extension L of an eigenfunction of a heavy particle of mass M with theextension ` of a lighter particle of mass m we obtain a ratio of

L

`=

(m

M

) 14

. (2.6)

This means that the larger the mass the more localized becomes the wave function. On theother hand, in a more localized wave function the particle is closer to the origin and theharmonic approximation is better justified.

2.2 Exercise 1. The heavy harmonic oscillator

The same conclusions as above can be drawn from the solution of the eigenvalue problemof the harmonic oscillator (2.3).

(i) Determine the full width at half height of the ground state wave function of theharmonic oscillator. How does it depend on the mass of the oscillator?

(ii) Consider the oscillator with a quartic correction term

H =p2

2m+Kx2

2+K ′x4

4!, (2.7)

where K ′ > 0. Show that the correction can be taken into account perturbativelyif the mass is large. For this purpose use ` as defined in (2.5) as a small parameter.Calculate the correction to the ground state energy in first oder perturbation theory.How does it depend on the mass of the oscillator?

2.3 Application to the ionic motion and adiabatic decoupling

Let M be a typical ion mass, e.g. the arithmetic average of all ion masses M =⟨Mj

⟩.

Then

κ =

(m

M

) 14

(2.8)

is an intrinsic length parameter for the bounded motion of the ions.We expect that the eigenvalue problem (1.15) has solutions which, on the scale of the

electronic wave functions, are strongly localized around certain equilibrium positions R(0).

Page 13: Introduction to Solid State Physics - arXiv

13

To take account of this expectation we introduce new coordinates u on the scale of theelectronic wave functions and relative to this equilibrium position, setting

R = R(0) + κu . (2.9)

For the wave functions we shall write

φn(u) = φn(R) . (2.10)

Using the new coordinates (2.9) we can make the κ dependence of the operators in (1.15)explicit. For this purpose we define the rescaled operators

Tu = −1

2

L∑j=1

M

Mj∂2uαj, (2.11a)

Anm(R) =L∑j=1

M

Mj

∫d3Nr ϕ∗n(r|R)

(∂Rαj ϕm

)(r|R)∂uαj , (2.11b)

Bnm(R) =1

2

L∑j=1

M

Mj

∫d3Nr ϕ∗n(r|R)

(∂2Rαjϕm)(r|R) , (2.11c)

which remain finite for κ → 0. With these definitions the eigenvalue problem (1.15)assumes the form[

κ2Tu + εn(R(0) + κu)− κ4Bnn(R(0) + κu)− E]φn(u)

= κ3∑

m∈N,m 6=n

(Anm(R(0) + κu) + κBnm(R(0) + κu)

)φm(u) . (2.12)

Recall that we assume that the wave functions φn are strongly localized around R(0),implying that the redefined functions φn are strongly localized around u = 0. For thisreason it makes sense to expand the operators in (2.12) which act on the functions φn in aTaylor series in κu and to solve the resulting eigenvalue problem perturbatively for small κ.The latter means to look for solutions in the form of formal series in κ,

E = E(0) + κE(1) + . . . , (2.13a)

φn(u) = φ(0)n (u) + κφ(1)

n (u) + . . . (2.13b)

Inserting these perturbation series into (2.12) and performing the Taylor expansion weobtain(

κ2Tu + εn(R(0)) + κε(1)n (u) + κ2ε(2)

n (u) + κ3ε(3)n (u) + κ4ε(4)

n (u)+

· · · − κ4Bnn(R(0))− · · · − E(0) − κE(1) − κ2E(2) − κ3E(3) − κ4E(4) − . . .)

×(φ(0)n (u) + κφ(1)

n (u) + κ2φ(2)n (u) + κ3φ(3)

n (u) + . . .)

= κ3∑

m∈N,m 6=n

(Anm(R(0)) + . . .

)(φ(0)m (u) + κφ(1)

m (u) + κ2φ(2)m (u) + . . .

). (2.14)

Page 14: Introduction to Solid State Physics - arXiv

14

Here we compare the coefficients in front of the powers of κ order by order. To theorder κ0 we obtain

(εn(R(0))− E(0))φ(0)n (u) = 0 . (2.15)

Let φ(0)` (u) 6= 0 for some ` ∈ N, ε`(R(0)) non-degenerate. Then

E(0) = ε`(R(0)) , φ(0)

n (u) = 0 ∀n 6= ` . (2.16)

For n 6= ` we conclude for the higher orders in κ that

O(κ) :(εn(R(0))− ε`(R(0))

)φ(1)n (u) = 0 ⇒ φ(1)

n (u) = 0 , (2.17a)

O(κ2) :(εn(R(0))− ε`(R(0))

)φ(2)n (u) = 0 ⇒ φ(2)

n (u) = 0 , (2.17b)

O(κ3) :(εn(R(0))− ε`(R(0))

)φ(3)n (u) = An`(R

(0))φ(0)` (u)

⇒ φ(3)n (u) =

An`(R(0))φ

(0)` (u)

εn(R(0))− ε`(R(0)). (2.17c)

The latter equation means that at order κ3 ∼ 1/1000 the wave function Ψ(r|R) (1.9) ofthe coupled electron ion system ceases to be a simple product of two factors.

For n = ` we find at order κ that

(ε(1)(u)− E(1)

(0)` (u) = 0 ⇒ E(1) = ε

(1)` (u) =

L∑j=1

∂ε`(R(0))

∂Rαjuαj . (2.18)

But E(1) must be independent of u which can only hold if

∂ε`(R(0))

∂Rαj= 0 . (2.19)

Then, necessarily,E(1) = 0 . (2.20)

For higher orders of κ we remain with the equation(Tu+ε

(2)` (u) + κε

(3)` (u) + κ2ε

(4)` (u)− κ2Bnn(R(0))− E(2) − κE(3) − κ2E(4)

)×(φ

(0)` (u) + κφ

(1)` (u) + κ2φ

(2)` (u)

)= κ4

∑m∈N,m 6=`

A`m(R(0))φ(3)m (u) + O(κ5)

= κ4∑

m∈N,m 6=`

A`m(R(0))Am`(R(0))φ

(0)` (u)

εm(R(0))− ε`(R(0))+ O(κ5) . (2.21)

Conceiving this as an equation up to the order κ2, we see that the right hand side can beconsistently neglected.

Page 15: Introduction to Solid State Physics - arXiv

15

2.4 Summary and interpretation

(i) Within the scheme of the above perturbation theory the solutions φn,`, En,` of theeigenvalue problem (

T + εn(R)−Bnn(R)− E)φ(R) = 0 (2.22)

consistently determine the eigenfunctions of the solid in product form

Ψn,`(r,R) = ϕn(r|R)φn,`(R) (2.23)

up to the fourth order expansion of εn(R) in κ and up to the zeroth order expansionof Bnn(R) in κ. ϕn(r|R) is interpreted as a conditional probability amplitude andφn,`(R) as an ionic wave function. The product form then means that the electronsfollow the motion of the ions.

(ii) The approximation (2.22), (2.23) is called the Born-Oppenheimer approximationand originated in molecular physics [2, 3]. Another common name is ‘the adiabaticapproximation’.

(iii) Within the Born-Oppenheimer approximation the equilibrium positions of the ionsare determined by the condition

∂εn(R(0))

∂Rαj= 0 . (2.24)

In solids the ions arrange themselves in regular lattices.

(iv) Consistently up to the second order in κ the eigenvalue problem (2.22) takes theform (

Tu + ε(2)n (u)− E

)φ(0)(u) = 0 (2.25)

where

ε(2)n (u) =

1

2

L∑j,k=1

∂2εn(R(0))

∂Rαj ∂Rβk

uαj uβk . (2.26)

This is called the harmonic approximation. Since only φ(0) is taken into account inthe harmonic approximation, it is consistent to consider the electronic wave functiononly to lowest order ϕ(r|R(0)) as well. In this approximation the electrons move ina static lattice determined by the equilibrium positions of the ions (like the earth wasapproximately moving around a static sun in our entrance example).

(v) Notice that, in spite of the ratio of electron to ion mass being very small, our actualexpansion parameter κ ∼ 1/10 is only moderately small.

(vi) Our perturbative analysis becomes questionable, if the electronic levels are degener-ate or close to degenerate (it is well justified for the electronic ground state of aninsulator, but problematic for a metal).

(vii) Many properties of solids can be understood qualitatively and often also quantita-tively within the harmonic approximation. Most of the treatment of solids in thislecture will be based on it. It explains e.g. the scattering of light or neutrons, thepropagation of sound and the specific heat.

Page 16: Introduction to Solid State Physics - arXiv

16

(viii) An example of an effect which cannot be explained within the harmonic approxima-tion is the thermal extension of a solid. It is a higher order effect and therefore small.Still it can be understood within the adiabatic approximation if we proceed to thefourth order expansion of εn(R).

2.5 Exercise 2. A simple application of Born-Oppenheimer

A heavy particle (M ) and a light particle (mM ) move inside an infinitely high potentialwell of width L. The particles experience an attractive interaction described by the potentialW (r −R) = −λ · δ(r −R), where λ > 0 and R and r are the positions of the heavy andlight particle, respectively.

Calculate the spectrum of the corresponding Hamiltonian

H = − ~2

2m

∂2

∂r2− ~2

2M

∂2

∂R2− λ · δ(r −R) (2.27)

in Born-Oppenheimer approximation:

(i) In step one transform the Hamiltonian into a dimensionless form and neglect thekinetic energy of the heavy particle. Determine the (unnormalized) eigenfunctionsand an equation for the energy eigenvalues ε(R) of the light particle. It may turn outto be useful to distinguish the cases ε < 0, ε = 0 and ε > 0. Recall the relation

∂ϕ

∂r(R+ 0)− ∂ϕ

∂r(R− 0) = −2mλ

~2ϕ(R) , (2.28)

for the jump of the derivative of the wave function caused by the δ-function.

(ii) In step two sketch the energy ε(R) of the light particle for the ground state and forthe first excited state as a function of the position of the heavy particle. What is themeaning of ε(R) and how should we include (qualitatively) the energy levels of theheavy particle into our picture?

L3 Crystal lattices

Within the adiabatic approximation the electrons and ions form bound states, molecules orsolids, in which the ions oscillate around the minima R(0) of the effective potentials εn(R)determined by the mutual Coulomb interaction of the ions and the energy eigenvalues ofthe electrons for fixed ion positions. We derived this statement in the previous lecturepresupposing that such minima exist. The argument would have been more convincing ifwe would have been able to prove the existence of minima starting with the Hamiltonian(1.1). Such undertaking seems to be out of reach with our current methods. Still, in nature,atoms always form bound states at low enough temperatures (the only notable exceptionbeing helium which condenses into a ‘super fluid’ before solidification). If many atoms areput together in stoichiometric ratios they form periodic structures called crystals. Crystalsare formed, since (i) two atoms prefer a certain binding length and (ii) no direction ispreferred in the large (this is very much like packing balls into a box and carefully shakingit such that the balls find their equilibrium positions).

In this and in the following lecture we introduce the terminology and certain mathe-matical structures needed for the description of crystal lattices.

Page 17: Introduction to Solid State Physics - arXiv

17

3.1 The Bravais lattice

The most important structural feature of a crystal is that it can be thought of as beinggenerated by periodic repetitions of a finite elementary structure, the unit cell, in threespace directions.

Let a1,a2,a3 ∈ R3, 〈a1,a2 × a3〉 = det(a1,a2,a3) 6= 0. Then

B =x ∈ R3

∣∣x = `1a1 + `2a2 + `3a3, `j ∈ Z, (3.1)

conceived as an Abelian group, is called a (3d) Bravais lattice. B is called the Bravaislattice of a crystal, if B is the group of all translations which map the (infinitely extended)crystal onto itself. Similarly we can define Bravais lattices in any number of dimensions.In the examples and exercises we will frequently work with d = 1 and d = 2.

Remark. We shall develop part of the theory of infinite crystals which is expected to givea realistic description, if the ratio of the number of ions at the surface to the number of ionsin the bulk is small. For a typical macroscopic total number of L ∼ 1024 ions, of the orderof L2/3 ∼ 1016 of them are at the surface, and the ratio is ∼ 10−8.

Bravais lattice vectors aj , j = 1, 2, 3, that generate the Bravais lattice as an Abeliangroup with respect to vector addition are called primitive (lattice) vectors. They are notunique.

Lemma 1. Let aj , j = 1, 2, 3, a set of primitive vectors of a Bravais lattice. Then

bi =

3∑j=1

mijaj , i = 1, 2, 3, primitive ⇔ mij ∈ Z and |detm| = 1 .

Proof. ⇒: primitive⇒ m−1ij ∈ Z⇒ detm−1 = 1/ detm ∈ Z ⇒ (since detm ∈ Z):

detm = ±1.⇐: Cramer’s rule and detm = ±1 ⇒ m−1

ij ∈ Z ⇒ primitive.

Since | detm| = 1 it follows that

Vu = | det(b1,b2,b3)| = |det(a1,a2,a3)| , (3.2)

the volume of the parallelepiped spanned by a set of primitive vectors, is independent oftheir choice.

By definition, the unit cell of a crystal with Bravais lattice B is a simply connectedfinite volume (of size Vu) which covers R3 through translation by B. We often use aspecial unit cell, the Wigner-Seitz cell, which reflects the symmetries of the Bravais lattice,and is defined as

W =x ∈ R3

∣∣‖x‖ ≤ ‖x− g‖ ∀g ∈ B \ 0. (3.3)

Geometrically this is the set of points in R3 for which the closest Bravais lattice point isthe origin. It can be constructed by drawing lines from the origin to the neighbouring sitesin the Bravais lattice and erecting the perpendicular bisectors on these lines.

Exercise: Draw the Wigner-Seitz cells for 2d Bravais lattices composed of equilateralsquares and triangles.

Page 18: Introduction to Solid State Physics - arXiv

18

3.2 The reciprocal lattice

The reciprocal lattice which we shall define now is a lattice which, in a sense, is dual tothe Bravais lattice. It is one of the most important notions in solid state physics and willaccompany us throughout this lecture.

Given a Bravais lattice B and a set of primitive vectors aj , j = 1, 2, 3, generating it,we would like to know how to decompose any x ∈ R3 with respect to the aj , i.e., wewould like to know its coordinates with respect to the basis a1,a2,a3. Suppose that bj ,j = 1, 2, 3, exist such that

〈aj ,bk〉 = 2πδjk . (3.4)

Then

2πxj = 〈x,bj〉 , x =〈x,bj〉aj

2π, (3.5)

where we have employed the summation convention in the second equation. If (3.4) issatisfied, then aj3j=1 and bj3j=1 are called reciprocal (dual) to each other. The lattice

B =x ∈ R3

∣∣x = `1b1 + `2b2 + `3b3, `j ∈ Z

(3.6)

is called the reciprocal lattice (associated with the Bravais lattice B).It is easy to solve (3.4) for the bj . For this purpose we rewrite it in matrix form

(a1,a2,a3)t(b1,b2,b3) = 2πI3 , (3.7)

where I3 is the 3× 3 unit matrix, and use Cramer’s rule,

(b1,b2,b3) = 2π((a1,a2,a3)−1

)t=

Vu(a2 × a3,a3 × a1,a1 × a2) . (3.8)

Without restriction of generality we have assumed here that det(a1,a2,a3) > 0.

3.3 Properties of the reciprocal lattice

(i) Involutivity. Equation (3.4) implies that B = B. The reciprocal of the reciprocallattice is the original Bravais lattice.

(ii) Brillouin zone. The Wigner-Seitz cell of the reciprocal lattice is called the (first)Brillouin zone. It plays an important role in solid state physics.

(iii) Volume of the Brillouin zone. Fix a set of primitive reciprocal lattice vectors bj anddenote the volume of the parallelepiped spanned by these vectors by

VR = det(b1,b2,b3) . (3.9)

Taking the determinant on the left and right hand side of the first equation (3.7) wesee that

VR =(2π)3

Vu(3.10)

is the volume of the unit cells of the reciprocal lattice, which is the same as thevolume of the Brillouin zone.

Page 19: Introduction to Solid State Physics - arXiv

19

(iv) Lattice planes. A lattice plane S ⊂ R3 associated with a Bravais lattice B is a planefor which ∃ a1,a2,a3 ∈ B primitive, ∃ ` ∈ Z such that

x`;m,n = `a1 +ma2 + na3 ∈ S , ∀m,n ∈ Z . (3.11)

Let b1 = 2π(a2 × a3)/Vu the reciprocal to a1. Then

S =x ∈ R3

∣∣〈x,b1〉 = 2π`. (3.12)

Thus, for every lattice plane there is a primitive reciprocal lattice vector b1 ∈ Band an ` ∈ Z such that (3.12) holds. If we vary ` in (3.12) we obtain a family ofequidistant lattice planes. Conversely, given any primitive vector b1 ∈ B and any` ∈ Z, the plane S defined by (3.12) is a lattice plane. Thus, families of lattice planesare in one-to-one correspondence with primitive vectors of the reciprocal lattice.

(v) Miller indices. Lattice planes play an important role in the spectroscopy of solids,We shall see in the course of this lecture that waves impinging on a crystal arereflected as if they were reflected by families of lattice planes. In spectroscopy thefamilies of lattice planes are usually labeled by the so-called Miller indices definedrelative to a fixed triple b1,b2,b3 of primitive vectors of the reciprocal lattice.According to lemma 1 every primitive vector b ∈ B can be uniquely presented asan integer linear combination

b = m1b1 +m2b2 +m3b3 , mj ∈ Z . (3.13)

The triple (m1,m2,m3) is then the Miller index of the family of lattice planesassociated with b. Accordingly, one speaks of the (m1,m2,m3) plane (e.g., of the(1,0,0) plane). By convention, a bar is used instead of a minus sign, such that e.g.(1,−1,0) = (1, 1, 0). As a reference triple b1,b2,b3 one usually uses primitivevectors of minimal possible length.

3.4 Exercise 3. Real space interpretation of the Miller indices

Let mj 6= 0, j = 1, 2, 3, the Miller indices of a family of lattice planes. Convince yourselfthat they indicate in which three points

xj =ajmj

, j = 1, 2, 3 , (3.14)

the lines along the aj directions cut the lattice plane with ` = 1 in (3.12). Note that in theexcluded cases with one or two of the mj being equal to zero there is no intersection in thecorresponding direction. The lines are parallel to the lattice plane.

3.5 Lattice periodic functions

Many quantities that characterize the state of a crystal have the same periodicity as thecrystal itself. For this reason we will often have to deal with periodic functions whoseperiods are the primitive vectors of a Bravais lattice. These are often conveniently describedby their Fourier series.

Let us recall Fourier series in one spatial dimension. In this case the unit cell isnecessarily an interval [0, a], where a > 0 is called the lattice spacing or lattice constant,

Page 20: Introduction to Solid State Physics - arXiv

20

and the primitive vector is equal to a. A lattice periodic function f : R→ C is a functionsatisfying

f(x) = f(x+ a) ∀x ∈ R . (3.15)

A natural basis for the expansion of such functions can be constructed as follows. Letϕ ∈ [0, 2π) and

z = eiϕ = ei(ϕ+2π) = ei 2πa

(ϕa2π

+a) . (3.16)

Setting

x =ϕa

2π, b =

a, km = mb (3.17)

we see thatab = 2π (3.18)

implying that b generates the reciprocal lattice, and that the functions

zm = eikmx , m ∈ Z , (3.19)

are linear independent and periodic with period a.If the series

f(x) =∑m∈Z

Akm eikmx (3.20)

converges uniformly, f is periodic with period a and continuous, and

1

a

∫ a

0dx e−iknx f(x) =

∑m∈Z

Akm1

a

∫ a

0dx ei(km−kn)x = Akn . (3.21)

By means of the latter formula we can associate a sequence of Fourier coefficients(Akn)n∈Z and a Fourier series (3.20) with every complex valued function f that is in-tegrable on [0, a]. In the following we will understand Fourier series in such a formalsense. But when we will be dealing with concrete examples, we will have in mind that theconvergence properties of Fourier series are a delicate matter and that, in general, neitheruniform nor pointwise convergence is guaranteed.

In order to construct Fourier series that have the periodicity defined by a Bravaislattice B, we fix a set of primitive vectors a1,a2,a3 and a set of reciprocal vectorsb1,b2,b3 satisfying (3.4). Then

zj = ei〈bj ,x〉 = ei〈bj ,x+ak〉 , j, k = 1, 2, 3 , (3.22)

and the monomials

z`1zm2 z

n3 = ei〈`b1+mb2+nb3,x〉 , `,m, n ∈ Z , (3.23)

are linear independent and periodic with all a ∈ B being periods. In analogy with the 1dcase we may define k`mn = `b1 +mb2 + nb3 and the Fourier series

f(x) =∑

`,m,n∈ZAk`mn ei〈k`mn,x〉 =

∑k∈B

Ak ei〈k,x〉 , (3.24)

whereAk =

1

Vu

∫U

d3x e−i〈k,x〉 f(x) (3.25)

and the integral is over a unit cell U of volume Vu.With these remarks on Fourier series we have obtained an interpretation of the recipro-

cal lattice. The reciprocal lattice is a lattice in Fourier space dual to the real space Bravaislattice.

Page 21: Introduction to Solid State Physics - arXiv

21

L4 Crystal symmetries

4.1 The crystal lattice

We define a (physical) crystal lattice as the set M of the positions (measured as expectationvalues) of the ions of a solid in its ground state. The translational symmetry of the crystal(lattice) is described by the corresponding Bravais lattice. In general there are several ionsin a unit cell of the Bravais lattice. The positions of the ions inside a unit cell determine theso-called lattice basis. If every unit cell contains only one ion, the crystal is called simple.In this case it can be identified with its Bravais lattice. A crystal that is not simple is calleda crystal with basis.

4.2 The Euclidean group

The Euclidean group is the group of all maps R3 7→ R3 which leave the distance be-tween any two arbitrary points invariant. It consists of all pairs (A,a) of orthogonaltransformations A ∈ O(3) and translations a ∈ R3,

(A,a)x = Ax + a . (4.1)

Then

(A,a)(B,b)x = (A,a)(Bx + b) = ABx +Ab + a = (AB,Ab + a)x , (4.2)

implying that two group elements are multiplied according to the rule

(A,a)(B,b) = (AB,Ab + a) . (4.3)

4.3 The symmetry group of a crystal

The symmetry group R of a crystal M is defined as

R =

(A,a) ∈ E(3)∣∣(A,a)M = M

, (4.4)

i.e., as the subgroup of E(3) that leaves M invariant. If we identify b with (id,b) forevery b ∈ B, the Bravais lattice of M , we see that B is a subgroup of R (since it leavesM invariant and is a group).

Remark. (id,a) ∈ R implies that a ∈ B, but (A,a) ∈ R, A 6= id does not imply thata ∈ B. In crystals with basis smaller translations may occur which are parts of so-calledglide reflections or screw rotations. For an example see Figure 1.

Lemma 2. Let R be the symmetry group of a crystal with Bravais lattice B. Then B ⊂ Ris a normal subgroup of R.

Proof. Let b ∈ B, (A,a) ∈ R. ⇒ (A,a)(id,b) = (A,Ab + a) = (id, Ab)(A,a) , ⇒(id, Ab)M = M ⇒ Ab ∈ B. Thus, (A,a)(id,b)(A,a)−1 ∈ (id, B) for every b ∈ B,meaning that the Bravais lattice is an invariant (= normal) subgroup of R.

Page 22: Introduction to Solid State Physics - arXiv

22

a

reflect

Figure 1: 2d example of a symmetry involving a translation which is not in the Bravaislattice. A glide reflection: glide by a and reflect at the dashed line. Clearly, reflectingwithout gliding is not a symmetry of the sketched point configuration, but leaves its Bravaislattice invariant.

4.4 The point group of a crystal

For a crystal M with symmetry group R define

R0 =A ∈ O(3)

∣∣∃ g ∈ R s. th. g = (A,a)

(4.5)

the set of all ‘O(3) parts’ of R.

Lemma 3. R0 is a subgroup of O(3), the so-called point group of the crystal.

Proof. (i) A,B ∈ R0 ⇒ ∃a,b ∈ R3 s. th. (A,a), (B,b) ∈ R ⇒ (AB,Ab + a) ∈R ⇒ AB ∈ R0. (ii) (A,a)−1 = (A−1,−A−1a) ∈ R ⇒ A−1 ∈ R0. (iii) id ∈ R0.(i)-(iii)⇒ R0 is a group.

Remark. In general, R0M 6= M , i.e. the point group of M does not necessarily leave Minvariant.

We have seen in the proof of lemma 2 that b ∈ B, (A,a) ∈ R ⇒ Ab ∈ B. Since,A ∈ R0, it follows that R0B ⊂ B, B is invariant under R0. This has two importantimplications:

(i) The set of all point groups must be restricted, since not all subgroups of O(3) canleave a Bravais lattice invariant.

(ii) Bravais lattices can be classified according to the point groups which leave theminvariant.

4.5 Remarks on point groups

(i) All point groups are subgroups of O(3) by construction. Thus, A ∈ R0 ⇒detA = ±1. If detA = 1 ∀A ∈ R0, then R0 is called a point group of the firstkind, otherwise a point group of the second kind. Point groups of the first kindconsist of only rotations.

Page 23: Introduction to Solid State Physics - arXiv

23

(ii) The inversion i ∈ O(3) is defined by ix = −x ∀x ∈ R3. ⇒ det i = −1. Forpoint groups of the second kind we distinguish point groups containing i from pointgroups not containing i.

(iii) Since b ∈ B ⇒ −b ∈ B for every Bravais lattice vector, the symmetry groups ofthe Bravais lattices are point groups of the second kind which do contain i.

(iv) Which rotations are possible? The fact that any point group must leave a Bravaislattice invariant restricts the possible rotation angles. It means that every rotationD ∈ R0 must map primitive vectors on vectors in B,

Daj = mjkak , mjk ∈ Z . (4.6)

Define a basis transformation M in R3 by

Maj = ej , ⇒ MDM−1ej = mjkek ,

⇒ trD = trMDM−1 = mjk〈ej , ek〉 = mjj ∈ Z . (4.7)

Recall how the rotation angle is related to the trace of a rotation matrix. The trace isinvariant under coordinate transformation. Hence, we may calculate it in a coordinatesystem in which the axis of rotation is the z-axis. Denoting the rotation angle by ϕwe obtain

trD = tr

cos(ϕ) sin(ϕ)− sin(ϕ) cos(ϕ)

1

= 1 + 2 cos(ϕ) . (4.8)

Combining (4.7) and (4.8) we conclude that allowed angles ϕ must satisfy thecondition 2 cos(ϕ) ∈ Z, or

cos(ϕ) = 0,±1

2,±1 . (4.9)

Thus, the only admissible values of ϕ ∈ [0,2π) are

ϕ = 0,π

3,π

2,2π

3, π . (4.10)

The corresponding rotation axes are called 6-fold, 4-fold, 3-fold, 2-fold.

(v) Point groups do not only contain only rotation axes of finite order, they are alsofinite groups. Comparison with the known finite subgroups of SO(3) leaves 11point groups of the first kind compatible with (4.10). From these we can constructaltogether 32 point groups. Their number is, in particular, finite.

4.6 Classification of all point and space groups

The full symmetry groups R of crystals M are discrete subgroups of E(3) which contain aBravais latticeB as a normal subgroup. Their number is finite as well. In mathematics (andcrystallography) such groups are called space groups. They have been completely classifiedand can be described by symmetry elements like rotations, reflections, glide reflectionsand screw rotations. The table gives an overview over the group theoretic classification ofthe Bravais lattices and crystals.

Page 24: Introduction to Solid State Physics - arXiv

24

Bravais lattices crystals

point groups 7 crystal systems (4 in 2d) 32 crystal classes (13 in 2d)

space groups 14 Bravais classes (5 in 2d) 230 (17 in 2d)

It is very instructive to have a look at least at the pictorial representations of the crystalsystems and Bravais classes. They are shown in Figure 2.

4.7 Exercise 4. Lattice planes in the cubic face-centered lattice

As we have seen in section 3.3 all families of lattice planes of a Bravais lattice can becharacterized by normal vectors, which can be expanded in a basis b1,b2,b3 of primitivevectors of the reciprocal lattice. For a (perpendicular) distance d of the planes the reciprocallattice vector k =

∑3i=1mibi is of length 2π/d. Since the mj have no common divisor, k

is the shortest reciprocal vector perpendicular to the planes.

(i) Show that the density of lattice points per unit area in the lattice planes is d/Vu,where Vu if the volume of the unit cell spanned by a1,a2,a3.

(ii) Show that the reciprocal lattice of the face-centered cubic lattice with lattice constanta is a body-centered cubic lattice with lattice constant 4π/a. By definition thelattice constant a is the edge length of the cube which envelops the unit cell of theface-centered cubic lattice with primitive vectors

a1 =a

2(ey + ez) , a2 =

a

2(ex + ez) , a3 =

a

2(ex + ey) .

ex, ey, ez is the canonical orthonormal basis of R3. The corresponding primitivevectors for the body-centered cubic lattice with lattice constant a′ are

a1 =a′

2(−ex + ey + ez) , a2 =

a′

2(ex − ey + ez) , a3 =

a′

2(ex + ey − ez) .

(iii) Find the Miller indices (m1,m2,m3) of that plane of the face-centered cubic latticewhich has the highest density of lattice points. Here it may be helpful to use theconnection between the density and the reciprocal lattice vector k.

4.8 Exercise 5. Face centered tetragonal structure

Why does the face centered tetragonal structure not appear in the list of the 14 Bravaisclasses? How does this lattice fit into one of the 14 Bravais classes?

L5 The action of the Bravais lattice on states

5.1 Shift operators, lattice momentum and Bloch’s theorem

For a set of primitive vectors a1,a2,a3 ∈ B define the corresponding shift operatorsUaj , acting on a single-particle space of states, by

UajΨ(x) = Ψ(x + aj) . (5.1)

Page 25: Introduction to Solid State Physics - arXiv

25

Crystal system Group primitive base-centered body-centered face-centered

Triclinic Ci

Monoclinic C2h

Orthorhombic D2h

Tetragonal D4h

Rhombohedral D3d

Hexagonal D6h

Cubic Oh

Figure 2: The crystal systems and Bravais classes (from Wikipedia, the free encyclopedia).A parallelepiped representing the point-group symmetry of a Bravais lattice has six pa-rameters, three lengths of its edges and three angles. If all edge lengths and all angles aremutually distinct, the symmetry is minimal (only the inversion). This is the triclinic crystalclass in the table. Considering all possible degeneracies (two right angles, to equal edgelengths etc.) one runs through all the listed symmetry classes, the crystal systems. Some ofthem can be realized by several Bravais lattices, giving the different Bravais classes. Thesecond column in the table contains the name of the point group in so-called Schonfliesnotation.

Page 26: Introduction to Solid State Physics - arXiv

26

ThenU−1aj = U+

aj , [Uaj , Uak ] = 0 , j, k = 1, 2, 3 . (5.2)

For any Bravais lattice vector R = `a1 +ma2 + na3 the operator

UR = U `a1Uma2

Una3(5.3)

is therefore uniquely defined and naturally acts as

URΨ(x) = Ψ(x + R) (5.4)

on single-particle wave functions.Equation (5.2) implies that the Uaj have a joint system of eigenfunctions. If Ψ is such

an eigenfunction, then

UajΨ(x) = Ψ(x + aj) = ω(aj)Ψ(x) . (5.5)

Since Uaj is unitary, |ω(aj)| = 1 implying that ∃ vj ∈ R such that ω(aj) = ei2πvj . LetR = `a1 +ma2 + na3, k = v1b1 + v2b2 + v3b3. Then

Ψ(x + R) = URΨ(x) = ei2π(`v1+mv2+nv3) Ψ(x) = ei〈k,R〉Ψ(x) . (5.6)

Thus, for every common eigenfunction Ψ of the three generators Uaj of lattice transla-tions ∃k ∈ R3 such that (5.6) holds for all R ∈ B. The vector k is a triple of quantumnumbers characterizing the eigenstates of the lattice translation operators in very muchthe same manor as the momentum p is a triple of quantum numbers that characterize theeigenstates of the operator of infinitesimal translations, the momentum operator. For thisanalogy k is called the lattice momentum.

Let g ∈ B, R ∈ B. Then 〈g,R〉 = m2π for some m ∈ Z and ei〈k+g,R〉 = ei〈k,R〉.This means that k and k + g characterize the same eigenstate of the lattice translationoperator, or that the lattice momentum is defined only modulo reciprocal lattice vectors. Forthis reason we may restrict the domain of definition of k to any unit cell of the reciprocallattice. This domain is conventionally taken as the first Brillouin zone, which explains theimportance of the latter.

In the language of group theory the lattice momenta k label the irreducible repre-sentations of the Bravais lattice. Since the Bravais lattice is an Abelian group, all of itsirreducible representations must be one-dimensional. They act by multiplication withcomplex numbers as can be seen in equation (5.6).

The Hamiltonian of the solid (1.1) is invariant under any infinitesimal translation.For this reason the center of mass momentum of the solid is conserved. In nature thetranslation symmetry is affected by a mechanism called spontaneous symmetry breaking.After separating the center of mass motion the ground state of the Hamiltonian (1.1) is lesssymmetric than the Hamiltonian itself. Instead of the full translation symmetry it exhibits adiscrete translation symmetry with an underlying Bravais lattice B. Effective Hamiltoniansdescribing the dynamics above the ground state again have the reduced symmetry describedby a Bravais lattice. The simplest example for a class of such effective Hamiltonians isthe Hamiltonian a single electron in a lattice periodic potential. The study of this class ofHamiltonians is called band theory. We will have a closer look at it below. In any case,single-particle Hamiltonians H which are invariant under the action of a Bravais lattice,

[H,Uaj ] = 0 , j = 1, 2, 3 , (5.7)

Page 27: Introduction to Solid State Physics - arXiv

27

play an important role in solid state physics. As we have seen, their wave functions canbe labeled by the lattice momentum quantum numbers. This is the statement of Bloch’stheorem.

Theorem 1. Bloch [1]. The eigenfunctions of a single-particle Hamiltonian H , periodicwith respect to a Bravais lattice B, can be labeled by lattice momenta k ∈ BZ, whereBZ ⊂ B is the Brillouin zone associated with B. An eigenfunction Ψk of H then has thefollowing properties with respect to translations by Bravais lattice vectors,

Ψk(x + R) = ei〈k,R〉Ψk(x) (5.8)

for all R ∈ B.

Let Ψk,α an eigenstate with lattice momentum k of a lattice periodic Hamiltonian H .Here we denote all other quantum numbers needed to specify the state by α. According toBloch’s theorem uk,α(x) = e−i〈k,x〉Ψk,α(x) is a lattice periodic function, uk,α(x+R) =uk,α(x). This implies the following corollary to Bloch’s theorem.

Corollary 1. The eigenfunctions of a lattice periodic single-particle Hamiltonian H areof the form

Ψk,α(x) = ei〈k,x〉 uk,α(x) , (5.9)

where k ∈ BZ is a lattice momentum vector and uk,α is a lattice periodic function.

Hence, we may think of the eigenfunctions of a lattice periodic Hamiltonian as ofamplitude-modulated plane waves, for which the modulation has the periods of the corre-sponding Bravais lattice.

5.2 Periodic boundary conditions

For the calculation of thermodynamic quantities in the framework of statistical mechanics(in particular) it is necessary to count states. For this reason we prefer systems of finite sizewhich have a discrete spectrum. In solid state physics this can be enforced by introducing‘boundaries’ (by putting the system into a box). After having counted the states oneconsiders the limit, when the systems size goes to infinity (the thermodynamic limit).

In general, boundaries are incompatible with lattice translations induced by a Bravaislattice. They break the translational symmetry and invalidate Bloch’s theorem. A way outof this dilemma is by employing periodic boundary conditions.

For a d-dimensional systems periodic boundary conditions can be realized by startingwith a parallelepiped and identifying opposite faces. This amounts to bending the paral-lelepiped to a torus in d+ 1 dimensions. For this reason periodic boundary conditions arealso sometimes called toroidal boundary conditions.

Imposing periodic boundary conditions on a 1d system of L sites with lattice constanta we find for a state with lattice momentum k that

Ψk(x+ aL) = Ψk(x) = eikaL Ψk(x) . (5.10)

For this to hold, the lattice momentum must be restricted to the values

k = m2π

aL, (5.11)

Page 28: Introduction to Solid State Physics - arXiv

28

where m ∈ Z in such a way that k lies in the Brillouin zone. The reciprocal lattice isgenerated by b = 2π/a. The Brillouin zone is the interval BZ = [−π/a, π/a), andk ∈ BZ ⇔ −L/2 ≤ m < L/2. Thus, for every L ∈ N there are L inequivalent ks inthe Brillouin zone.

The argument is similar in any number of dimensions. In order to obtain the latticemomentum quantization condition, e.g. in 3d, we expand k in a basis of primitive vectors ofthe reciprocal lattice, k = v1b1+v2b2+v3b3. Then, for a state Ψk of lattice momentum k,

Ψk(x + Laj) = ei〈k,Laj〉Ψk(x) = Ψk(x) , (5.12)

requiring that 〈k, Laj〉 = 2πvjL = 2πmj for mj ∈ Z. Restricting k to the first Brillouinzone means

vj =mj

Lmod 1 ⇔ −L

2≤ mj <

L

2, (5.13)

i.e., there are L3 inequivalent lattice momenta in the first Brillouin zone. Let us rephrasethis statement in the following form.

Lemma 4. There are as many lattice momenta in the Brillouin zone that are compatiblewith periodic boundary conditions as unit cells in the crystal.

Remark. In 3d periodic boundary conditions cannot be physically realized. However,the density of states of a macroscopically large system (∼ 1023 particles) is practicallyindependent of the boundary conditions. As long as we are not interested in the boundariesthemselves, periodic boundary conditions are justified.

Remark. Mathematically periodic boundary conditions imply that we are dealing withfunctions that are periodic with periods of the large parallelepiped spanned by La1, La2,La3. Hamiltonians must be defined in a way that is compatible with this periodicity. Onthe corresponding space of states the generators Uaj of translations by primitive vectorsturn into generators of the cyclic group of order L,

ULaj = id , j = 1, 2, 3 , (5.14)

which equivalently might have served as a starting point for introducing periodic boundaryconditions.

L6 Phonons – spectrum and states

6.1 The Hamiltonian of the lattice vibrations in harmonic approximation

In lecture L2 we have discussed the Born-Oppenheimer approximation. We have seen that,in this approximation, the motion of the much heavier ions decouples from the motion ofthe electrons up to fourth order in an expansion in the deviations u of the ion positionsfrom their equilibrium values R(0). In most of this lecture we shall be dealing with idealsolids. By definition, the equilibrium positions of the ions R(0) in an ideal solid are thepoints of a crystal lattice. Solids in nature can be very close to ideal. The idealization of aperfect crystal is a good starting point to describe real solids.

Consider a crystal with Bravais lattice B and N ions per unit cell. In such a crystal itmakes sense to label the coordinates of the vector u as uα(R), where R ∈ B and

α = (r, j) ∈ I = 1, . . . , N × x, y, z . (6.1)

Page 29: Introduction to Solid State Physics - arXiv

29

Here r = 1, . . . , N counts the ions in a unit cell, and j = x, y, z denotes their Cartesiancoordinates. For the dimensionless ion masses (cf. Section 2.3) we introduce the notation

µα = µ(r,j) =Mr

M(6.2)

Then the operator Tu, equation (2.11a), of the kinetic energy of the ions takes the form

Tu = −1

2

∑R∈B

∑α∈I

1

µα∂2uα(R) . (6.3)

We shall treat the motion of the ions within the harmonic approximation (2.25). Thiswill allow us to develop a rather simple and general theory which nevertheless describesmany of the experimental observations quite accurately. Within the harmonic approxima-tion the potential energy of the ions can be written as

V (u) =1

2

∑R,S∈B

∑α,β∈I

uα(R)Kαβ(R,S)uβ(S) , (6.4)

where

Kαβ(R,S) =∂2ε0(X)

∂Xα(R)∂Xβ(S)

∣∣∣∣X=R(0)

(6.5)

is the so-called force matrix (see (2.26)).Thus, the Hamiltonian of the lattice vibrations in harmonic approximation is

H = −1

2

∑R∈B

∑α∈I

1

µα∂2uα(R) +

1

2

∑R,S∈B

∑α,β∈I

uα(R)Kαβ(R,S)uβ(S) . (6.6)

It is a quadratic form in the position operators of the ions and the corresponding derivatives.We will diagonalize this quadratic form. This will reduce the spectral problem of theHamiltonian to the spectral problem of independent harmonic oscillators describing thequantized normal modes of the ideal harmonic solid. In order to control the number of thenormal modes, we will employ periodic boundary conditions as introduced in the previouslecture.

6.2 Implications of the translational invariance

Fix a set of primitive vectors a1,a2,a3 ⊂ B. Define the action of B on functions v onB obeying periodic boundary conditions by

Uajv(R) = v(R + aj) . (6.7)

The space spanned by such functions is a finite dimensional vector space. For the actionof Uaj on this space there is an L ∈ N such that ULaj = id, j = 1, 2, 3. Thus, if ωj is aneigenvalue of Uaj we must have |ωj | = 1 (because of unitarity) and ωLj = 1. It followsthat

ωj = eimj2π

L = ei〈k,aj〉 (6.8)

for some k = (m1b1 +m2b2 +m3b3)/L ∈ BZ. As we have seen in the previous lecturethere are altogether L3 such vectors.

Page 30: Introduction to Solid State Physics - arXiv

30

It is easy to find the corresponding eigenfunctions of the shift operators. If vk is aneigenfunctions with lattice momentum k, and R = `a1 +ma2 + na3 ∈ B, then

vk(R) = vk(`a1 +ma2 + na3) = U `a1Uma2

Una3vk(0)

= ω`1ωm2 ω

n3 vk(0) = ei〈k,R〉 vk(0) (6.9)

for all R ∈ B. This determines vk up to normalization. We fix the normalization by setting

vk(0) =1√L3

. (6.10)

All joint eigenfunctions of the Uaj are of this form, and all these functions are jointeigenfunctions of the Uaj . Hence, they form a basis of on the space of complex valuedfunctions on B.

Remark. (i) We have just invented the (discrete) Fourier transformation.

(ii) With the choice (6.10) the usual Hermitian scalar product of two eigenfunctionstakes the values

〈vk, vq〉 =∑R∈B

v∗k(R)vq(R) =1

L3

∑R∈B

e−i〈k−q,R〉 = δk,q (6.11)

for any two k,q ∈ BZ.

6.3 Block diagonalization of the force matrix

We will first of all diagonalize the force matrix Kαβ(R,S) defined in (6.5). For thispurpose and for later use as well we list its main properties.

(i) Symmetry. From its very definition as a second derivative matrix and from thecommutativity or the partial derivatives we get at once that

Kαβ(R,S) = Kβα(S,R) . (6.12)

(ii) Translation symmetry.

Kαβ(R + a,S + a) = Kαβ(R,S) ∀ a ∈ B . (6.13)

Inserting here a = −S we obtain

Kαβ(R,S) = Kαβ(R− S, 0) = Kαβ(R− S) , (6.14)

where the second equation is a definition. The meaning is that forces between ionsdepend only on the relative positions of unit cells.

(iii) In crystal lattices with inversion symmetry we have in addition that

Kαβ(R) = Kαβ(−R) . (6.15)

Combining this with (6.12) we see that the force matrix in crystal lattices withinversion symmetry is symmetric in every unit cell,

Kαβ(R) = Kβα(R) . (6.16)

Page 31: Introduction to Solid State Physics - arXiv

31

The force matrix defines an operator on the space of functions with periodic boundaryconditions on B,

Kαβf(R) =∑S∈B

Kαβ(R− S)f(S) . (6.17)

It is easy to see that Kαβ commutes with the shift operators Uaj , j = 1, 2, 3,

UajKαβf(R) =∑S∈B

Kαβ(R + aj − S)f(S)

=∑S∈B

Kαβ(R− S)f(S + aj) = KαβUajf(R) , ⇔ [Kαβ, Uaj ] = 0 . (6.18)

Hence, Uaj and Kαβ possess a joint system of eigenfunctions. Since the vk, equation (6.9),form already an orthonormal basis of non-degenerate eigenfunctions, they must be alsoeigenfunctions of Kαβ ,

Kαβvk(R) = καβ(k)vk(R) . (6.19)

In order to block diagonalize the quadratic form (6.4) we expand the displacementsuα(R) into their Fourier modes,

uα(R) =∑

k∈BZξαkvk(R) . (6.20)

Using that uα(R) ∈ R and v∗k = v−k we see that

ξαk∗ = ξα−k . (6.21)

Inserting (6.20) into (6.4) and making use of (6.19), (6.21) we obtain

V (u) =1

2

∑α,β∈I

⟨uα, Kαβu

β⟩

=1

2

∑α,β∈I

∑q,k∈BZ

ξαq∗ξβk⟨vq, Kαβvk

⟩=

1

2

∑q∈BZ

∑α,β∈I

ξα−qκαβ(q)ξβq , (6.22)

the block diagonal form of the potential energy.

6.4 Transformation of the kinetic energy

Let us now apply the same transformation to the kinetic energy operator. First of all

〈uα, v−k〉 =∑

q∈BZξα−q〈vq, v−k〉 = ξαk =

∑S∈B

uα(S)v−k(S) , (6.23)

implying that

∂uα(R)=∑

k∈BZ

∂ξαk∂uα(R)

∂ξαk=∑

k∈BZv−k(R)

∂ξαk. (6.24)

It follows that

Page 32: Introduction to Solid State Physics - arXiv

32

Tu = −1

2

∑α∈I

1

µα

∑R∈B

∑q,k∈BZ

v−q(R)v−k(R)∂

∂ξαq

∂ξαk

= −1

2

∑α∈I

1

µα

∑q,k∈BZ

〈vq, v−k〉∂

∂ξαq

∂ξαk= −1

2

∑q∈BZ

∑α∈I

1

µα

∂ξα−q

∂ξαq. (6.25)

Here we have used (6.11) in the second equation. We see that the lattice Fourier transfor-mation has diagonalized Tu.

In the next step we want to completely diagonalize the force matrix while keeping thediagonal form of the kinetic energy operator. To achieve the latter goal, we first rescale thecomplex coordinates, setting

ηαq =√µα ξ

αq ⇒

∂ηαq=

1√µα

∂ξαq. (6.26)

Further defining

καβ(q) =καβ(q)√µαµβ

(6.27)

we obtain the following form of the Hamiltonian (6.6),

H =1

2

∑q∈BZ

−∑α∈I

∂ηα−q

∂ηαq+∑α,β∈I

ηα−qκαβ(q)ηβq

. (6.28)

6.5 Properties of the matrix κ(q)

Before we can proceed we have to understand the properties of the matrix κ(q).

(i) κ(q) is Hermitian, since

κ∗αβ(q) =⟨vq, Kαβvq

⟩∗=

∑R,S∈B

vq(R)Kαβ(R− S)v−q(S)

=∑

R,S∈Bvq(R)Kβα(S−R)v−q(S)

=∑

R,S∈Bv−q(R)Kβα(R− S)vq(S) =

⟨vq, Kβαvq

⟩= κβα(q) . (6.29)

Here we have used the symmetry of the force matrix in the third equation.

(ii) κ(q) is non-negative. This follows, since the potential energy V is assumed to havea total minimum for u = 0 with V = 0. Hence

V (u) =1

2

∑α,β∈I

ηα−qκαβ(q)ηβq ≥ 0 . (6.30)

(iii) κ(q) and κ(−q) are similar matrices. First of all

κ∗αβ(q) =∑

R,S∈Bvq(R)Kαβ(R− S)v−q(S) = καβ(−q) , (6.31)

Page 33: Introduction to Solid State Physics - arXiv

33

since the force matrix is real. Then also

κ∗αβ(q) = καβ(−q) . (6.32)

According to (i) and (ii) the matrix κ(q) can be diagonalized by a unitary trans-formation and has a non-negative spectrum ω2

α(q)α∈I for every q ∈ BZ. Letyα(q)α∈I the set of corresponding orthonormal eigenvectors. Then

κ∗(q)yα(q)∗ = ω2α(q)yα(q)∗ = κ(−q)yα(q)∗ (6.33)

implying that ω2α(q)α∈I is the spectrum of κ(−q). Hence, κ(q) and κ(−q) are

similar matrices.

Since, on the other hand,

κ(−q)yα(−q) = ω2α(−q)yα(−q) (6.34)

by definition of the eigenvectors and eigenvalues, the identification

yα(−q) = yα(q)∗ (6.35)

(which is one possible choice of indexing the eigenvectors of κ(−q) once theeigenvectors of κ(q) are given) implies that

ω2α(q) = ω2

α(−q) . (6.36)

6.6 Exercise 6. Classical harmonic chain with various boundary conditions

The lectures on lattice vibrations will be accompanied by a set of exercises on the classicalharmonic chain with broken translation invariance. We shall study the influence of fixedand open boundary conditions and of a mass defect. There is a good deal to learn fromthese exercises, namely something about the irrelevance of the boundary conditions asfar as bulk thermodynamic properties are concerned, but also something about the typicaleffects of impurities, such as the appearance of localized states and impurity levels insidethe band gap.

We start with an important technical device, the transfer matrix, and with the effectof fixed and open boundary conditions. For this purpose consider N masses m1, . . . ,mN

coupled to a linear, harmonic chain by N − 1 springs with force constants k > 0. Periodicboundary conditions are realized by an additional identical spring connecting the massesm1 and mN . For fixed boundary conditions, the masses m1 and mN are coupled withsprings of spring constants k to a rigid wall, while open boundaries are realized if themasses m1 and mN are not at all coupled to each other.

In this exercise we shall consider the particular case of equal masses m1 = . . . =mN = m. We want to analyze the harmonic chain by the so-called transfer matrix method.Upon slight modifications, it is possible to treat the different boundary conditions in asimilar way.

(i) Find the equations of motion for the deviations xn from the equilibrium positions inthe periodic case. Employing the ansatz xn(t) = eiωt xn the equations of motionimply an eigenvalue problem of the form Tx = Ω2x with x = (x1, . . . , xN )t.Obtain the matrix T , and show how Ω depends on ω,m and k.

Page 34: Introduction to Solid State Physics - arXiv

34

(ii) Show, that, with the substitution ψ(n) = xn, ϕ(n) = ψ(n − 1), the eigenvalueproblem in (i) can be reformulated as(

ψ(n+ 1)ϕ(n+ 1)

)= Ln(Ω2)

(ψ(n)ϕ(n)

), Ln(Ω2) = L(Ω2) =

(2− Ω2 −1

1 0

).

In the periodic case Ln(Ω2) is independent of the site index. Calculate the eigen-values and the corresponding eigenvectors of L(Ω2). Because L(Ω2) acts like atranslation operator, it is useful to write the eigenvalues in the form e±iκ. Diagonalizethe equation (

ψ(n+ 1)ϕ(n+ 1)

)= Ln(Ω2)

(ψ(1)ϕ(1)

).

How can Ω2 be expressed in terms of κ?

(iii) The periodic boundary conditions turn into ψ(N+1) = ψ(1) and ϕ(N+1) = ϕ(1).From this determine all possible eigenfrequencies ω!

(iv) Which modification is required for fixed boundaries? Determine all possible eigen-frequencies ω in this case.

(v) Show that the modifications necessary for open boundaries lead to the equation

(1− Ω2,−1)LN−2(Ω2)

(1− Ω2

1

)= 0 .

With this, calculate again all possible eigenfrequencies ω. What is the physicalmeaning of the solution ω = 0? How are the eigenfrequencies of the open chain andof the chain with fixed boundaries connected with each other?

L7 Phonons – spectrum and states continued

7.1 Reduction of the Hamiltonian to a diagonal quadratic form

For every q ∈ BZ we define a unitary 3N × 3N matrix

Y (q) = (y(1,x)(q), . . . ,y(N,z)(q)) . (7.1)

This matrix diagonalizes κ(q),

κ(q) = Y (q) diag(ω2

(1,x), . . . , ω2(N,z)

)Y +(q) (7.2)

and, because of (6.35), has the property that

Y (q)∗ = Y (−q) . (7.3)

Lettingxq = Y +(q)ηq ⇔ xαq =

⟨yα(q),ηq

⟩(7.4)

we see that ∑α,β∈I

ηα−qκαβ(q)ηβq = ηt−qκ(q)ηq =∑α∈I

ω2α(q)xα−qx

αq , (7.5)

Page 35: Introduction to Solid State Physics - arXiv

35

while for the kinetic term

∑α∈I

∂ηα−q

∂ηαq=

∑α,β,γ∈I

∂xβ−q∂ηα−q

∂xγq∂ηαq

∂xβ−q

∂xγq

=∑

α,β,γ∈IY +β

α(−q)Y +γα(q)

∂xβ−q

∂xγq=∑α∈I

∂xα−q

∂xαq. (7.6)

Here we have used (7.3) in the last equation. Altogether, we have transformed H into adiagonal quadratic form,

H =1

2

∑q∈BZ

∑α∈I

− ∂

∂xα−q

∂xαq+ ω2

α(q)xα−qxαq

. (7.7)

The term in the bracket can be interpreted as the Hamiltonian of a 1d harmonic oscillatorwith ‘complex coordinates’.

7.2 Zero modes

Before going on we have to discuss the question whether the functions ωα(q) can be zero.What we can say is that there are always at least three values of α for which ωα(0) = 0.These special ‘modes’ are connected with the center of mass motion of the solid. Theirexistence can be inferred from the translation invariance of the Hamiltonian (6.6) whichis inherited from the full Hamiltonian (1.5) of the solid. For (6.6) translation invariancemeans invariance under the transformation

u(r,j)(R) 7→ u(r,j)(R) + εj , (7.8)

for all εj ∈ R and every (r, j) ∈ I . This transformation leaves the kinetic energy (6.3)trivially invariant. For the potential energy (6.4) we infer that

∂εjV∣∣εj=0

=

1

2

∑R,S∈B

N∑r,s=1

∑`=x,y,z

(K(r,j)(s,`)(R− S)u(s,`)(S) + u(r,`)(R)K(r,`)(s,j)(R− S)

)=

∑R,S∈B

N∑r,s=1

∑`=x,y,z

K(r,j)(s,`)(R− S)u(s,`)(S) = 0 (7.9)

for arbitrary u(s,`)(S) ∈ R. Here we have used the symmetry (6.12) of the force matrix inthe second equation. Setting all but one of the displacements equal to zero and this oneequal to one we obtain the relation

∑R∈B

N∑r=1

K(r,j)(s,`)(R) = 0 (7.10)

for the force matrix, which holds for all (s, `) ∈ I and j = x, y, z. On the other hand

καβ(0) =⟨v0, Kαβv0

⟩=

1

L3

∑R,S∈B

Kαβ(R− S) =∑R∈B

Kαβ(R) . (7.11)

Page 36: Introduction to Solid State Physics - arXiv

36

Setting α = (r, j), β = (s, `) summing over r and using (7.10) we conclude that

N∑r=1

√µrµs κ(r,j)(s,`)(0) = 0 (7.12)

for all (s, `) ∈ I and j = x, y, z. Thus, there are three independent linear relations betweenthe rows of the matrix κ(0) which therefore has at least a threefold eigenvalue zero. Wemay order the spectrum of καβ(0) in such a way that the corresponding eigenvectors arey(1,j)(0), j = x, y, z. The corresponding ‘normal coordinates’ are x(1,j)

0 = 〈y(1,j)(0),η0〉.Using this notation, the Hamiltonian (7.7) splits into

H = Tcm +1

2

∑q∈BZ,α∈I

(q,α) 6=(0,(1,j))

− ∂

∂xα−q

∂xαq+ ω2

α(q)xα−qxαq

, (7.13)

where

Tcm = −1

2

∑j=x,y,z

∂2

∂(x

(1,j)0

)2 (7.14)

can be interpreted a the kinetic energy of the center of mass motion of the crystal.The center of mass motion of the crystal is unbounded. If there were any other ‘zero

modes’, i.e., a higher than threefold degeneracy of the eigenvalue zero, then there would beanother eigenvector yα(q) corresponding to unbounded motion. This would necessarilyinvolve an unbounded relative motion of different parts of the crystal, meaning that thecrystal would disintegrate. In the following we shall exclude this possibility and concentrateon stable crystals. We shall also discard the center of mass motion. Then we remain withthe Hamiltonian of the proper lattice vibrations which we denote

Hph = H − Tcm =1

2

∑(q,α)∈Q

− ∂

∂xα−q

∂xαq+ ω2

α(q)xα−qxαq

. (7.15)

Here we have introduced the notation

Q =

(q, α) = BZ × I∣∣ωα(q) 6= 0

. (7.16)

The subindex ‘ph’ refers to ‘phonon’ which is the name of a quantized normal mode of thelattice.

7.3 Diagonalization of the phonon Hamiltonian

To accomplish a complete diagonalization of the Hamiltonian Hph we introduce theoperators

aαq =

√ωα(q)

2xαq +

1√2ωα(q)

∂xα−q, (7.17a)

a+αq =

√ωα(q)

2xα−q −

1√2ωα(q)

∂xαq(7.17b)

Page 37: Introduction to Solid State Physics - arXiv

37

for all (q, α) ∈ Q. They satisfy the commutation relations (exercise: check it!)[aαq, a

βk

]= 0 =

[a+α

q, a+βk

],[aαq, a

+βk

]= δq,kδ

αβ . (7.18)

Inverting (7.17) we obtain

xαq =aαq + a+α

−q√2ωα(q)

,∂

∂xαq=

√ωα(q)

2

(aα−q − a+α

q

). (7.19)

The latter equation implies that

ω2α(q)xα−qx

αq −

∂xα−q

∂xαq

=ωα(q)

2

(aα−q + a+α

q

)(aαq + a+α

−q)− ωα(q)

2

(aαq − a+α

−q)(aα−q − a+α

q

)= ωα(q)

(a+α−qa

α−q + a+α

qaαq + 1

), (7.20)

whenever ωα(q) 6= 0. Here we have used the evenness of the functions ωα, equation (6.36),and the commutation relations (7.18). Inserting (7.20) into (7.15) and using once more thatωα is an even function of q we arrive at

Hph =∑

(q,α)∈Q

ωα(q)(a+α

q aαq + 1

2

). (7.21)

Thus, Hph is decomposed into a sum of independent harmonic oscillators.

7.4 Creation and annihilation operators in terms of the original displace-ments

Going step by step backwards, we express the creation and annihilation operators of thephonons in term of the original displacement variables and their associated momentumoperators

pα(R) = −i∂

∂uα(R). (7.22)

We obtain

xαq =∑β∈I

Y +αβ(q)ηβq =

∑β∈I

Y +αβ(q)√µβ ξ

βq

=∑β∈I

Y +αβ(q)√µβ 〈uβ, v−q〉 =

1√L3

∑R∈B

∑β∈I

e−i〈q,R〉√µβ Y +αβ(q)uβ(R)

(7.23)

and

∂xα−q=∑β∈I

∂ηβ−q∂xα−q

∂ηβ−q=∑β∈I

Y βα (−q)√µβ

∂ξβ−q=∑β∈I

Y +αβ(q)√µβ

∂ξβ−q

=1√L3

∑R∈B

∑β∈I

e−i〈q,R〉 Y+αβ(q)√µβ

ipβ(R) . (7.24)

Page 38: Introduction to Solid State Physics - arXiv

38

Inserting the latter two equations into the definitions (7.17) of the annihilation and creationoperators we obtain

aαq =1√L3

∑R∈B

∑β∈I

e−i〈q,R〉 Y +αβ(q)

[√µβωα(q)

2uβ(R) +

ipβ(R)√2µβωα(q)

], (7.25a)

a+αq =

1√L3

∑R∈B

∑β∈I

ei〈q,R〉 Y βα (q)

[√µβωα(q)

2uβ(R)− ipβ(R)√

2µβωα(q)

]. (7.25b)

In this form it is obvious that aαq and a+αq are mutually adjoint operators.

7.5 Construction of the eigenstates

Let

Ψ0(u) = exp

−1

2

∑(k,β)∈Q

ωβ(k)xβ−kxβk

. (7.26)

Then

∂Ψ0(u)

∂xα−q=

(−ωα(q)xαq

2−ωα(−q)xαq

2

)Ψ0(u) = −ωα(q)xαqΨ0(u)

⇔ aαqΨ0(u) = 0 (7.27)

for all (q, α) ∈ Q. Hence,

HphΨ0(u) = E0Ψ0(u) , E0 =1

2

∑(q,α)∈Q

ωα(q) . (7.28)

Ψ0 is the ground state, since it is the ground state for every 1d harmonic oscillator in thesum (7.21).

Comparing (7.26) and (7.15) and recalling the original definition (6.4) of the harmonicpotential, we obtain the ground state wave function as a function of the displacements u ofthe ions,

Ψ0(u) = e−V (u) = exp

−1

2

∑R,S∈B

∑α,β∈I

uα(R)Kαβ(R,S)uβ(S)

, (7.29)

which is a natural generalization of the 1d case.A general phonon state is generated by the multiple action of phonon creation operators

a+αq on the ground state which, for this reason, is also sometimes called the phonon

vacuum (exercise: repeat the construction of excited states for the 1d harmonic oscillatorbased on the Heisenberg algebra (7.18)). Such states are parameterized by maps Q→ N0,(q, α) 7→ nαq. Accordingly we shall denote them as

Ψn(u) =∏

(q,α)∈Q

(a+α

q

)nαq Ψ0(u) . (7.30)

It follows from the commutation relations (7.18) that

HphΨn(u) = EnΨn(u) , En =∑

(q,α)∈Q

ωa(q)(nαq + 1

2

). (7.31)

Page 39: Introduction to Solid State Physics - arXiv

39

The eigenvectors yα(q) of κ(q), together with the corresponding eigenfrequenciesωα(q), determine the creation operators a+α

q, equation (7.25b), since Y βα (q) = (yα)β . In

analogy with the expression for the quantized electro-magnetic field we shall call thempolarization vectors.

Let us summarize the insight we have gained so far in the following

Theorem 2. In order to obtain the spectrum (7.31) and the eigenstates (7.29), (7.30) of thevibrational motion of the ions in a solid in harmonic approximation, it suffices to calculatethe dispersion relations ωα(q) and the polarization vectors yα(q). For this purpose onefirst calculates the matrix

καβ(q) =∑R∈B

e−i〈q,R〉Kαβ(R)√µαµβ

(7.32)

and then its eigenvectors yα(q) and eigenvalues ω2α(q).

The input here is the force matrix. In applications it comes from quantum chemicalcalculations or from simple heuristic models. Note that in (7.32) every matrix elementκαβ(q) is represented as a (finite) Fourier series (cf. section 3.5) defining it as a periodicfunction in reciprocal space with periods in B. We expect the forces between ions to decayrapidly with distance and the convergence of the Fourier series (7.32) in the thermodynamiclimit to be uniform, implying that the limit function is differentiable in q.

L8 Phonons – examples and general properties

8.1 Example 1 – the harmonic chain

To start with we reconsider the harmonic chain within the framework of the general theory.This is a 1d problem with one ion per unit cell, thus no indices α, β are required and µ = 1.Denoting the lattice spacing by a we obtain

q =n

L· 2π

a, R = `a (8.1)

for the quantized lattice momenta q and the Bravais lattice vectors R.The model force matrix is

K(R,S) = K(`a,ma) = K((`−m)a

)= ω2

0(2δ`,m − δ`,m+1 − δ`,m−1) , (8.2)

where we have to keep the periodic boundary conditions in mind. This is a 1× 1 matrix.The polarization vector is y = 1 and

κ(q) =∑`

e−in2πL` ω2

0 (2δ`,0 − δ`,1 − δ`,−1) = ω202(1− cos(2πn/L)

)= 2ω2

0

(1− cos(qa)

)= 4ω2

0 sin2(qa/2) = ω2(q) . (8.3)

It follows thatω(q) = 2ω0| sin(qa/2)| (8.4)

which is called the dispersion relation of the harmonic chain with nearest-neighbourinteractions.

Page 40: Introduction to Solid State Physics - arXiv

40

α2 α2 α1

u1(R)

m2 m1 m2 m1α1

u2(R− a) u1(R + a)u2(R)

Figure 3: Sketch of a diatomic chain with alternating masses m1 and m2 and alternatingforce constants α1 and α2.

8.2 Example 2 – diatomic chain with alternating forces

The classical model for this configuration is given by the equations of motion

µ1u1(R) = α1

(u2(R)− u1(R)

)− α2

(u1(R)− u2(R− a)

),

µ2u2(R) = α2

(u1(R+ a)− u2(R)

)− α1

(u2(R)− u1(R)

), (8.5)

whereµj =

2mj

m1 +m2, j = 1, 2 , (8.6)

are the dimensionless masses and α1, α2 dimensionless force constants (see Figure 3). Thecorresponding force matrix is (F = − gradV )

K(R,S) =

((α1 + α2)δR,S −α1δR,S − α2δR,S+a

−α1δR,S − α2δR,S−a (α1 + α2)δR,S

). (8.7)

It follows that

κ(q) =

(α1 + α2 −α1 − α2 e−iqa

−α1 − α2 eiqa α1 + α2

)(8.8)

and

κ(q) =

α1+α2µ1

−α1−α2 e−iqa√µ1µ2

−α1−α2 eiqa√µ1µ2

α1+α2µ2

. (8.9)

As it should be, this is a Hermitian 2× 2 matrix. We have to calculate its eigenvalues andeigenvectors.

For the eigenvalues λ± of a 2 × 2 matrix A we have the general formula (exercise:check it!)

λ± =trA

2±√(

trA

2

)2

− detA . (8.10)

For the matrix κ(q) we calculate

tr κ(q) =2(α1 + α2)

µ1µ2, (8.11a)

det κ(q) =α1α2

µ1µ24 sin2

(qa2

). (8.11b)

It follows that

Page 41: Introduction to Solid State Physics - arXiv

41

-3 -2 -1 1 2 3

0.5

1.0

1.5

2.0

ω

ω+

ω−

q

Figure 4: Dispersion relation (8.12) of the diatomic chain. Upper branch ω+(q), lowerbranch ω−(q). Parameters m1 = 1,5, m2 = 2,0, α1 = 1,3, α2 = 0,9, a = 1. The lowerbranch is called the acoustic branch, the upper branch is called the optical branch. Thebranches are separated by the ‘band gap’ which is the distance between the dashed lines.

ω2±(q) =

α1 + α2

µ1µ2±√(

α1 + α2

µ1µ2

)2

− α1α2

µ1µ24 sin2

(qa2

)

=(α1 + α2)(m1 +m2)2

4m1m2

√1− 4α1α2

(α2 + α2)2

4m1m2

(m1 +m2)2sin2

(qa2

). (8.12)

Recalling that 4xy ≤ (x+ y)2 ⇔ 0 ≤ (x− y)2 for all x, y ∈ R we may conclude that

0 <4α1α2

(α2 + α2)2

4m1m2

(m1 +m2)2≤ 1 (8.13)

as it must be for ω2±(q) to be real.

The two branches ω± of the dispersion relation are sketched in Figure 4. Note that thelower branch is going to zero linearly as q goes to zero,

ω−(q) = vsq + O(q2) , vs =

√α1α2

2(α1 + α2)a . (8.14)

This is the limit of long wave lengths. For this reason the branch is called the acousticbranch, vs is the sound velocity.

The branch ω+ is called the optical branch. The normal modes at small q in the acousticbranch correspond to motions when the two atoms in the unit cell move in phase, whilesmall values of q in the optical branch correspond to motions when the two atoms moveagainst each other. This can be seen by looking at the polarization vectors, which we leaveas an exercise. As we see in Figure 4 the frequencies in the optical branch are higher thatin the acoustic branch. In real solids typical optical branches correspond to frequencies inthe infrared.

Page 42: Introduction to Solid State Physics - arXiv

42

The numbersW± = maxω±(q)−minω±(q) (8.15)

are called the band width of the optical and acoustic branches. They quantify the rangesof available frequencies or bands. In real solids, like in our Figure 4, optical bands havetypically smaller band widths than acoustic bands. The number

g = minω+(q)−maxω−(q) (8.16)

is called the band gap. It corresponds to a range of ‘forbidden frequencies’. As we shallsee the existence of bands and band gaps explain many of the characteristic feature ofsolids observed in experiments.

It is interesting to see, how the monoatomic chain of example 1 is recovered forα1 = α2 = α and m1 = m2. For this special choice of parameters

ω2±(q) = 2α

(1± cos

( qa2

)), (8.17)

since |q| ≤ π/a. This can be rewritten as

ω2−(q) = 4α sin2

( qa4

), (8.18a)

ω2+(q) = 4α cos2

( qa4

)= 4α sin2

((q ± 2π

a )a4). (8.18b)

We see that ω+ is the same function as ω−, shifted by ±2πa . The band gaps have vanished,

and by shifting ω+ on the interval [−π/a,0] by 2π/a to the right and the same functionon the interval [0, π/a] by 2π/a to the left, we obtain the function ω− on the doubledBrillouin zone [−2π/a,2π/a]. Thus, we have two equivalent descriptions, two branchesof the dispersion relation on the original Brillouin zone [−π/a, π/a] or one branch on[−2π/a,2π/a]. The doubling of the Brillouin zone corresponds to a bisection of the unitcell in the Bravais lattice (exercise: draw the pictures!).

8.3 General properties of phonons in a 3d lattice

(i) There are 3N branches ωα(q) of the dispersion relation for a crystal lattice with Natoms per unit cell.

(ii) In the thermodynamic limit the lattice momenta q densely fill the Brillouin zone, andthe matrix κ(q) as defined in equation (7.32) becomes a continuously differentiablefunction of q. Then, due to the implicit function theorem, the functions ω2

α(q)become differentiable functions of q. As also follows from (7.32) they are naturallyextended as periodic functions on the reciprocal lattice B.

(iii) There are precisely three acoustic branches for which ωα(0) = 0. All other branchesare optical branches with minωα > 0. Since any continuous functions assumes itsextremum on compact sets, every phonon band has finite band width.

(iv) Of the three acoustic branches one has longitudinal, the others have transversalpolarization. In general the longitudinal acoustic modes are faster than the transversalacoustic modes (reversing force is larger for pressure waves than for shear waves,becomes clear when thinking about transition to fluid).

Page 43: Introduction to Solid State Physics - arXiv

43

(v) The dispersion relations of the phonons are invariant under the action of the pointgroup R0 of the crystal,

ωα(q) = ωα(Gq) (8.19)

for all (q, α) ∈ Q, for all G ∈ R0.

Proof. Let G ∈ R0 ⊂ O(3). Then G acts naturally, as a rotation or a rotation followed byan inversion, on the Bravais lattice and on the displacements u(r,j)(R), j = x, y, z, of theindividual ions from their equilibrium positions. The latter action combines into the actionof a representation D of R0 on the vectors of displacement u,

u′α(GR) =

∑β∈I

Dαβ (G)uβ(R) . (8.20)

This transformation leaves the kinetic energy and the potential energy of the Hamiltonian(7.15) of the harmonic crystal separately invariant, as all ions simultaneously undergo thesame O(3) transformation, which does not affect their relative displacements. The crystallattice is not necessarily invariant under this transformation, but the effect on the crystallattice is at most a translation, since any point group operation can be seen as a combinationof a space group operation and a translation. This is another way of understanding thepoint group invariance of the Hamiltonian (7.15).

Let us work out the consequences of the invariance for the representation D. First ofall, setting R′ = GR,

∂uα(R)=∑β∈I

∂u′β(R′)

∂uα(R)

∂u′β(R′)=∑β∈I

Dβα(G)

∂u′β(R′), (8.21)

and hence

Tu = −1

2

∑R∈B

∑α∈I

1

µα

∂2

∂uα(R)2

= −1

2

∑R∈B

∑α∈I

1

µα

∑β,γ∈I

Dβα(G)Dγ

α(G)∂

∂u′β(R′)

∂u′γ(R′)

= −1

2

∑R∈B

∑β,γ∈I

[∑α∈I

Dβα(G)Dγ

α(G)

µα

]∂

∂u′β(R)

∂u′γ(R), (8.22)

where we have used the invariance of B under R0 in the third equation. Form invarianceof the kinetic energy now means that

∑α∈I

Dβα(G)Dγ

α(G)

µα=δβγ

µβ. (8.23)

Similarly, the potential energy transforms like

V (u) =1

2

∑R,S∈B

∑α,β∈I

uα(R)Kαβ(R− S)uβ(S)

Page 44: Introduction to Solid State Physics - arXiv

44

=1

2

∑R,S∈B

∑α,β∈I

uα(GR)Kαβ

(G(R− S)

)uβ(GS)

=1

2

∑R,S∈B

∑α,β∈I

u′α(R)

[∑γ,δ∈I

Dγα(G)Kγδ

(G(R− S)

)Dδβ(G)

]u′β(S) . (8.24)

Then form invariance of this expression implies that

Kαβ(R− S) =∑γ,δ∈I

Dγα(G)Kγδ

(G(R− S)

)Dδβ(G) . (8.25)

It follows that∑γ,δ∈I

Dγα(G)κγδ(Gq)Dδ

β(G) =∑R∈B

e−i〈Gq,R〉∑γ,δ∈I

Dγα(G)Kγδ(R)Dδ

β(G)

=∑R∈B

e−i〈Gq,GR〉∑γ,δ∈I

Dγα(G)Kγδ(GR)Dδ

β(G) = καβ(q) . (8.26)

Here we have used (8.25) and the fact that 〈Gq, GR〉 = 〈q,R〉 in the last equation.Equation (8.26) is equivalent to saying that

καβ(q) =∑γ,δ∈I

Dγα(G)κγδ(Gq)Dδ

β(G) , (8.27)

where

Dαβ(G) =

√µαµβDαβ (G) (8.28)

and therefore∑β∈I

Dαβ(G)Dtβ

γ (G) =∑β∈I

Dαβ(G)Dγ

β(G) =√µαµγ

∑β∈I

Dαβ (G)Dγ

β(G)

µβ= δαγ . (8.29)

This means that κ(q) and κ(Gq) are similar matrices which implies our claim.

8.4 Exercise 7. Linear chain with a mass defect

For equal masses the solution of the periodic chain in Exercise 6 leads to the acousticphonons of the simple one-dimensional lattice. If the masses are different, however, thereis, in general, no simple closed solution, not even of the one-dimensional problem.

In the following we shall study the influence of a mass defect. We assume that allmasses but one are equal to m and that the remaining mass is equal to m(1 + µ). As inExercise 6 the ansatz of an harmonic time dependence leads to an eigenvalue problem ofthe form(

ψ(1)

ϕ(1)

)=

(ψ(N + 1)

ϕ(N + 1)

)= LN · · ·L1

(ψ(1)

ϕ(1)

), Ln =

(2− Ω2

n −1

1 0

). (8.30)

Here ψ(n) and ϕ(n) are defined as in Exercise 6, and Ω2n = mnω

2/k withm1 = m(1+µ)and mj = m for j = 2, . . . , N .

Page 45: Introduction to Solid State Physics - arXiv

45

(i) Solve the eigenvalue problem (8.30) and show that κ in Ω2 = 4 sin2(κ/2) satisfiesthe transcendental equation

tan

(Nκ

2

)+ µ tan

(κ2

)= 0 (8.31)

or eiκ = ±1 or eiNκ = 1.

(ii) In order to solve (8.31) graphically we transform it into a more convenient form. Forthis purpose set z = eiκ. First prove that

1

zN − 1=

1∏Nj=1(z − zj)

=1

N

N∑j=1

zjz − zj

(8.32)

and obtain an analogous relation for z−1j . Here the zj , j = 1, . . . , N , are the N th

roots of unity. Then show that (8.31) is equivalent to

N

µ+

N∑j=1

Ω2

Ω2 − Ω2j

= 0 , (8.33)

where the Ωj are the eigenfrequencies of the problem with equal masses. Nowdiscuss (8.33) graphically. Which equations for the frequencies do you obtain forthe particular cases when µ = −1 or µ =∞?

(iii) How many solution do you obtain from (8.33)? Where are the missing solutions andwhat is their interpretation?

8.5 Exercise 8. Mass defect in the thermodynamic limit

In exercise 7 the consequence of a mass defect on the spectrum of the harmonic chain withperiodic boundaries was analyzed. All eigenfrequencies except the translation mode areeither determined by

N∑j=1

Ω2

Ω2 − Ω2j

= −Nµ, (8.34)

whereinN is the number of masses and Ω2j are the eigenvalues of the periodic chain without

mass defect, or they agree with one of the Ω2j . For a negative mass defect −1 < µ < 0

there is one mode, i.e. one solution of (8.34), outside of the domain [0,4] of the Ω2j .

(i) Calculate the eigenfrequency of this mode in the thermodynamic limit N → ∞directly from (8.34).

(ii) Consider the infinite harmonic chain with mj = m, j 6= 0 and m0 = m(1 + µ).The deviations from the equilibrium positions of the masses are denoted xn. Use theansatz

xn(t) = exp(−q |n| − iωt) (8.35)

and determine how q and ω have to be chosen, to get a solution of the equation ofmotion of the chain. Why does this solution only make sense for negative massdefects?

Page 46: Introduction to Solid State Physics - arXiv

46

8.6 Exercise 9. Harmonic oscillations of a two-dimensional lattice

Consider a two-dimensional square lattice composed of identical ions of mass m underperiodic boundary condition. Every ion interacts with nearest and next-nearest neighbours.The spring constants of the harmonic potential are given by β1 for nearest neighbours andby β2 for next-nearest neighbours. All other interactions are assumed to be negligible.Furthermore all motions of the ions are confined to the lattice plane. Set up the force-matrixKαβ(R,S) and compute κ(q). Diagonalize κ(q). How does the frequency depend on thewave vector q? Plot the dispersion relation in (q,0)- and in (q, q)-direction.

L9 Statistical mechanics of the harmonic crystal

9.1 Partition function and free energy

The thermodynamic properties of the harmonic crystal are completely determined by thecanonical partition function

Zph = tr

e−HphT

= e−

FphT . (9.1)

Here Fph is the free energy of the harmonic crystal, and we are using units such that theBoltzmann constant kB = 1. Inserting (7.21) into (9.1) we obtain

Zph =∏

(q,α)∈Q

tr

exp

−ωα(q)

T

(a+α

q aαq + 1

2

)

=∏

(q,α)∈Q

e−

ωα(q)2T

∞∑k=0

e−ωα(q)kT

=

∏(q,α)∈Q

1

2 sh(ωα(q)

2T

) . (9.2)

It follows thatFph = E0 + T

∑(q,α)∈Q

ln(

1− e−ωα(q)T

). (9.3)

Here we have used (9.1) and (7.28).

9.2 The Bose-Einstein distribution

From (9.2) and (9.3) we also conclude that

∂Fph

∂ωα(q)= − T

Zph

∂Zph

∂ωα(q)=

tr(a+α

q aαq + 1

2

)e−

HphT

tr

e−HphT

=⟨a+α

q aαq + 1

2

⟩=

1

2+

1

eωα(q)T −1

. (9.4)

Recalling thatnαq = a+α

q aαq (9.5)

is the occupation number operator that measures the occupancy of the mode (q, α), wemay interpret

〈nαq〉 =1

eωα(q)T −1

(9.6)

Page 47: Introduction to Solid State Physics - arXiv

47

as a function of T as the average occupancy of the mode (q, α) in a canonical ensemble oftemperature T . Seen as a functions of q and α this functions measures the distribution attemperature T of the quanta of vibration energy over the modes. Equation (9.6) defines thefamous Bose-Einstein distribution.

In physics we call excitations of a ‘quantum field’, that carry energy and momentum,particles. The particles associated with the quantized lattice vibrations are called phonons.Within the approximation of the harmonic crystal the phonons do not interact. Accordingto (9.6) they form an ideal gas of non-conserved Bosons, very much like the photons whichare the quanta of the electro-magnetic field. If the number of phonons would be conserved,a chemical potential that would control their number would appear in (9.6).

9.3 The density of states

In the theory of ideal quantum gases the free energy and all derived thermodynamicquantities in the thermodynamic limit are usually written as functionals of the density ofstates. We would like to briefly recall its definition and its use in approximating sums likein (9.3) by integrals. Usually the latter is done in a two-step procedure. In step one sumsover lattice momenta are converted into integrals. In step two integrals over momenta aretransformed into integrals over energies by means of the density of states.

Recall that the volume of the Brillouin zone is (cf. equation (3.10)) VR = (2π)3/Vu,where Vu is the volume of the unit cell, and that there are L3 lattice momenta in theBrillouin zone, if we introduce periodic boundary conditions as described in section 5.2.Then the volume of the crystal is V = L3Vu. The volume per lattice momentum in theBrillouin zone or volume element is

d3q =VRL3

=(2π)3

VuL3=

(2π)3

V. (9.7)

Thus, the free energy (9.3) can be approximated as

Fph = E0 +V T

(2π)3

∑α∈I

∫BZ

d3q ln(

1− e−ωα(q)T

). (9.8)

We would like to rewrite the integral on the right hand side of (9.8) as an integral overenergies. For this purpose we define the counting function

fα(ω) =1

V

∑q∈BZ

Θ(ω − ωα(q)) −−−−−→V→∞

1

(2π)3

∫BZ

d3q Θ(ω − ωα(q)) , (9.9)

where α ∈ I and Θ is the Heaviside step function. The density of states of the αth phononbranch is equal to the number of states in [ω, ω + ∆ω] divided by V∆ω or, for ∆ω → 0,

gα(ω) = f ′α(ω) =1

(2π)3

∫BZ

d3q δ(ω − ωα(q)) . (9.10)

Introducing the total density of states as

g(ω) =∑α∈I

gα(ω) (9.11)

we can rewrite the expression (9.8) for the free energy in the form

Fph = E0 + V T

∫ ∞0

dω g(ω) ln(1− e−

ωT). (9.12)

Page 48: Introduction to Solid State Physics - arXiv

48

This form will be the starting point for the discussion of the specific heat of the harmoniccrystal below.

Before entering this discussion we will have a closer look at the density of states g(ω).The integral on the right hand side of (9.10) can be interpreted as representing the ‘area’ ofa surface S(ω) in reciprocal space, where ω is implicitly defined by

σ(q0) = ω − ωα(q0) = 0 . (9.13)

In the vicinity of this surface we introduce local coordinates

∆q = n∆q⊥ + t1∆q1‖ + t2∆q2‖ , (9.14)

where n is a unit vector normal to the surface and tj∆q1‖, j = 1, 2, are parallel unitvectors. Then

σ(q0 + ∆q) = −〈gradωα(q0),∆q〉+ · · · = −‖ gradωα(q0)‖∆q⊥ + . . . (9.15)

and thus

gα(ω) =1

(2π)3

∫S(ω)

dS

∫d∆q⊥δ(−‖ gradωα(q0)‖∆q⊥ + . . . )

=1

(2π)3

∫S(ω)

dS

‖ gradωα(q0)‖ . (9.16)

This formula can be used to actually calculate the density of states, given the dispersionrelations of the phonons.

9.4 Van-Hove singularities

The density of states g(ω) exhibits singularities where gradω = 0. These are calledvan-Hove singularities. Their character depends on the space dimension and is different for3d, 2d and 1d lattices. We can understand them by expanding ω in the vicinity of a criticalpoint k0, gradω(k0) = 0, where

ω = ω0 +1

2

⟨(k− k0),Γ(k− k0)

⟩(9.17)

and Γ is the matrix of the second derivatives of ω at k0. Setting x = k − k0 we havegradω = Γ(k− k0) = Γx. Hence, for values of ω close to a critical point,

gα(ω) =1

(2π)3

∫〈x,Γx〉=2(ω−ω0)

dS

‖Γx‖ . (9.18)

The matrix Γ can be diagonalized by an orthogonal transformation. Such a transformationleaves the surface element dS invariant. Hence, we may assume that Γ = diag(γ1, γ2, γ3).Thus,

gα(ω) =1

(2π)3

∫γ1x2

1+γ2x22+γ3x2

3=2(ω−ω0)

dS√γ2

1x21 + γ2

2x22 + γ2

3x23

. (9.19)

When discussing this integral, we have distinguish several cases. The critical pointmay be a minimum, a maximum or a saddle point depending on the signature of Γ whichis defined as

sign Γ = diag(sign γ1, sign γ2, sign γ3) . (9.20)

Page 49: Introduction to Solid State Physics - arXiv

49

We denote the four different cases by m = (+,+,+), M = (−,−,−), S1 = (+,+,−)and S2 = (+,−,−). Clearly m corresponds to a minimum, M to a maximum, and S1,S2 to two kinds of saddle points. In 3d the singularities are of square root type in all fourcases. The details will be worked out in exercise 10.

Theorem 3. Van Hove [14]. In 3d the density of states g has at least one singularity oftype S1, and one of type S2. The derivative at the upper edge of the spectrum is −∞.

9.5 Exercise 10. Van-Hove singularities

The density of states per unit volume of the αth phonon branch in a crystal lattice in ddimensions is given by

gα(ω) =1

(2π)d

∫Sα(ω)

dS

‖ gradωα(k)‖ , (9.21)

where Sα(ω) is the surface ωα = const. in the reciprocal space. The total density of statesper unit volume is the sum over all branches.

Determine the four distinct types of singularities of the density of states for spacedimension d = 3. For this purpose expand ωα close to a critical point ω0,

ωα − ω0 = γ1x21 + γ2x

22 + γ3x

23 , (9.22)

and discuss the four different cases associated with different choices of the relative sign ofthe coefficients γ1, γ2 and γ3. Hint: the saddle point cases require the introduction of acut-off.

9.6 Exercise 11. Density of states in linear chains

The dimensionless dispersion relations of the monoatomic linear chain and of the linearchain with alternating masses with 0 ≤ κ < 2π and ratio of masses µ = m/M are givenby

Ω2(κ) = 4 sin2 (κ/2) and (9.23)

Ω2(κ) = 1 + µ±√

(1 + µ)2 − 4µ sin2 (κ/2) . (9.24)

Calculate and sketch the densities of states. Which types of singularities do appear?

L10 Specific heat of the harmonic crystal

Given the free energy of the phonons as a functional of the density of states (9.12) we cancalculate their internal energy,

Eph = Fph + TSph = Fph − T∂Fph

∂T

= E0 − V T 2 ∂

∂T

∫ ∞0

dω g(ω) ln(1− e−

ωT)

= E0 + V

∫ ∞0

dωg(ω)ω

eωT −1

. (10.1)

Here Sph in the first equation is the entropy of the harmonic lattice vibrations. The integrandon the right hand side of the last equation has a clear interpretation as the “density of statesg(ω) × energy ω × thermal occupation.”

Page 50: Introduction to Solid State Physics - arXiv

50

The quantity that is measured in experiments is the specific heat

CV =∂Eph

∂T= V

∫ ∞0

dω g(ω)

(ω/2T

sh(ω/2T )

)2

. (10.2)

It is a linear functional of the density of states. Due to (9.11) the specific heat of thephonons is the sum of the contributions from all branches of the dispersion relation. Fromany model for the force matrix we can calculate the dispersion ωα, then the density ofstates gα and finally the specific heat by means of the above equation.

10.1 Low-temperature specific heat

Many bulk characteristic properties of solids at low and high temperatures are ratheruniversal. A prime example, which we shall consider now, is provided by the contributionof the phonons to the specific heat. In order to understand its low-T behaviour we use(10.1) to present it as

CV = −V ∂

∂TT 2 ∂

∂T

∫ ∞0

dω g(ω) ln(1− e−

ωT). (10.3)

The density of states has a Taylor series expansion around ω = 0. Let ω0 > 0 be its radiusof convergence and fix δ such that 0 < δ < ω0. ⇒∫ ∞

0dω g(ω) ln

(1− e−

ωT)

=

∫ δ

0dω g(ω) ln

(1− e−

ωT)

+ O(T∞)

=∞∑n=0

g(n)(0)Tn+1

n!

∫ δ/T

0dx xn ln

(1− e−x

)+ O

(T∞). (10.4)

Using the Taylor expansion of the logarithm we can estimate the integral on the right handside as∫ δ/T

0dx xn ln

(1− e−x

)=

∫ ∞0

dx xn ln(1− e−x

)+ O

(T∞)

= −∞∑k=1

1

k

∫ ∞0

dx xn e−kx +O(T∞)

= −∞∑k=1

1

kn+2

∫ ∞0

dx xn e−x +O(T∞)

= −ζ(n+ 2)Γ(n+ 1) + O(T∞), (10.5)

where Γ is the gamma function and ζ Riemann’s zeta function. Inserting (10.5) and (10.4)into (10.3) we obtain the low-T expansion of the specific heat,

CV = V

∞∑n=0

g(n)(0)(n+ 1)(n+ 2)ζ(n+ 2)Tn+1 , (10.6)

which holds up to exponentially small corrections in the temperature.Equation (10.6) holds separately for every phonon branch. If we replace g by gα, we

obtain the contribution CV,α of the phonon branch number α to the specific heat. Forevery optical branch the density of states at ω = 0 is identically zero and so are all thecoefficients in its Taylor expansion around this point. Thus,

CV,α = O(T∞)

(10.7)

Page 51: Introduction to Solid State Physics - arXiv

51

for every optical branch. In other words, the low-T specific heat of the phonons is entirelydetermined by the three acoustic branches.

Consider an acoustic phonon branch with (isotropic) sound velocity v. Then

ω(q) = v‖q‖ (10.8)

for small q. The corresponding counting function (9.9) for small ω is

f(ω) =1

(2π)3

∫BZ

d3q Θ(ω − v‖q‖) =1

2π2

∫ ω/v

0dq q2 . (10.9)

Hence, for each acoustic branch with sound velocity vα, the density of states at small ω is

g(ω) =ω2

2π2v3. (10.10)

According to (10.6) the contribution to the specific heat is

CV,α =2π2V T 3

15v3α

+ O(T 4). (10.11)

Here we have used that ζ(4) = π4/90. Summing over the three acoustic branches weobtain the total low-T specific heat of the ions in harmonic approximation,

CV =2π2V T 3

5〈v3〉h+ O

(T 4), (10.12)

where 〈·〉h stands for the harmonic mean of the sound velocities.Equation (10.12) is the T 3 law for the specific heat of insulators (in conductors the

electrons contribute significantly to CV ). In simple solids it holds up to temperatures ofabout 10K. Notice that (10.12) implies that the low-T specific heat can be determined bymeasuring the sound velocities, or, taking it the other way round, that the average soundvelocity can be obtained from a specific heat measurement.

10.2 Internal energy and specify heat at high temperatures

Recall that, for |x| < 2π,

x

ex−1= 1− x

2+

∞∑n=1

B2n

(2n)!x2n , (10.13)

where the B2n are the Bernoulli numbers. This allows us to derive a convergent high-Texpansion of the internal energy of the phonon gas. Inserting (10.13) into (10.1) we obtain

Eph = E0 + V T

∫ ∞0

dω g(ω)

1− ω

2T+

∞∑n=1

B2n

(2n)!

(ωT

)2n

= E0 + 3Nat

T − 〈ω〉g

2+

∞∑n=1

B2n

(2n)!

〈ω2n〉gT 2n−1

. (10.14)

Page 52: Introduction to Solid State Physics - arXiv

52

Here Nat = NL3 is the total number of ions and 〈·〉g denotes the average with respect tothe probability density

pg(ω) =g(ω)∫∞

0 dω g(ω). (10.15)

Thus, 〈ωm〉g is the mth moment of the probability density. In the derivation of (10.14) wehave used that ∫ ∞

0dω g(ω) = lim

ω→+∞

∑α∈I

fα(ω) =3NL3

V(10.16)

which follows from (9.9).Equation (10.14) implies the high-T series representation

CV = −3Nat

∞∑n=0

(2n− 1)B2n

(2n)!〈ω2n〉gT−2n = 3Nat −

Nat〈ω2〉g4T 2

+ O(T−4

)(10.17)

for the specific heat. Here we have used that B0 = 1 and B2 = 1/6. The leading ordercontribution corresponds to the Dulong-Petit law

CV = 3Nat (10.18)

indicating a constant heat capacity at high temperature, whence the name ‘heat capacity’.Note that this high-temperature limit is approached monotonously from below.

10.3 Debye interpolation

We have seen that the specific heat of the phonon gas shows universal behaviour at lowand high temperatures. Let us seek for a simple model interpolating between these twouniversal regimes. For this purpose we take the low-frequency form of the density of statesand cut it off at a frequency ωD in such a way that the normalization condition (10.16) issatisfied. Let us further introduce an effective sound velocity

v =⟨v3⟩ 1

3h. (10.19)

Then our model density of states is

gD(ω) =3ω2

2π2v3 Θ(ωD − ω) , (10.20)

where the cut-off frequency ωD is fixed by the condition∫ ∞0

dω gD(ω) =ω3D

2π2v3 =3Nat

V(10.21)

implying that

ωD = v

(6π2Nat

V

) 13

= N13 v rR , (10.22)

where

rR =

(VR

4π/3

) 13

(10.23)

Page 53: Introduction to Solid State Physics - arXiv

53

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 20

0.2

0.4

0.6

0.8

1

1.2

T/ωD

CV/3N

at

Figure 5: Debye model of the specific heat of the phonon gas in solids.

is a reciprocal length parameter sometimes called the ‘radius of the Brillouin zone’. gD iscalled Debye density of states and the frequency ωD the Debye frequency. These notionsgo back to the Dutch-American noble laureate Peter Debye [5].

The internal energy for the Debye density of states is

Eph = E0 +9NatT

4

ω3D

∫ ωDT

0dω

ω3

eω −1. (10.24)

Defining the Debye function

D(x) =3

x3

∫ x

0dt

t3

et−1(10.25)

we may recast the internal energy as

Eph = E0 + 3NatTD(ωD/T ) , (10.26)

while the corresponding specific heat takes the form

CV = 3Nat

D(ωDT

)− ωD

TD′(ωDT

). (10.27)

This is called the Debye formula or the Debye interpolation formula. According tothis formula the specific heat of the phonon gas is a universal function of ωD/T (seeFigure 5). With respect to the phonon gas contribution to the specific heat different solidsare distinguished by a single parameter, the Debye frequency. Because it is simple and atthe same time covers the basic features of the temperature dependence of the specific heat,the Debye model is prevailing in experimental solid state physics.

10.4 The anomaly of the harmonic crystal

Usually the basic tasks in statistical physics are to derive the specific heat and the equationof state, expressing the pressure p = −∂F/∂V as a function of T and V . For the harmoniccrystal one can find the statement that the pressure is temperature independent,( ∂p

∂T

)V

= 0 , (10.28)

Page 54: Introduction to Solid State Physics - arXiv

54

in several places in the textbook literature. This should be related to the scale invariance ofthe equations of motion in harmonic potentials, but at the moment I do not know of anyreally convincing derivation in the general case.

Equation (10.28) implies several thermodynamic anomalies. Since(∂V∂T

)p

= −( ∂p∂T

)V( ∂p

∂V

)T

= 0 , (10.29)

it follows, for instance, that the thermal expansion coefficient

α =1

V

(∂V∂T

)p

= 0 . (10.30)

This contradicts our experience that solids usually expand when they are heated. Weconclude that the thermal expansion of solids is a feature that must be attributed the higherorder corrections to the harmonic crystal. Another anomaly implied by (10.28) is, forinstance, CV = Cp, the coincidence of the specific heats at constant pressure and constantvolume.

The simplest model system for which we can verify (10.28) is a classical chain ofparticles of equal masses m that interact with their nearest neighbours through a pairpotential V (r), where r is the distance between the particles. Let us consider N + 1such particles with coordinates xn, n = 0, . . . , N , and nearest-neighbour distances rn =xn − xn−1, n = 1, . . . , N . Then the Hamiltonian of the system is

H =N∑n=0

p2n

2m+

N∑n=1

V (rn) , (10.31)

where the pn are the momenta canonically conjugate to the xn. Note that this systemhas no periodic boundary conditions. Instead of applying periodic boundary conditionsfor the interaction we apply an external mechanic pressure to control the length of thesystem. This can be achieved by adding a term (xN − x0)p to the Hamiltonian (10.31). Asa configuration space K for the particles we consider a ring of finite length L on which the‘springs’ connecting neighbouring particles can be arbitrarily expanded by moving themseveral times relative to each other around the ring, but the center of mass of all particles isconfined to the ring. This construction is necessary in order to regularize the integral

Z =1

(N + 1)!

∫RN+1

dN+1p

(2π~)N+1

∫K

dN+1x exp

− 1

T

(H + (xN − x0)p

)(10.32)

representing the classical partition function. It would be otherwise divergent. By construc-tion

−T∂p lnZ = 〈xN − x0〉 = ` (10.33)

is the average length of the system at given T and p which justifies the interpretation of pas the pressure.

It is not difficult to calculate the integral on the right hand side of (10.32). Themomentum integration reduces to Gaussian integrals, whereas the remaining integrals overthe configuration space can be dealt with after introducing Jacobi coordinates

s =1

N + 1

N∑n=0

xn , rn = xn − xn−1 , n = 1, . . . , N . (10.34)

Page 55: Introduction to Solid State Physics - arXiv

55

This transformation is linear, and it is easy to see that its Jacobi matrix equals one. Takingthis into account we obtain

Z =L

(N + 1)!

(mT

2π~2

)N+12(∫ ∞−∞

dr e−(V (r)+pr)/T

)N. (10.35)

It follows that

` = −NT∂p ln

(∫ ∞−∞

dr e−(V (r)+pr)/T

). (10.36)

In general the latter expression does depend on T , but if we insert the harmonicpotential

V (r) =mω2

0

2(r − a)2 , (10.37)

where a is the ‘rest length’ of the potential, then

` = N

(a− p

mω20

)(10.38)

which we interpret as the thermodynamic length of the system. If we now solve for thepressure p, we end up with

p = mω20

(a− `

N

)(10.39)

which, indeed, does not depend on T . Note that, unlike the pressure of gases, the pressurehere can have either sign, a negative pressure contracting the chain, if it is expanded overits equilibrium length.

L11 Neutron scattering

11.1 Thermal neutrons

Neutron scattering experiments in solids are performed with so-called thermal neutrons.These are neutrons with energies between 5 - 10 meV (T ' 60 - 1000K, λ ' 0,4 - 0,1 nm).Thermal neutrons are one of the most important probes for studying the properties of solidsand fluids. This has several reasons.

(i) The neutron is electrically neutral. It penetrates deeply into the solid, can come closeto the atomics nuclei and is scattered by nuclear forces.

(ii) Because of their relatively large mass the de Broglie wave length of thermal neutronsis of the order of magnitude of atomic distances in solids and fluids.

(iii) Their energy is of the order of magnitude of the energy of elementary excitations insolids. If a neutron is scattered inelastically its relative change of energy is generallylarge enough to be resolved experimentally. Hence, neutrons do not only resolve thestructure of solids, but can be also used to measure their excitation spectra, e.g. thedispersion relations of phonons.

(iv) Neutrons carry a magnetic moment which is sensitive against intra-atomic magneticfields. Therefore they can be used to measure short-wavelength magnetic structures(antiferromagnetism!) and magnetic excitations.

Page 56: Introduction to Solid State Physics - arXiv

56

target

counter

source

(energy resolved)

current density‖j‖ = vρ =number of particlescrossing unit area Aper unit time

assumed stationary (counts per unit time)∆n(ϑ, ϕ,E) = counting rate

Figure 6: Schematic view of a typical scattering experiment. An incident particle currenthits a target and is scattered in all directions. The currents of the scattered particles througha sphere centered around the target are measured. They determine the differential crosssection.

The traditional source of neutrons are nuclear reactors. Note that neutrons which come outof a nuclear reactor have a Maxwell velocity distribution leading to a neutron current perunit time for neutrons with velocity v of

φ(v)dv =ρ√π

( m2T

) 32v3 e−

mv2

2T dv , (11.1)

where m is the neutron mass and ρ their density (exercise: derive (11.1)).

11.2 Scattering experiments and cross sections

The typical geometry of a scattering experiment is sketched in Figure 6. A stationarycurrent of incident particles impinges on a target and is scattered. Detectors in equaldistance from the target measure the current scattered in direction (ϑ, ϕ), where ϑ, ϕ arespherical coordinates. We denote the counting rate at (ϑ, ϕ) for particles with energies inan infinitesimal interval around E by ∆n(ϑ, ϕ,E). In a stationary situation the countingrate is expected to be proportional to the current density ‖j‖ of the incident current j.Hence, the counting rate normalized by the current density of the incident current,

∆σ =∆n

‖j‖ , (11.2)

is expected to be independent of the current and to characterize the target. If we furtherdivide by the solid angle ∆Ω = sin(ϑ)∆ϑ∆ϕ and by the width ∆E of the energy interval,we obtain a quantity called the differential cross section,

∆σ

∆Ω∆E=

∆n

∆Ω∆E‖j‖ . (11.3)

In the limit of infinite energy and angle resolution this function is denoted

d2σ

dΩdE(ϑ, ϕ,E) . (11.4)

Page 57: Introduction to Solid State Physics - arXiv

57

11.3 Cross section and transition rate

We shall assume that the incident neutron beam is monochromatic and coherent. Quantummechanically it is then represented by a plane wave

Ψ(x) =ei〈k,x〉√V

(11.5)

with wave vector k and ‘normalization volume V ’ (we shall use units such that ~ = 1).The corresponding current is

j(x) = − i

2m

(Ψ∗(x)∂xΨ(x)−Ψ(x)∂xΨ∗(x)

)=

k

mV. (11.6)

The interaction of the neutrons with the target determines the transition rate P (k,k′)for a transition from an incoming wave with wave vector k to a scattered wave withwave vector k′ under the influence of the perturbation caused by the target (recall thetime-dependent perturbation theory from the quantum mechanics lecture). Since d3k′

(2π)3/V

is the number of states in the volume d3k′ around k′ we can express the counting rate as

dn = P (k,k′)d3k′

(2π)3/V= P (k,k′)

V k′2dk′dΩ

(2π)3=

mV

(2π)3P (k,k′)k′dEdΩ . (11.7)

Using (11.6) and (11.7) in (11.3) we have expressed the differential cross section in termsof the transition rate,

d2σ

dΩdE(ϑ, ϕ,E) =

k′

k

(mV )2

(2π)3P (k,k′) . (11.8)

This formula connects the basic measurable quantity on the left hand side with a quantitydepending on the microscopic properties of the target on the right hand side.

Let us recall how the transition rate appears in quantum mechanics. It is usually firstencountered in the context of time-dependent perturbation theory and Fermi’s ‘golden rule’which states in our case that

P (k,k′) = 2π∑f

δ(εi − εf )|〈Ψf , UΨi〉|2 (11.9)

is the rate for transitions Ψi → Ψf , where Ψi = Ψk(x)φi is a fixed initial state andΨf = Ψk′(x)φf runs through all possible final states with fixed k′. Here φi and φf denoteeigenstates of the ions (recall that the neutrons interact with the ions!). Hence, the energiesof initial and final states εi and εf are

εi = Ei +k2

2m, εf = Ef +

k′2

2m, (11.10)

where Ei and Ef are the energies of the ionic states. Introducing the notation ω =(k2 − k′2)/2m the energy difference in (11.9) takes the form

εi − εf = Ei − Ef + ω . (11.11)

The potential U that describes the interaction of the neutrons with the ions is of the form

U(x) =∑R∈B

N∑r=1

u(x− xr(R)

), (11.12)

where the xr(R) are the position vectors of the ions. In order to simplify the notation wesuppress from now on the sum over the ion positions in the unit cell. It will always comewith the sum over the Bravais lattice an can be restored at any later stage if necessary.

Page 58: Introduction to Solid State Physics - arXiv

58

11.4 Form factor and structure factor

Substituting the explicit form of the potential and of the factorized wave functions we cancalculate the matrix elements in (11.9),

〈Ψi, UΨf 〉 =1

V

∑R∈B

∫d3x e−i〈q,x〉

∫ [∏S∈B

d3x(S)

]φ∗i u

(x− x(R)

)φf

=1

V

∑R∈B

∫d3x e−i〈q,x〉 u(x)

∫ [∏S∈B

d3x(S)

]φ∗i e−i〈q,x(R)〉 φf , (11.13)

where q = k − k′. We see that the integrals factorize into a factor that depends on theinteraction potential between the neutrons and the ions and a factors that depends on thecrystal structure. Denoting

b(q) =

∫d3x e−i〈q,x〉 u(x) , (11.14a)

⟨φi, e

−i〈q,x(R)〉 φf⟩

=

∫ [∏S∈B

d3x(S)

]φ∗i e−i〈q,x(R)〉 φf (11.14b)

we obtain the following factorized form of the transition rate,

P (k,k′) =2π

V 2|b(q)|2

∑f

δ(εi − εf )∣∣∣∑R∈B

⟨φi, e

−i〈q,x(R)〉 φf⟩∣∣∣2 . (11.15)

Here |b(q)|2 is called the atomic form factor, since it depends only on the interaction of theindividual ions with the neutrons. The sum over f divided by Nat is called the dynamicstructure factor S(q, ω). It contains the information about the structure and the dynamicsof the ions. Inserting the expression (11.15) for the transition rate into (11.8) we obtain aformula for the differential cross section,

d2σ

dΩdE(ϑ, ϕ,E) =

k′

k

m2

(2π)2|b(q)|2

∑f

δ(εi − εf )∣∣∣∑R∈B

⟨φi, e

−i〈q,x(R)〉 φf⟩∣∣∣2 . (11.16)

Note that this formula is very general.

(i) With slight modifications it holds for fluids as well.

(ii) It solely relies on Fermi’s golden rule (= time dependent perturbation theory + Bornapproximation).

11.5 Rewriting the structure factor

Using the ‘Fourier inversion formula’

δ(ω) =

∫ ∞−∞

dt

2πe−iωt (11.17)

we obtain the following expression for the dynamic structure factor.

S(q, ω) =1

Nat

∫ ∞−∞

dt

∑f

e−i(Ei−Ef+ω)t∑

R,S∈B

⟨φi, e

−i〈q,x(S)〉 φf⟩⟨φf , e

i〈q,x(R)〉 φi⟩

Page 59: Introduction to Solid State Physics - arXiv

59

=1

Nat

∫ ∞−∞

dt

2πe−iωt

∑R,S∈B

⟨φi, e

−i〈q,x(S)〉 ei〈q,x(R,t)〉 φi⟩. (11.18)

Here we have used that

e−i(Ei−Ef )t⟨φf , e

i〈q,x(R)〉 φi⟩

=⟨φf , e

iHt ei〈q,x(R)〉 e−iHt φi⟩

=⟨φf , e

i〈q,x(R,t)〉 φi⟩, (11.19)

where H is an effective Hamiltonian for the ion-ion interaction. The expectation valueunder the sum on the right hand side of (11.18) is called a dynamical two-point correlationfunction. The appearance of such a correlation function is generic. As we shall see withfurther examples below, most spectroscopic experiments and transport experiments insolids measure two- or four-point correlation functions.

L12 Neutron scattering continued

12.1 Disorder

So far we have assumed that the scattering potentials of all nuclei are equal. This is onlytrue if the solid (the fluid) consists of a single isotope and if all nuclear spins are aligned(nuclear spin ferromagnet) or zero. In reality the latter conditions are almost never satisfied.We typically rather have to deal with mixtures of different isotopes and with nuclear spinparamagnets. This means that we have to modify our above derivation, replacing

u(x− x(R))→ uR(x− x(R)) , b(q)→ bR(q) . (12.1)

In particular bR(q) remains under the sum over the Bravais lattice in the expression for thedifferential cross section, which now reads

d2σ

dΩdE(ϑ, ϕ,E)

=k′

k

m2

(2π)2

∫ ∞−∞

dt

2πe−iωt

∑R,S∈B

bR(q)b∗S(q)⟨φi, e

−i〈q,x(S)〉 ei〈q,x(R,t)〉 φi⟩. (12.2)

In order to simplify this expression again, we assume that the potentials uR arerandomly distributed. We shall indicate the disorder average by brackets 〈·〉d. We assumethat the distribution underlying the average is such that the mean value of bR(q) istranslation invariant and, for this reason, use the notation

〈b(q)〉d = 〈bR(q)〉d . (12.3)

It further reasonable to assume that potentials at different lattice sites are uncorrelated,implying that

〈bR(q)b∗S(q)〉d = 〈bR(q)〉d〈b∗S(q)〉d = |〈b(q)〉d|2 (12.4)

for R 6= S. Let us further introduce the averaged coherent and incoherent atomic formfactors

σcoh =m2

π|〈b(q)〉d|2 , σinc =

m2

π

(〈|b(q)|2〉d − |〈b(q)〉d|2

). (12.5)

Page 60: Introduction to Solid State Physics - arXiv

60

Averaging the differential cross section over the disorder then results in

d2σ

dΩdE(ϑ, ϕ,E) =

k′

k

∫ ∞−∞

dt

2πe−iωt

∑R,S∈B

σcoh

⟨φi, e

−i〈q,x(S)〉 ei〈q,x(R,t)〉 φi⟩

+∑R∈B

σinc

⟨φi, e

−i〈q,x(R)〉 ei〈q,x(R,t)〉 φi⟩

. (12.6)

So far we have assumed that the ions are in a pure initial state φi. This is not realisticin experiments on a macroscopic sample. In a macroscopic sample the ions have to bedescribed by a density matrix. If the crystal can exchange energy with its environment itwill be the density matrix of the canonical ensemble in which each state φi is occupiedwith probability e−Ei/T /Z, where Z is the canonical partition function. Performing thecanonical ensemble average and denoting the canonical expectation values by 〈·〉T weobtain the final formula for the theory of neutron scattering.

12.2 Summary of the theory of neutron scattering – general case

First of all the dynamic structure factor at finite temperature becomes

S(q, ω) =1

Nat

∫ ∞−∞

dt

2πe−iωt

∑R,S∈B

⟨e−i〈q,x(S)〉 ei〈q,x(R,t)〉⟩

T. (12.7)

The differential cross section splits into a coherent part and an incoherent part,

d2σ

dΩdE=

d2σ

dΩdE

∣∣∣∣coh

+d2σ

dΩdE

∣∣∣∣inc

, (12.8)

where

d2σ

dΩdE

∣∣∣∣coh

=σcoh

k′

kNatS(q, ω) , (12.9a)

d2σ

dΩdE

∣∣∣∣inc

=σinc

k′

k

∫ ∞−∞

dt

2πe−iωt

∑R∈B

⟨e−i〈q,x(R)〉 ei〈q,x(R,t)〉⟩

T. (12.9b)

These formulae are widely used in order to analyze the data of neutron scatteringexperiments. Before specializing them to crystal structures we would like make a fewgeneral comments.

(i) In our derivation of (12.7) we have used the letters R and S in x(R), x(S) as mereparticle labels, which is the reason why equation (12.7), if properly read, also definesthe dynamic structure factor of a fluid or a glass. When we used the notation R ∈ Bwe had in mind to apply the formula to a mono-atomic (simple) lattice, but so far Bwas rather an index set which may refer to any labeling of the ions. Similar formulaehold, in particular, for a lattice with basis.

(ii) The factors σcoh and σinc depend, in the experimentally relevant range of neutronwave lengths of the order of inter-atomic distances only weakly on q, since thepotentials uR vary on a scale of the size of the nuclei. For this reason σcoh and σinc

can often be treated as constants.

Page 61: Introduction to Solid State Physics - arXiv

61

(iii) The most interesting contribution to the differential cross section is the dynamicstructure factor S(q, ω). It is completely determined by the properties of the sampleand independent of the properties of the neutrons.

(iv) The incoherent cross section sums contributions from the same nuclei at differenttimes. It carries no information about the structure of the sample. If all bR(q) areidentical, the incoherent part vanishes. Its occurance is a direct consequence of thedisorder in the system.

12.3 Theory of neutron scattering – specialization to crystal structures

What happens if we specialize (12.7) to a mono-atomic crystal is, that the Hamiltonian isthen invariant under the action of the Bravais lattice B, implying that⟨

e−i〈q,x(S)〉 ei〈q,x(R,t)〉⟩T

=⟨e−i〈q,x(0)〉 ei〈q,x(R−S,t)〉⟩

T. (12.10)

Inserting this into (12.7) and also using that, x(R) = R + u(R), where u(R) is thedeviation from the equilibrium position at R, we obtain

S(q, ω) =

∫ ∞−∞

dt

2πe−iωt

∑R∈B

ei〈q,R〉⟨e−i〈q,u(0)〉 ei〈q,u(R,t)〉⟩T. (12.11)

The dynamic structure factor of the ions in a crystal lattice is the spatio-temporal Fouriertransform of the dynamical two-point function

⟨e−i〈q,u(0)〉 ei〈q,u(R,t)〉⟩

T.

Similarly, the incoherent part of the differential cross section simplifies to

d2σ

dΩdE

∣∣∣∣inc

=σinc

k′

kNat

∫ ∞−∞

dt

2πe−iωt

⟨e−i〈q,u(0)〉 ei〈q,u(0,t)〉⟩

T, (12.12)

the Fourier transformation in time of the auto correlation function⟨e−i〈q,u(0)〉 ei〈q,u(0,t)〉⟩

T.

12.4 Dynamic structure factor of the harmonic crystal

The harmonic approximation brings more simplifications about. If A and B are any twooperators that depend linearly on the deviations u(R) of the ions from their equilibriumposition and linearly on the conjugate momentum p(R), then⟨

eA eB⟩T

= e12〈A2+2AB+B2〉T , (12.13)

where the canonical averages are calculated with the Hamiltonian of the harmonic crystal(see exercise 12.8). Applying this formula to A = −i〈q,u(0)〉 and B = i〈q,u(R, t)〉 anddenoting

2W (q) =⟨〈q,u(0)〉2

⟩T

=⟨〈q,u(R, t)〉2

⟩T, (12.14)

we obtain the following formula for the dynamic structure factor of the harmonic crystal,

S(q, ω) =e−2W (q)

∫ ∞−∞

dt e−iωt∑R∈B

ei〈q,R〉 e〈〈q,u(0)〉〈q,u(R,t)〉〉T . (12.15)

The function W (q) is called the Debye-Waller factor.

Page 62: Introduction to Solid State Physics - arXiv

62

12.5 Expansion into multi-phonon processes

A harmonic crystal has 3Nat eigenstates which can be excited independently. Denotethe corresponding occupation numbers n1, . . . , n3Nat . Transition matrix elements can beclassified according to the differences in occupation numbers between the initial state niand final state n′i.

Elastic processes. ni = n′i, all occupation numbers remain unaltered, no exchange ofenergy between crystal and neutron.

Single-phonon processes. ∃ α ∈ 1, . . . , 3Nat such that ni = n′i ∀ i 6= α, n′α = nα±1,the occupation number of a single mode is changed due to the scattering process.

Multi-phonon processes. The occupation numbers of several modes are changed.

We shall see below that the nth term in the expansion

e〈〈q,u(0)〉〈q,u(R,t)〉〉T = 1 +⟨. . .⟩T

+1

2

⟨. . .⟩2

T+ . . . (12.16)

can be identified with the n-phonon processes.

12.6 Coherent elastic neutron scattering

The contribution of the first term in (12.16) to the dynamic structure factor is

S0(q, ω) = Nat e−2W (q) δ(ω)∑K∈B

δK,q . (12.17)

Recall that the argument of the delta function is ω = (k2 − k′2)/2m. Hence, the conditionω = 0 imposed by the delta function in (12.17) is the condition of energy conservationfor the scattered neutrons which justifies the interpretation of the zeroth order term as theelastic scattering term. Since k, k′ > 0, the zeroth order structure factor (12.17) can onlybe non-vanishing if k = k′. Consequentially, we obtain the expression

d2σ

dΩdE

∣∣∣∣coh

=σcoh

4πN2

at e−2W (k−k′) δ(ω)∑K∈B

δK,k−k′ (12.18)

for the elastic contribution to the coherent differential cross section of the harmonic crystal.The sum over Kronecker deltas on the right hand side means that elastic scattering takesplace, if the wave vectors k of the incident neutron beam and k′ of the scattered beamdiffer by a reciprocal lattice vector K,

k− k′ = K ∈ B . (12.19)

This is the famous Bragg condition which also holds in X-ray diffraction experiments.

12.7 Interpretation of the Bragg condition

The interpretation of the Bragg condition in reciprocal space is depicted in Figure 7. TheBragg condition is often formulated as a relation between the distance d of lattice planesin the original Bravais lattice, the angle ϑ between incident and scattered neutron and

Page 63: Introduction to Solid State Physics - arXiv

63

k k′

k− k′ 6= K

ϑ

no elastic scattering

k

beamdirection

observer

ϑ k′

K

Figure 7: Illustration of the Bragg condition (12.19) in the reciprocal space.

the wave length λ = 2π/k of the incident neutron. Recalling from section 3.3 that thereciprocal lattice vector K corresponds to a family of lattice planes perpendicular to K insuch a way that K = ‖K‖ = n · 2π/d for some n ∈ N and taking into account that elasticscattering implies k = k′, we conclude that

K2 = k2 + k′2 − 2kk′ cos(ϑ)

= k24 sin2(ϑ/2) = (n2π/d)2

⇒ nλ = 2d sin(ϑ/2) . (12.20)

Relative to the Bravais lattice we can interpret theBragg condition in the following way.

(i) The scattering occurs as if it would happen atthe lattice planes according to the reflection lawof geometric optics.

(ii) Waves which are reflected by parallel planes in-terfere constructively, meaning the differencein optical distance between two ‘rays’ reflectedfrom consecutive lattice planes must be an inte-ger multiple of the wave length,

2`− b = 2`− 2a cos(ϑ/2)

= 2`− 2` cos2(ϑ/2) = 2` sin2(ϑ/2)

= 2d sin(ϑ/2) = nλ . (12.21)

k′

K

k

b

ad

ϑ2

12.8 Exercise 12. A proof of equation (12.13)

In the derivation of the formula (12.15) for the dynamic structure factor of the harmoniccrystal we have used equation (12.13). Here we would like to provide a step-by-stepderivation (cf. [11]).

Consider a system of harmonic oscillators with Hamiltonian

H =∑i

ωi(a+i ai + 1

2

), (12.22)

Page 64: Introduction to Solid State Physics - arXiv

64

where the ai and a+j are annihilation and creation operators with commutation relations

[ai, aj ] = 0 = [a+i , a

+j ], [ai, a

+j ] = δij . Let

A =∑i

(αiai + βia

+i

), B =

∑i

(γiai + δia

+i

), (12.23)

where αi, βj , γk, δl ∈ C. Denote the canonical ensemble average with Hamiltonian (12.22)by 〈·〉. Our goal is to prove that⟨

eA eB⟩

= e12〈A2+2AB+B2〉 . (12.24)

The proof will be based on the formula

eA eB = e12

[A,B] eA+B (12.25)

which should be familiar from the introductory lecture on quantum mechanics and followsfrom the Baker-Campbell-Hausdorff formula (exercise: recall the derivation of bothformulae). It holds whenever adA[A,B] = adB[A,B] = 0 (where the adjoint action of anoperator X on any operator Y is defined by adX Y = [X,Y ]).

(i) DefineC =

∑i

ciai , D =∑i

dia+i . (12.26)

Use (12.25) to show that

⟨eC eD

⟩= e[C,D]

⟨eD eC

⟩= e[C,D]

⟨ee

adH/T C eD⟩. (12.27)

(ii) LetCn = en adH/T C , n ∈ N0 . (12.28)

Show that ⟨eC eD

⟩= e[

∑nk=0 Ck,D]

⟨eCn+1 eD

⟩. (12.29)

(iii) Show that (formally) limn→∞Cn = 0 and that the series∑∞

k=0Cn has a formallimit. Denoting this limit by S and using that

⟨eD⟩

= 1, equation (12.29) thenimplies that ⟨

eC eD⟩

= e[S,D] . (12.30)

(iv) Show in a similar way as above that

〈CD〉 = [S,D] . (12.31)

Thus, ⟨eC eD

⟩= e〈CD〉 . (12.32)

(v) Use (12.25) and (12.32) to prove (12.24).

Page 65: Introduction to Solid State Physics - arXiv

65

L13 Inelastic neutron scattering

13.1 Debye-Waller factor

If we solve equations (7.25) for uα(R) we obtain

uα(R) =1√L3

∑(q,β)∈Q

ei〈q,R〉 Y αβ (q)

aβq + a+β−q√

2µαωβ(q), (13.1)

the deviations of the ions from their equilibrium positions in a center of mass frame ofreference (the zero modes are excluded in the summation, see section 7.2).

Let us assume for simplicity that we are dealing with a mono-atomic lattice. Then theindices α, β in (13.1) take values x, y, z, the dimensionless masses reduce to µα = 1, andL3 = Nat. The scalar products defining the Debye-Waller factor (12.14) become

〈q,u(0)〉 =1√Nat

∑k∈BZk 6=0

∑α=x,y,z

〈q,yα(k)〉√2ωα(k)

(aαk + a+α

−k), (13.2)

where the yα(k) are the polarization vectors defined below (7.31). It follows that

2W (q) =⟨〈q,u(0)〉2

⟩T

=1

Nat

∑k,k′∈BZk,k′ 6=0

∑α,β=x,y,z

〈q,yα(k)〉√2ωα(k)

〈q,yβ(k′)〉√2ωβ(k′)

⟨(aαk + a+α

−k)(aβk′ + a+β

−k′)⟩T

=1

Nat

∑k∈BZk 6=0

∑α=x,y,z

〈q,yα(k)〉〈q,yα(−k)〉2ωα(k)

cth(ωα(k)

2T

)

=1

Nat

∑k∈BZk6=0

∑α=x,y,z

|〈q,yα(k)〉|22ωα(k)

cth(ωα(k)

2T

). (13.3)

Here we have used (cf. (9.6))⟨(aαk + a+α

−k)(aβk′ + a+β

−k′)⟩T

= δαβδk,−k′⟨aαka

+αk + a+α

kaαk

⟩T

= δαβδk,−k′⟨2nαk + 1

⟩T

= δαβδk,−k′ cth(ωα(k)

2T

)(13.4)

in the second equation and (6.35) in the third equation. Taking the thermodynamic limit in(13.3) we end up with

2W (q) =Vu

(2π)3

∑α=x,y,z

∫BZ

d3k|〈q,yα(k)〉|2

2ωα(k)cth(ωα(k)

2T

). (13.5)

Here two remarks are in order.

(i) In general (13.5) cannot be rewritten by means of the density of states, since thepolarization vectors yα(k) depend on k not necessarily through ωα(k).

Page 66: Introduction to Solid State Physics - arXiv

66

(ii) The Debye-Waller factor describes the weakening of the coherent scattering dueto thermal fluctuation and as a function of the change of momentum q, i.e. as afunction of the scattering angle. It appears as a prefactor e−2W (q) in the dynamicstructure factor. Since W (q) is quadratic in q, the weakening is larger for larger q.In elastic scattering, where q must equal a reciprocal lattice vector K, the scatteringis strong in the direction corresponding to minimal K or maximal distance betweenthe associated family of lattice planes. Recall (cf. exercise 4.7) that this family hasmaximal density of lattice points within a plane.

As a function of the temperature the scattered intensity weakens with growing T ,since

cth(ωα(k)

2T

)∼ 2T

ωα(k)(13.6)

as T →∞, leading to a linear temperature dependence ofW (q) and to a suppressionof the scattered intensity exponentially in T .

13.2 Exercise 13. Debye-Waller factor of a cubic lattice

For a cubic lattice consisting of N atoms the Debye-Waller factor can be written as

2W (q) =V q2

N

∫ ∞0

ω2E(ω)g(ω) , (13.7)

wherein g(ω) is the density of states (per branch), q is the difference of momenta and E(ω)is the average of the energy of the phonons with frequency ω,

E(ω) = ω

[n(ω, T ) +

1

2

], n(ω, T ) = [exp(ω/T )− 1]−1 . (13.8)

(i) Calculate the Debye-Waller factor for a cubic Bravais lattice by using the Einsteinmodel for the phonons. This is a model for an optical phonon branch, where allions oscillate with the same frequency ω0; thus all N states are located at ω0, andthe density of states is given by g(ω) = (N/V )δ(ω − ω0). Determine 2W (q) forT ω0 or T ω0, respectively.

(ii) What is the Debye-Waller factor for the Debye model? What follows for 2W (q) inthe limits T ωD and T ωD?

13.3 Coherent inelastic scattering – the single-phonon contribution

Let us now consider the contribution of the second term in (12.16) to the dynamic structurefactor,

S1(q, ω) =e−2W (q)

∫ ∞−∞

dt e−iωt∑R∈B

ei〈q,R〉〈〈q,u(0)〉〈q,u(R, t)〉〉T . (13.9)

As we shall see, this term corresponds to the single-phonon contributions. This can beunderstood from a closer inspection of the two-point function on the right hand side. Weinfer from (13.1) that

〈q,u(R)〉 =1√Nat

∑k∈BZk 6=0

∑α=x,y,z

ei〈k,R〉 〈q,yα(k)〉√2ωα(k)

(aαk + a+α

−k). (13.10)

Page 67: Introduction to Solid State Physics - arXiv

67

This evolves in time into

〈q,u(R, t)〉 = eiHt〈q,u(R)〉 e−iHt

=1√Nat

∑k∈BZk 6=0

∑α=x,y,z

ei〈k,R〉 〈q,yα(k)〉√2ωα(k)

(aαk e−iωα(k)t +a+α

−k eiωα(k)t). (13.11)

Hence, performing a similar calculation as in (13.3),

〈〈q,u(0)〉〈q,u(R, t)〉〉T (13.12)

=1

Nat

∑k,k′∈BZk,k′ 6=0

∑α,β=x,y,z

ei〈k,R〉 〈q,yα(k)〉√2ωα(k)

〈q,yβ(k′)〉√2ωβ(k′)

× δαβδk,−k′(〈nαk〉T e−iωα(k)t +〈nαk + 1〉T eiωα(k)t

)=

1

Nat

∑k∈BZk 6=0

∑α=x,y,z

ei〈k,R〉 |〈q,yα(k)〉|22ωα(k)

e−iωα(k)t

eωα(k)/T −1+

eiωα(k)t

1− e−ωα(k)/T

=1

Nat

∑k∈BZk 6=0

∑α=x,y,z

|〈q,yα(k)〉|22ωα(k)

ei(〈k,R〉−ωα(k)t)

eωα(k)/T −1+

e−i(〈k,R〉−ωα(k)t)

1− e−ωα(k)/T

. (13.13)

Here we have substituted −k for k in the second sum in the last equation and have usedthat ωα(k) and |〈q,yα(k)〉|2 are even functions of k.

Inserting the latter equation into (13.9) we obtain the single-phonon contribution to thedynamic structure factor in the form

S1(q, ω) = S1,+(q, ω) + S1,−(q, ω) , (13.14)

where

S1,+(q, ω) =e−2W (q)

∫ ∞−∞

dt e−iωt∑R∈B

ei〈q,R〉

× 1

Nat

∑q′∈BZq′ 6=0

∑α=x,y,z

|〈q,yα(q′)〉|22ωα(q′)

e−i(〈q′,R〉−ωα(q′)t)

1− e−ωα(q′)/T

= e−2W (q)∑

q′∈BZq′ 6=0

∑α=x,y,z

|〈q,yα(q′)〉|22ωα(q′)

δ(ω − ωα(q′))

1− e−ωα(q′)/T

∑K∈B

δK,q−q′ (13.15)

and

S1,−(q, ω)

= e−2W (q)∑

q′∈BZq′ 6=0

∑α=x,y,z

|〈q,yα(q′)〉|22ωα(q′)

δ(ω + ωα(q′))

eωα(q′)/T −1

∑K∈B

δK,q+q′ . (13.16)

Page 68: Introduction to Solid State Physics - arXiv

68

These are those contributions to the dynamic structure factor, where precisely onephonon is excited (emitted) or absorbed. We shall denote the corresponding contributionsto the differential cross section

d2σ

dΩdE

∣∣∣∣coh,1±

=σcoh

k′

kS1,±(q, ω) . (13.17)

Like the elastic cross section these cross sections are ‘discontinuous’ and describe apattern of bright spots at certain energy and momentum transfers. The delta functionsand Kronecker deltas in (13.15) and (13.16) force the dispersion relations of the scatteredneutron and the phonons to match in the following sense.

Emission of a phonon. In S1,+ the momenta of the neutron and the phonon involved inthe scattering process must satisfy

q = k− k′ = q′ + K , (13.18)

where k− k′ is the momentum transfer to the lattice and q′ is the momentum takenby the phonon, i.e. the phonon takes the momentum transferred to the lattice andreduced to the Brillouin zone. For the energies we have the matching condition

ω =k2 − k′2

2m= ωα(q′) = ωα(q−K) = ωα(k− k′) , (13.19)

where we have used the periodicity of ωα with respect to the reciprocal lattice in thelast equation. Since ωα ≥ 0 we see that k ≥ k′ in this process. The lattice absorbsenergy, ‘a phonon is emitted.’

The temperature dependence of the emission is encoded in the factor 1/(1 −e−ωα(q′)/T ) in S1,+. This factor decreases as T decreases but stays finite forT → 0+.

Absorption of a phonon. For S1,− the momentum balance is

q = k− k′ = −q′ + K (13.20)

which we interpret such that the neutron takes momentum q′ −K. The relation forthe exchange of energy between the scattered neutron an the phonon involved in theprocess is

−ω =k′2 − k2

2m= ωα(q′) = ωα(K− q) = ωα(k′ − k) . (13.21)

Here k′ ≥ k, and energy is absorbed by the neutron, which is interpreted as a phononbeing absorbed in the process.

In this case the temperature dependence comes in through a factor 1/(eωα(q′)/T −1)which vanishes as T → 0+. At zero temperature no phonon remains in the systemand there is nothing left to be absorbed.

Page 69: Introduction to Solid State Physics - arXiv

69

0.5 1 1.5 2 2.5 3 3.5

−2

0

2

4

k′

12m(k2 − k′2)12m(k′2 − k2)

ω`(k− k′n)

ωt(k− k′n)counting rate

Figure 8: Schematic picture explaining how phonon dispersion relations can be inferredfrom the differential cross section.

13.4 Measuring dispersion relations of phonons

Dispersion relations of phonons can be inferred from the differential cross section. Assumethat a beam of mono-energetic neutrons impinges on a crystal and the velocity distributionof the neutrons scattered in a fixed direction n is measured . Strong scattering in direction ntakes place whenever k′ = k′n satisfies one of the resonance conditions (13.19) or (13.21)for emission or absorption. The situation is then as sketched in Figure 8. In the example wehave assumed a simple lattice with a longitudinal acoustic phonon with dispersion ω` (grayline) and two degenerate transversal acoustic phonons with dispersion ωt (dashed gray line).The blue line represents the energy loss of the scattered neutron if it has momentum k′.The scattering is likely only if there is a phonon of momentum k− k′n modulo reciprocallattice vectors, which can be emitted at this energy. Similarly, the dashed blue line showsthe energy gain of the neutron if a phonon is absorbed in the scattering process. One of theresonance conditions is satisfied if one of the dispersion curves intersects with the blue ordashed blue line. The red line represents schematically the corresponding cross-section. Ifthe energy of the incident neutrons is varied the blue and dashed blue lines sweep over thedispersion curves of the phonons and the full dispersion relation in n direction is mappedout.

L14 Electronic excitations in solids

14.1 Hamiltonian of the electrons in adiabatic approximation

We saw in section L2 that the adiabatic principle implies a decoupling of the lattice andelectronic degrees of freedom. To leading order, the dynamics of the electrons is governedby the Hamiltonian

Hel(R) = −1

2

N∑j=1

∂2rαj

+N∑j=1

L∑k=1

−Zk‖rj −Rk‖︸ ︷︷ ︸VI(rj)

+∑

1≤j<k≤N

1

‖rj − rk‖︸ ︷︷ ︸Velel

+M , (14.1)

Page 70: Introduction to Solid State Physics - arXiv

70

where the ions are fixed to their equilibrium positions Rk and

M =∑

1≤j<k≤L

ZjZk‖Rj −Rk‖

(14.2)

is the electro-static energy of the ions, the so-called Madelung energy. The latter plays norole for the dynamics of the electrons. Thus, to leading order in the adiabatic approximation,the electrons in a crystal are described by a repulsive Coulomb gas with interaction Velel

that is filled into the periodic potential VI generated by the ions sitting on their equilibriumpositions.

Because of the mutual Coulomb interaction of the electrons, this is still an interactingmany-body quantum system, and imperturbable optimism is required to believe that it canever be solved exactly or numerically with sufficient accuracy. At least at the current stageof our knowledge drastic further approximations are necessary in order to be able to makeany quantitative prediction.

14.2 Reduction to a single-particle problem

(i) Only the valence electrons (electrons outside closed shells) contribute significantlyto the typical properties of solids, because they are responsible for the chemicalbonding and are distributed over the solid. For this reason we shall interpret VI asthe potential of ions with completely filled shells (Coulomb with reduced chargenumber Zj ; example sodium, alkali metal, one valence electron, Z = 1).

(ii) If the electron-electron interaction Velel could be neglected, the ‘electronic problem’would be reduced to the problem of a single particle in the periodic potential VI.

(iii) Instead of simply neglecting the electron-electron interaction, we shall try to split itinto an ‘effective single-particle contribution’ which modifies the periodic potentialVI and a residual ‘many-body contribution’ which for the understanding of certainquantities can be neglected in many cases, at least in the first instance. Some ideasabout the systematic derivation of an effective single-particle description will beexplained below.

(iv) An intuitive explanation of how the reduction to a single-particle problem is pos-sible is through the notions of ‘screening’ and ‘mean fields’. The full many-bodyHamiltonian Hel is invariant under the action of the Bravais lattice associated withthe equilibrium positions of the ions, [Hel, UR] = 0 ∀ R ∈ B. Hence, followingthe same reasoning as for the derivation of Bloch’s theorem in section 5.1, we seethat the lattice momentum k ∈ BZ is a good quantum number for the many-bodyeigenfunctions and that those transform like

Ψk(r1 + R, . . . , rN + R) = ei〈k,R〉Ψk(r1, . . . , rN ) , (14.3)

R ∈ B, under the action of the Bravais lattice. It follows that

|Ψk(r1 + R, . . . , rN + R)|2 = |Ψk(r1, . . . , rN )|2 (14.4)

for all R ∈ B. The electronic charge density associated with this state,

Page 71: Introduction to Solid State Physics - arXiv

71

D(x) = −e∫

d3Nr |Ψk(r1, . . . , rN )|2N∑j=1

δ(x− rj)

= −eN∫

d3(N−1)r |Ψk(x, r2, . . . , rN )|2 , (14.5)

clearly inherits the invariance under translations by Bravais lattice vectors, since

D(x + R) = −e∫

d3Nr |Ψk(r1, . . . , rN )|2N∑j=1

δ(x− rj + R)

= −e∫

d3Nr |Ψk(r1 + R, . . . , rN + R)|2N∑j=1

δ(x− rj) = D(x) , (14.6)

where we have used (14.4) in the third equation. We imagine this as the chargedensity ‘screening’ the ionic potential VI and combining together with it to aneffective periodic potential which would be felt by an additional electron insertedas a probe into the solid. This ‘screened periodic potential’ provides an intuitivesingle-particle description for the electron gas in a crystal: a particle moving in themean field of all other particles.

(v) Another way of thinking about the existence of an effective single-particle descriptionof the solid is the following. Consider the excitations of the full many-particle system(14.1). If there are excitations with charge quantum numbers ±e which do havea dispersion, i.e. excitations for which a definite change of (lattice) momentum ofthe many-body system always comes with one and the same definite change ofenergy, then we call them quasi-particle excitations or simply particles. It is alwayspossible to define a single-particle Hamiltonian (in momentum representation) thathas exactly the same dispersion relation (the same spectrum) as the full many-particleHamiltonian (14.1). This Hamiltonian may be seen as an effective single-particleHamiltonian of the full system. How well it describes the full system dependson the details. Unlike e.g. in the harmonic crystal which realizes an ideal gas ofphonons, the two- and multi-particle excitations of the full electronic system arenot just superpositions of single-particle excitations. Still, the effective interactionsbetween the quasi-particles may be weak, in which case the effective single-particledescription will give a good description of at least the thermodynamic properties ofthe system.

It is beyond the current capabilities of theoretical physics to prove the existence ofquasi particles for the Hamiltonian (14.1). But many experiments show that solidsgenerically admit quasi-particle electronic excitations with charge quantum numbers∓e. These are called ‘electrons’ and ‘holes’ (the holes are the solid-state analoguesof the positrons). Experiments show in addition that the interaction between severalelectrons or holes or between the electrons and the holes can often be neglected inthe first instance.

Page 72: Introduction to Solid State Physics - arXiv

72

14.3 Particles in a periodic potential

The above discussion should provide enough motivation to study the general problem of aparticle moving in a periodic potential. The corresponding Hamiltonian is

H = −1

2∂2x + V (x) (14.7)

with x ∈ R3 and V (x) = V (x + R) for all R ∈ B. The study of this Hamiltonian willlead us to the extraordinarily successful band model of solids.

Let us start with some general remarks.

(i) The one dimensional Kronig-Penney model with potential

V (x) = V0

∑`∈Z

a δ(x− `a) , (14.8)

where a is the lattice constant and V0 the strength of the interaction, is the onlysimple but non-trivial model of particles in a periodic potential which admits a closedanalytic solution. The next simple case V (x) = −V0 cos(2πx/a) involves Mathieufunctions. Its understanding requires already some mathematical effort.

(ii) There is no other choice than trying to understand the general properties of particlesin a periodic potential by the common means of mathematics or theoretical physics.As far as the mathematical part is concerned it would be instructive to study Floquettheory as part of the theory of ordinary differential equations. For time limitationswe refrain from touching this interesting subject and rather go ahead with typicalmethods of theoretical physics which are based on perturbation theory.

14.4 Exercise 14. Kronig-Penney model

The Kronig-Penney model is a simple one-dimensional model for the understanding of theband structure in solids. Consider an electron of mass m moving in a periodic potential

V (x) = V0

∞∑n=−∞

a δ(x− na) . (14.9)

Here V0 is the strength of the potential and a the lattice constant. Note that either sign ofV0 makes sense. For V0 > 0 the potential is repulsive, for v0 < 0 it is attractive.

(i) Introduce dimensionless units such that the time-independent Schrodinger equationtakes the form

Hϕ(y) =

[− ∂2

∂y2+ 2c

∞∑n=−∞

δ(y − n)

]ϕ(y) = q2ϕ(y) . (14.10)

Which connections exist between x and y, c and V0, and between q2 and the energyE?

(ii) Because of the periodicity of the potential, the Hamiltonian (14.10) commutes withthe shift operator defined by Tϕ(y) = ϕ(y + 1). Thus, H and T have a commonsystem of eigenfunctions, i.e., the eigenvalue equations

Hϕ(y) = q2ϕ(y) , (14.11a)

Page 73: Introduction to Solid State Physics - arXiv

73

Tϕ(y) = eikϕ(y) (14.11b)

can be solved simultaneously. Determine the solutions of (14.11) as a function ofk and q. Which equation connects k with q? For the calculation it is sufficient toconsider the Schrodinger equation with the general solution ϕ(y) = Aeiqy +Be−iqy

in the interval [0,1]. Note that q can take real as well as imaginary values.

(iii) Discuss the dispersion relation cos(k) = cos(q) + (c/q) sin(q) following from (ii)graphically. Observe that k has to be real in order for ϕ(y) to be bounded andnormalizable. Therefore the eigenstates of H and T are restricted on certain energybands.

14.5 Almost free electrons

The qualitative features of the motion of particles can be understood from time-independentperturbation theory. Since V (x) is periodic, it has the Fourier series representation (cf.section 3.5)

V (x) =∑g∈Bg 6=0

Vg ei〈g,x〉 , Vg =1

Vu

∫U

d3x e−i〈g,x〉 V (x) . (14.12)

Here we took the liberty to set V0 = 0 which just fixes the zero point of the energy. In orderto be able to apply perturbation theory we assume that V (x) is a weak periodic potentialin the sense that Vg = λvg, where |vg| is uniformly bounded in λ and |λ| 1.

Since [H,UR] = 0 for all R ∈ B the assumptions of the Bloch theorem (cf. section 5.1)are fulfilled. Hence, the eigenfunctions Ψk of H have definite lattice momentum k ∈ BZand are of the form

Ψk(x) = ei〈k,x〉 uk(x) , (14.13)

where uk(x) = uk(x + R) for all R ∈ B and thus has a Fourier series representation

uk(x) =∑g∈B

uk,g e−i〈g,x〉 , uk,g =1

Vu

∫U

d3x ei〈g,x〉 uk(x) . (14.14)

Inserting this back into (14.13) we see that

Ψk(x) =∑g∈B

uk,g ei〈k−g,x〉 . (14.15)

If we substitute this representation into the Schrodinger equation forH and use (14.12),we obtain∑g′′∈B

(1

2‖k−g′′‖2−ε(k)

)uk,g′′ e

−i〈g′′,x〉+∑

g′,g′′∈Bg′ 6=0

λvg′uk,g′′ ei〈g′−g′′,x〉 = 0 . (14.16)

Here we multiply by ei〈g,x〉 /Vu and integrate over x over the unit cell. Then(1

2‖k− g‖2 − ε(k)

)uk,g + λ

∑g′∈Bg′ 6=g

vg′−guk,g′ = 0 . (14.17)

Page 74: Introduction to Solid State Physics - arXiv

74

−3 −2 −1 1 2 3

2

4

6

8

10

12

ka

ε(k) degeneraciesε0(k)

ε2π/a(k)

ε−2π/a(k)

Figure 9: The dispersion relation of free particles in 1d, conceived as particles in aperiodic potential. The parabolic dispersion relation splits up in branches, obtained by‘back-folding’ into the Brillouin zone. The picture shows the three lowest lying branchesfor the dispersion ε(k) = k2/2m with m = 4. The red dots denotes degenerate points inthe spectrum which belonged to different momenta in the non-periodic picture, but to thesame lattice momentum in the periodic picture.

This is an eigenvalue problem for the vector uk of the Fourier components uk,g, g ∈ B.We will study it perturbatively in λ.

The reference point are free electrons, λ = 0. Then(1

2‖k− g‖2 − ε(k)

)uk,g = 0 ∀ g ∈ B . (14.18)

Clearly, the solutions of this eigenvalue problem are

εh(k) =1

2‖k− h‖2 , uk,g = δg,h (14.19)

for all h ∈ B. They are parameterized by reciprocal lattice vectors.Example: free electrons in 1d. If a > 0 is the lattice constant, then

B =g ∈ R

∣∣∣g = n2π

a, n ∈ Z

, BZ =

[−πa,π

a

]. (14.20)

It follows that

ε0(k) =k2

2, ε± 2π

a(k) =

1

2

(k ∓ 2π

a

)2, . . . , (14.21)

where k ∈ BZ, are the different branches of the dispersion relation. The situation issketched in Figure 9.

For non-zero potential we assume that the eigenvalues and the Fourier coefficients ofthe wave functions can be expanded in an asymptotic series in λ,

εh(k) =1

2‖k− h‖2 + λε

(1)h (k) + λ2ε

(2)h (k) + O

(λ3), (14.22a)

Page 75: Introduction to Solid State Physics - arXiv

75

uk,g = δg,h + λu(1)k,g + O

(λ2). (14.22b)

We choose the normalization of uk such that u(1)k,h = 0. Then (14.22) in (14.17) for g = h

implies−λε(1)

h (k)− λ2ε(2)h (k) + λ

∑g′∈Bg′ 6=h

vg′−hλu(1)k,g′ = O

(λ3). (14.23)

Thus,ε

(1)h (k) = 0 , ε

(2)h (k) =

∑g∈Bg 6=h

vg−hλu(1)k,g , (14.24)

whence, for g 6= h,(1

2‖k− g‖2 − 1

2‖k− h‖2 − λ2ε

(2)h (k)− . . .

)(λu

(1)k,g + . . .

)+

λ∑g′∈Bg′ 6=g

vg′−g(δg′,h + λu

(1)k,g′ + . . .

)= 0 , (14.25)

implying that

u(1)k,g =

2vh−g‖k− h‖2 − ‖k− g‖2 (14.26)

if ‖k− g‖ 6= ‖k−h‖. If we insert this back into (14.23) and use that v−g = v∗g we obtainthe second order corrections to the energies,

εh(k) =1

2‖k− h‖2 + λ2

∑g∈Bg 6=0

2|vg|2‖k− h‖2 − ‖k− h− g‖2 + O

(λ3). (14.27)

As we see, for a weak periodic potential, the energy in second order perturbation theorycan be expressed in terms of the Fourier coefficients of the potential.

If‖k− h‖ = ‖k− h− g‖ (14.28)

for some g ∈ B, g 6= 0 and some k ∈ BZ, then the expression (14.27) is singular at thatspecific value of k. In order to interpret this problem set h = 0. Then (14.28) reduces to

‖k‖ = ‖k− g‖ . (14.29)

If g runs through the nearest-neighbour sites of the origin in B, this equation describes theboundaries of the Brillouin zone. Hence, we expect that the exact solution of the problemexhibits a strong deviation of the dispersion relation from the dispersion relation of freeelectrons at the boundary of the Brillouin zone. For h 6= 0 additional singular manifoldsappear due to back-folding into the Brillouin zone.

For the interpretation of the perturbative result we further recall that the perturbationtheory it applied to each individual energy level, in our case for fixed h and k. Equation(14.27) is valid for all h, k which do not satisfy (14.28). If (14.28) is satisfied for a pairh, k, we have to modify our calculation, applying the scheme of degenerate perturbationtheory instead.

Page 76: Introduction to Solid State Physics - arXiv

76

L15 Particles in a periodic potential

15.1 Degenerate levels

Fix k and assume that there are precisely two vectors h,h′ ∈ B, h 6= h′, such that

‖k− h‖ = ‖k− h′‖ . (15.1)

Then (14.17) has two degenerate zeroth order solutions with energies

εh(k) =1

2‖k− h‖2 =

1

2‖k− h′‖2 = εh′(k) . (15.2)

and Fourier coefficientsuk,g = aδg,h + bδg,h′ (15.3)

where a,b ∈ C are to be determined. The space of solutions is two-dimensional. Weassume again an asymptotic dependence of the dispersion relations on the interactionparameter λ as in (14.22a). The corresponding ansatz for the Fourier coefficients (14.22b)has to be modified due to the degeneracy,

uk,g = aδg,h + bδg,h′ + O(λ) . (15.4)

Substituting this into (14.17) for g = h,h′ and comparing coefficients at order λ we obtain

−ε(1)(k)a+ vh′−hb = O(λ) ,

−ε(1)(k)b+ vh−h′a = O(λ) . (15.5)

The solvability condition for this homogeneous system implies that ε(1)(k) = ±|vh−h′ |.Thus, the two degenerate energy levels split into

ε±(k) =1

2‖k− h‖2 ± λ|vh−h′ |+ O(λ2) . (15.6)

By way of contrast to the non-degenerate case, the corrections to the eigenstates arenow of first order in λ. This means that close to the Brillouin zone boundaries the analyticstructure of the corrections change. As mentioned above, this makes only sense if weconsider a finite system under periodic boundary conditions. Still, we take it as anotherindication that the strongest effect on the dispersion relation of a free particle by a weakperiodic perturbation is at the boundaries of the Brillouin zone.

The possible values of the coefficients a, b follow from

∓ a±|vh−h′ |+ b±vh′−h = 0 ⇔ a±b±

= ± vh′−h|vh−h′ |

= ± e2iδ

⇐(a+

b+

)=

1√2

(eiδ

e−iδ

),

(a−b−

)=

1√2

(eiδ

− e−iδ

). (15.7)

Remark. We have considered a two-fold degeneracy. At special symmetric points at theboundaries of the Brillouin zone in more than one dimension higher degeneracies (likefour or six) may occur.

Page 77: Introduction to Solid State Physics - arXiv

77

−3 −2 −1 1 2 3

2

4

6

ka

ε(k)

Figure 10: The opening of band gaps at degenerate points of the dispersion relationschematically for almost free, periodically perturbed electrons in 1d. The dashed linesrepresent the dispersion of free particles parameterized by the lattice momentum and are thesame as in Figure 9. The blue line represent the deformation of the free-particle dispersionunder the influence of a periodic perturbation.

15.2 Example d = 1

(i) Let h = 0 and h′ = 2π/a. Then a degeneracy occurs at k = π/a, and

ε±(πa

)=

1

2

(πa

)2± λ|v2π/a| . (15.8)

(ii) Let h = 2π/a and h′ = −2π/a. Then the spectrum is degenerate at k = 0,

ε±(0) =1

2

(2π

a

)2± λ|v4π/a| . (15.9)

The two cases are sketched in Figure 10. We observe the ‘opening of band gaps’ at thedegenerate points in the dispersion relation. Band gaps are most characteristic for thephenomenology of crystalline solids.

In order to develop more intuition we discuss the wave functions connected with thefirst case above. These are

Ψ±(x) = a± eiπx/a +b± e−iπx/a +O(λ) =1√2

(ei(πx/a+δ)± e−i(πx/a+δ)

)+ O(λ)

=√

2

cos(πx/a+ δ)

i sin(πx/a+ δ)+ O(λ) . (15.10)

It follows that|Ψ±(x)|2 = 1± cos

(2(πx/a+ δ)

). (15.11)

Page 78: Introduction to Solid State Physics - arXiv

78

The periodic potential V , on the other hand, has the Fourier series representation

V (x) = λ(v−2π/a e−i2πx/a +v2π/a ei2πx/a

)+ higher Fourier modes

= 2λ|v2π/a| cos(2(πx/a+ δ)

)+ higher Fourier modes . (15.12)

Drawing (15.11) and (15.12) in the same picture and choosing λ < 0 (attractive effectivepotential near origin) we see that for Ψ+ the particle is in the average in the valleys of thepotential, whereas it is more on the hills for Ψ−. Accordingly, ε+(π/a) < ε−(π/a) in thiscase.

15.3 The tight-binding method

So far we have studied the formation of energy bands starting from free electrons (planewaves) subject to a periodic perturbation. Now we would like to turn to the oppositeextreme. We shall start with atomic wave functions and ask what happens, when the atomsare brought close to each other. For simplicity we assume a mono-atomic lattice with Nat

atoms.

(i) Let φa(x) an atomic wave function of an electron. In order to satisfy Bloch’s theoremwe consider the linear combination

ϕak(x) =1√Nat

∑R∈B

ei〈k,R〉 φa(x−R) , (15.13)

k ∈ BZ.

(ii) Further define

j(R−R′) =

∫V

d3x φ∗a(x−R′)φa(x−R) , (15.14a)

h(R−R′) =

∫V

d3x φ∗a(x−R′)(Hφa)(x−R) , (15.14b)

where H is the one-particle Hamiltonian (14.7). Since the atomic wave functions de-cay exponentially fast with the distance from the nucleus, we expect these functionsto behave as

j(R−R′) ∼ δR,R′ , (15.15a)

h(R−R′) ∼ δR,R′h(0) , (15.15b)

if the interatomic distance becomes large. It follows that, asymptotically for largeinteratomic distances,

〈ϕak, ϕak′〉 = δk,k′ , (15.16a)

〈ϕak, Hϕak′〉 = h(0)δk,k′ . (15.16b)

Thus, for large interatomic distances, the functions ϕak are a set of approximateeigenfunctions of H corresponding to an Nat-fold degenerate atomic level.

Page 79: Introduction to Solid State Physics - arXiv

79

(iii) This highly degenerate level splits under the influence of the mutual perturbationsof the atoms, if they come closer to each other. In order to take into account theperturbation we determine the norms and energy expectation value in these states,

‖ϕak‖2 =

∫V

d3x |ϕak(x)|2 =1

Nat

∑R,R′∈B

ei〈k,R−R′〉 j(R−R′)

= 1 +∑R∈BR6=0

ei〈k,R〉 j(R) . (15.17)

Note that the sum on the right hand side vanishes for large lattice spacing. For theexpectation value of the energy we obtain in a similar way

E(k) =

∫V

d3xϕ∗ak(x)Hϕak(x)

‖ϕak‖2=

1

‖ϕak‖2Nat

∑R,R′∈B

ei〈k,R−R′〉 h(R−R′)

=1

‖ϕak‖2∑R∈B

ei〈k,R〉 h(R) =1

‖ϕak‖2(h(0) +

∑R∈BR6=0

ei〈k,R〉 h(R)

), (15.18)

where again the sum in the brackets on the right hand side vanishes for large latticespacing.

(iv) Let us now assume that the functions j(R) and h(R) decrease rapidly with increas-ing distance from the origin in B. Then

1 j(Rnn) j(Rnnn) . . . , (15.19a)

|h(0)| h(Rnn) h(Rnnn) . . . , (15.19b)

where ‘nn’ refers to nearest neighbours to the origin, ‘nnn’ to next-to-nearestneighbours etc. It follows that

E(k) = h(0) +∑

R∈Rnn⊂B

ei〈k,R〉(h(R)− h(0)j(R))

+ nnn terms . (15.20)

This formula describes the so-called ‘tight-binding bands’ which give a realisticdescription of bands with ‘pronounced atomic character’ for which the electrons areclose to the atoms. We shall denote

t(R) = h(R)− h(0)j(R) . (15.21)

We conclude from the inversion symmetry of the Bravais lattice that t(R) = t(−R).For this reason tight-binding bands are sums over cosines.

(v) Example fcc lattice: The fcc lattice has 12 nearest neighbours to the origin locatedat

R =a

2

(±1,±1, 0)

(0,±1,±1)

(±1, 0,±1)

, (15.22)

Page 80: Introduction to Solid State Physics - arXiv

80

and t(R) = t by symmetry. Hence,

E(k) = h(0) + t∑

σ,σ′=±

(eia(σkx+σ′ky)/2 + eia(σky+σ′kz)/2 + eia(σkx+σ′kz)/2

)= h(0) + 4t

cos(akx

2

)cos(aky

2

)+ cos

(aky2

)cos(akz

2

)+ cos

(akz2

)cos(akx

2

). (15.23)

Note that the band width is proportional to t. This means that small overlaps of thewave function induces narrow bands. The tight-binding model is thus expected toprovide a good description of narrow energy bands in solids.

(vi) We close this section with two remarks. First, the tight binding bands can be betterjustified by introducing so-called Wannier orbitals instead of atomic orbitals (seebelow).

Second, perhaps the most important insight we can gain from the above reasoningis the intuitive physical picture. Under the influence of the mutual interaction anNat-fold degenerate energy level splits into a band with Nat states. This means thatbands can be classified according to the character of the underlying atomic orbitalsas s, p, d, f bands. Our calculation above makes sense for an isolated s-orbital. Forp, d, f orbitals we cannot start with a single atomic state φa, but have to take intoaccount a set φa(x)amax

a=1 of atomic states. This corresponds to a combination ofthe LCAO (linear combination of atomic orbitals) method of molecular physicswith the tight-binding method. Like in molecular physics hybrid orbitals (like s-dorbitals) may appear in the general case.

L16 Electrons in a periodic potential

16.1 Method of orthogonalized plane waves

The tight-binding method works for low-energy narrow bands formed by atomic states inwhich, in the average, the electrons are close to the ions. A method for the calculation ofmore realistic band structure is obtained by combining the tight binding method with themethod of almost free electrons. This is called the OPW (orthogonalized plane waves)method.

We assume that the low-lying states are known. They may be, for instance, sufficientlywell described by the tight-binding wave function (15.13),

Hϕak ' Etb(k)ϕak , Etb(k) =∑

Rnn∈Bei〈k,R〉 t(R) . (16.1)

LetP =

∑k∈BZ

Pk , (16.2)

the projector onto the subspace of the full Hilbert space that is spanned by the ϕak. Inorder to determine the remaining part of the spectrum it suffices to consider

Page 81: Introduction to Solid State Physics - arXiv

81

H(1− P )Ψ(x) = E(1− P )Ψ(x)

⇔ HΨ(x) + (E −H)PΨ(x) = HΨ(x) +∑

k∈BZ(E − Etb(k))PkΨ(x)

=

[p2

2+ V (x) +

∑k∈BZ

(E − Etb(k))Pk

]Ψ(x) = EΨ(x) . (16.3)

HereW (E,x) = V (x) +

∑k∈BZ

(E − Etb(k))Pk (16.4)

is called a ‘pseudo potential’. The pseudo potential is not a potential, since in positionrepresentation it is represented by an integral operator. Note that

W (E,x)− V (x) > 0 , (16.5)

since E > Etb. We interpret this in such a way that W (E,x) includes the effects ofscreening as discussed above. It is therefore a more appropriate starting point for aperturbation theory for almost free electrons.

16.2 Other methods

Within the ‘augmented plane wave method’ the Schrodinger equation is solved in spheresaround the ions and plane waves are fitted into the space between the spheres (were thepotential is assumed to be negligible).

The KKR (Korringa, Kohn, Rostocker) method is a variant of the augmented planewave method, where, in a first step, the Green function of the Laplace operator is used inorder to transform the Schrodinger equation into an integral equation.

16.3 Summary

(i) For the understanding of many of the electronic properties of solids it suffices to takeinto account the interaction of the electrons only in so far as they screen the attractivepotential of the core ions. The remaining problem is the problem of independentelectrons in a periodic potential.

(ii) Electrons in a periodic potential are characterized by the branches εn(k), n ∈ N, oftheir dispersion relation. As opposed to the spectral problem of phonons the numberof branches for the electrons is infinite. The branches are called energy bands, theirentirety is called the ‘band structure’ of the solid.

(iii) Like for the phonons the εn(k) become differentiable functions of k in the thermo-dynamic limit which exhibit the full translation symmetry of the reciprocal latticeand the full point group symmetry of the solid.

16.4 Exercise 15. Electronic band structure in one-dimensional solids byWKB

The Schrodinger equation for a non-relativistic electron of mass M in a 1d periodicpotential of period L reads

d2ψ

dx2+ [E − V (x)]ψ(x) = 0 . (16.6)

Page 82: Introduction to Solid State Physics - arXiv

82

Here the energy and the length are measured in units of ~2

2ML2 and L, respectively.We would like to solve equation (16.6) by means of the WKB-approximation and find

a condition which determines the band structure, i.e. all allowed energy values E.

B0

A0 F1

G1

A1

B1

a0 b

0a1x x 10

E

Setting p(x) =√E − V (x) we can express the general WKB-solution of the Schrodinger

equation for a single potential barrier V (x) (see figure) in the classically accessible regionsleft and right of the barrier as

E < Vmax :1√p(x)

exp[±i

∫ x

a0

dyp(y)] , x < a0 ,1√p(x)

exp[±i

∫ x

b0

dyp(y)] , x > b0

E > Vmax :1√p(x)

exp[±i

∫ x

0dyp(y)] , x < 0 ,

1√p(x)

exp[±i

∫ x

0dyp(y)] , x > 0 .

(i) The general solution is in each case a linear combination with coefficients A0, B0

or F1, G1 for the left and right classically accessible region, respectively. Thecoefficients are related by a 2 × 2 matrix M . Conclude from the time-inversioninvariance that M1

1 = M22∗, M2

1 = M12∗. For this purpose consider

M

(A0

B0

)=

(F1

G1

), M

(B0

A0

)=

(G1

F 1

). (16.7)

(ii) The current conservation |A0|2 − |B0|2 = |F1|2 − |G1|2 implies detM = 1. LetT be the transmission coefficient and R = 1− T the reflection coefficient with thecorresponding phase shifts eiµ and e−iν . Verify the representation

M =

1√T

eiµ −√

RT ei(µ+ν)

−√

RT e−i(µ+ν) 1√

Te−iµ

. (16.8)

(iii) A periodic continuation of the potential barrier leads to another representation of thesolution in the classically accessible domain b0 < x < a1,

E < Vmax :A1√p(x)

exp[+i

∫ x

a1

dyp(y)] +B1√p(x)

exp[−i

∫ x

a1

dyp(y)] , (16.9)

E > Vmax :A1√p(x)

exp[+i

∫ x

1dyp(y)] +

B1√p(x)

exp[−i

∫ x

1dyp(y)] . (16.10)

Which matrix connects A1, B1 with F1, G1, if we set φ(E) =∫ a1

b0dyp(y), E <

Vmax rsp. φ(E) =∫ 1

0dyp(y), E ≥ Vmax? Calculate the transfer matrix P whichlinks the coefficients A0, B0 and A1, B1 of two neighbouring cells.

Page 83: Introduction to Solid State Physics - arXiv

83

(iv) The full solution for the periodic potential stays bounded as long as P has eigenvaluesof absolute value 1. Show that this fact implies a condition that determines the bandstructure, ∣∣∣∣∣cos[φ(E) + µ]√

T (E)

∣∣∣∣∣ ≤ 1 . (16.11)

(v) For a potential of the form V (x) = V0 cos(2πx), V0 > 0, the Schrodinger equation(16.6) of the non-relativistic particle is equal to the Mathieu equations. The transitioncoefficient T (E) of a single potential barrier in WKB-approximation follows as

T (E) =1

1 + e2W, W (E) =

∫ b0

a0

dy|p(y)| , E < V0 ,

T (E) =1

1 + e−2W, W (E) =

∣∣∣∣ ∫ y2

y1

dyp(y)

∣∣∣∣ , E > V0 ,

where for E > V0 the integral has to be calculated on the direct line betweenboth imaginary reversal points. Represent φ(E) and W (E) by complete ellipticalintegrals of the first and second kind K(m) and E(m), with the dimensionlessparameter m = E+V0

2V0, E < V0 and m = 2V0

E+V0, E > V0, respectively!

(vi) Choosing different parameters m (and thus fixing E/V0) it is possible to to sketchthe regions in the V0,E-diagram for µ ≈ 0, where the solutions are bounded. Findsuch type of diagram in the literature.

16.5 The Fermi distribution

Electrons are Fermions. According to the Pauli principle many-electron wave functionsmust be totally anti-symmetric under the permutations of any two electrons. Consequen-tially, in a system of many non-interacting electrons (or holes) no two of them can be inthe same single-particle state. When we calculate the grand canonical partition functionwe therefore have to count every single-particle state as either unoccupied or occupied byjust one electron (or hole),

Ze =

∞∏n=0

∏k∈BZ

(1 + e−

εn(k)−µT

). (16.12)

Here 1 = e0 stands for an unoccupied state, while e−εn(k)−µ

T stands for an occupiedstate. Every factor on the right hand side of (16.12) represents the partition functioncorresponding to a single-electron state. Hence, the grand-canonical probability for havingthis state occupied is

f(εn(k)− µ) =e−

εn(k)−µT

1 + e−εn(k)−µ

T

=1

eεn(k)−µ

T +1. (16.13)

As we recall from our lecture on statistical mechanics, this is the Fermi distributionfunction.

Page 84: Introduction to Solid State Physics - arXiv

84

16.6 Grand canonical potential of the electron gas

As for every ideal gas of spin-12 Fermions we can immediately write down the total particle

number N , internal energy E and entropy S of the system as sums involving the Fermifunction,

N(T, µ) = 2∞∑n=0

∑k∈BZ

f(εn(k)− µ) , (16.14a)

E(T, µ) = 2∞∑n=0

∑k∈BZ

f(εn(k)− µ)εn(k) , (16.14b)

S(T, µ) = −2∞∑n=0

∑k∈BZ

f(εn(k)− µ) ln

(f(εn(k)− µ)

)+ (1− f(εn(k)− µ)) ln

(1− f(εn(k)− µ)

). (16.14c)

Here the factor of 2 accounts for the spin degree of freedom. The grand canonical potentialis obtained as

Ω(T, µ) = E(T, µ)− TS(T, µ)− µN(T, µ)

= 2T∞∑n=0

∑k∈BZ

f ln(1/f − 1) + f ln f + (1− f) ln(1− f)

= −2T

∞∑n=0

∑k∈BZ

ln(1 + e−

εn(k)−µT

). (16.15)

In a similar way as for the phonon gas we can rewrite the sums over all lattice momentaasymptotically for large volume V first as an integral over the Brillouin zone and then asan integral over all energies. For the second step we need to define an electronic density ofstates. We proceed as for the phonon gas and first of all introduce the density of states fora single branch of the dispersion relation,

gn(ε) =1

(2π)3

∫BZ

d3k δ(ε− εn(k)) =1

(2π)3

∫S(ε)

dS

‖ gradk εn(k)‖ , (16.16)

where S(ε) is the surface implicitly determined by the equation ε = εn(k). With this the(total) density of states is defined as

g(ε) = 2

∞∑n=0

gn(ε) . (16.17)

Again a factor of 2 is included to take the spin degrees of freedom into account.

Remark. Our statements about van-Hove singularities in section 9.4 remain valid forelectrons.

16.7 Fermi energy and Fermi surface

Two important notions in solid state physics are the ‘Fermi energy’ and the ‘Fermi surface’.Using the density of states introduced above we may write the particle number as a functionof temperature and chemical potential as

N(T, µ) = V

∫ ∞−∞

dε g(ε)f(ε− µ) . (16.18)

Page 85: Introduction to Solid State Physics - arXiv

85

Since

∂µN(T, µ) =V

4T

∫ ∞−∞

dεg(ε)

ch2((ε− µ)/2T )> 0 , (16.19)

equation (16.18) can be inverted at any T > 0 to give µ = µ(T,N) (the latter fact follows,of course, also from general arguments on the equivalence of thermodynamic ensembles inthe thermodynamic limit).

The Fermi energy EF is defined as the chemical potential at T = 0,

EF = limT→0+

µ(T,N) . (16.20)

It is clear from (16.18) and (16.19) that the Fermi energy depends only on the particledensity N/V and is a monotonically increasing function of the particle density. In acanonical ensemble description, i.e. if we fix the particle number and consider µ as afunction of N and T , we obtain the pointwise pointwise limit,

limT→0+

f(ε− µ(T,N)

)= Θ(EF − ε) , (16.21)

where Θ is the Heaviside function. This reflects the fact that in the ground state all single-particle states with energies up to EF are occupied, while those with energies larger thanEF are unoccupied.

For the electronic ground state of solids (within the band model) exist two alternatives,

(i) g(EF ) > 0 , (ii) g(EF ) = 0 , (16.22)

which come with drastically different phenomenologies. In case (i) the Fermi energy lieswithin a band. In case (ii) it is situated in a band gap (cf. Figure 11). In case (i) excitationsof arbitrarily small energy are possible. In case (ii), due to the Pauli principle, the smallestpossible excitation energy is equal to the band gap. In case (i) an arbitrarily small electricfield causes a current, in case (ii) this does not happen. This is our first explanation, withina single-particle picture, of the difference between conductors and insulators.

In conductors the Fermi energy is in at least one band, and the equations

εn(k) = EF n ∈ N (16.23)

determine a surface in reciprocal space which is called the Fermi surface SF . The volumeenclosed by the Fermi surface

IntSF =k ∈ R3

∣∣∣εn(k) ≤ EF , n ∈ N

(16.24)

is called the Fermi sphere. In general the Fermi surface has a complicated shape and topol-ogy (not necessarily simply connected). It is of crucial importance for the understandingof the transport properties of solids, in particular in the presence of a magnetic field.

L17 Low-temperature specific heat of the electron gas

A first quantitative example, showing that it makes a big difference whether or not the Fermienergy lies within a band, is the thermodynamics of band electrons at low temperature.

Page 86: Introduction to Solid State Physics - arXiv

86

0.1 0.2 0.3 0.4 0.5

−0.5

0.5

1

1.5

2

EF conductor

EF insulator

g(ε)

ε

Figure 11: Within the single-particle picture or band model a solid is a conductor or aninsulator depending on whether or not the Fermi energy is situated within a band or in aband gap.

Expressing the grand canonical potential (16.15) by means of the electronic density ofstates (16.16), (16.17) we obtain

Ω(T, µ) = −TV∫ ∞−∞

dε g(ε) ln(1 + e−

ε−µT). (17.1)

In order to prepare for the low-T analysis we rewrite this as

Ω(T, µ) = V

∫ µ

−∞dε g(ε)(ε− µ)− TV

∫ ∞−∞

dε g(ε) ln(1 + e−

|ε−µ|T). (17.2)

Since g(x) is the density of states of an electronic band structure, there are two interlacedsequences (an)n∈N, (bn)n∈N with an < bn < an+1 < bn+1 and g(x) 6= 0 ∀ x ∈ (an, bn),g(x) = 0 ∀ x ∈ [bn, an+1]. If µ is in one of the band gaps (bn, an+1), then the secondintegral on the right hand side of (17.2) vanishes exponentially fast for T → 0, which isnot the case, if µ is situated inside a band.

17.1 Specific heat of metals

Both cases require a separate asymptotic analysis. Let us start with the metallic caseµ ∈ (an, bn) for some n ∈ N. We assume that g is analytic in (an, bn). Then it has aTaylor series expansion around ε = µ with some finite radius of convergence δ > 0,

g(ε) =∞∑k=0

g(k)(µ)

k!(ε− µ)k , (17.3)

if |ε− µ| < δ. It follows that

Ω(T, µ)− V∫ µ

−∞dε g(ε)(ε− µ)

Page 87: Introduction to Solid State Physics - arXiv

87

= −TV∞∑k=0

g(k)(µ)

k!

∫ µ+δ

µ−δdε (ε− µ)k ln

(1 + e−

|ε−µ|T)

+ O(T∞)

= −V∞∑k=0

g(k)(µ)

k!T k+2

∫ δ/T

−δ/Tdx xk ln

(1 + e−|x|

)+ O

(T∞). (17.4)

Here we have substituted x = (ε− µ)/T in the second equation. The remaining integralvanishes for symmetry reasons if k is odd. If k = 2m we obtain∫ δ/T

−δ/Tdx x2m ln

(1 + e−|x|

)= 2

∫ ∞0

dx x2m ln(1 + e−x

)+ O

(T∞)

= 2∞∑n=1

(−1)n+1

n

∫ ∞0

dx x2m e−nx +O(T∞)

= 2(2m)!∞∑n=1

(−1)n+1

n2m+2+ O

(T∞)

= (2m)!(2− 2−2m)ζ(2m+ 2) + O(T∞), (17.5)

where ζ is Riemann’s zeta function. Inserting this back into (17.4) we obtain the low-Tasymptotic series

Ω(T, µ) = V

∫ µ

−∞dε g(ε)(ε− µ)

− V∞∑k=0

(2− 2−2k)ζ(2k + 2)g(2k)(µ)T 2k+2 + O(T∞)

(17.6)

for the grand canonical potential.The corresponding series for the particle number and entropy in a grand canonical

description are

N(T, µ) = −∂Ω(T, µ)

∂µ= V

∫ µ

−∞dε g(ε)

+ V∞∑k=0

(2− 2−2k)ζ(2k + 2)g(2k+1)(µ)T 2k+2 + O(T∞), (17.7a)

S(T, µ) = −∂Ω(T, µ)

∂T

= V∞∑k=0

(2− 2−2k)(2k + 2)ζ(2k + 2)g(2k)(µ)T 2k+1 + O(T∞). (17.7b)

The asymptotic series allow us to calculate the low-temperature expansion of thespecific heat order by order, using

CV = T∂S

∂T

∣∣∣∣N

= T

(∂S

∂T

∣∣∣∣µ

+∂S

∂µ

∣∣∣∣T

∂µ

∂T

∣∣∣∣N

). (17.8)

Equation (17.7a) can be used iteratively to obtain µ as a function of N and T . The equation

N

V=

∫ Ef

−∞dε g(ε) , (17.9)

Page 88: Introduction to Solid State Physics - arXiv

88

following from (17.7a) at T = 0, determines EF = µ(0, N) as a function of the density ofparticles N/V . Since only T 2 enters (17.7a), µ must be even in T ,

µ = EF + αT 2 + O(T 4). (17.10)

Inserting this into (17.7a) and using (17.9) we can calculate α,∫ µ

EF

dεg(ε)+ζ(2)g′(µ)T 2 = (µ−EF )g(µ)−g′(µ)(µ− EF )2

2+ζ(2)g′(µ)T 2+O

(T 6)

=(αg(EF ) + ζ(2)g′(EF )

)T 2 + O

(T 4)

= O(T 4)

⇒ α = −ζ(2)g′(EF )

g(EF )⇒ µ = EF − ζ(2)

g′(EF )

g(EF )T 2 + O

(T 4). (17.11)

Furthermore, using (17.7b) we get at once that

T∂S

∂T

∣∣∣∣µ

= 2ζ(2)TV g(EF ) + O(T 3), (17.12a)

T∂S

∂µ

∣∣∣∣T

= 2ζ(2)T 2V g′(EF ) + O(T 4). (17.12b)

Using (17.11) and (17.12a) in (17.8) and recalling that ζ(2) = π2/6 we finally arrive atthe sought for low-temperature asymptotics of the specific heat of a metal,

CV =π2

3TV g(EF ) + O

(T 3). (17.13)

Let us add a few comments in conclusion.

(i) The above low-T asymptotic expansion of the thermodynamic quantities of a Fermigas goes back to Sommerfeld [13] and is called the Sommerfeld expansion.

(ii) Equation (17.13) is an important result stating that the electronic contribution to thespecific heat of metals is linear in T and proportional to the density of states at theFermi energy.

(iii) Taking it the other way round, we see that we can experimentally determine thedensity of states of a metal close to its Fermi energy by measuring its specific heat.

(iv) Since typical electronic energies in solids, like band gaps or band widths, are of theorder of 1 eV, low-temperature expansions for the electrons in solids are usuallyvalid even above room temperature.

17.2 Specific heat of insulators

Starting once more from equation (17.2) we consider the specific heat of an insulator. Bydefinition a band insulator has g(EF ) = 0. The Fermi energy is located inside a bandgap, bn < EF < an+1 for some n ∈ N. In this situation the closest band below the Fermienergy, the one with index n here, is called ‘valence band’, the closest band above theFermi energy, the one with index n+ 1, ‘conduction band’. It follows from (17.2) that, forbn < µ < an+1,

Page 89: Introduction to Solid State Physics - arXiv

89

Ω(T, µ)− V∫ µ

−∞dε g(ε)(ε− µ)

∼ −TV∫ ∞−∞

dε g(ε) e−|ε−µ|T ∼ −TV

∫ bn

an

dε g(ε) eε−µT +

∫ bn+1

an+1

dε g(ε) e−ε−µT

= −TV

ebn−µT

∫ bn

an

dε g(ε) eε−bnT + e−

an+1−µT

∫ bn+1

an+1

dε g(ε) e−ε−an+1

T

= −T 2V

ebn−µT

0∫−(bn−an)/T

dx g(bn + Tx) ex + e−an+1−µ

T

(bn+1−an+1)/T∫0

dx g(an+1 + Tx) e−x. (17.14)

Here we have neglected contributions that are exponentially smaller than the displayedterms. In order to further simplify the remaining integrals we have to recall that for 3dsystems there are square-root type van-Hove singularities at the band edges. This meansthat there are αv, αc > 0 such that within the respective bands

g(bn+Tx) ∼ αv√−Tx(1+O(Tx)) , g(an+1+Tx) ∼ αc

√Tx(1+O(Tx)) . (17.15)

Substituting this into the integrals on the right hand side of (17.14) we obtain, for instance,

0∫−(bn−an)/T

dxg(bn+Tx) ex ∼√Tαv

∫ ∞0

dx√x e−x =

√TαvΓ(3/2) =

1

2

√πTαc (17.16)

and a similar expression for the other integral. Altogether we end up with

Ω(T, µ) ∼ V∫ µ

−∞dε g(ε)(ε− µ)−

√π

2T

52V

αv e−

µ−bnT +αc e−

an+1−µT

, (17.17)

which is the low-temperature asymptotics of the grand canonical potential for an insulator,when µ ∈ (bn, an+1).

Again the formulae for particle number and entropy in the grand canonical ensembleare obtained by taking derivatives (cf. (17.7)),

N(T, µ) ∼ V∫ µ

−∞dε g(ε)−

√π

2T

32V

αv e−

µ−bnT −αc e−

an+1−µT

, (17.18a)

S(T, µ) ∼√π

2T

12V

(µ− bn)αv e−

µ−bnT +(an+1 − µ)αc e−

an+1−µT

. (17.18b)

For T → 0+ in (17.18a) we still get that the integral on the right hand side with upperlimit EF is equal to the total particle number, but now this equality does not fix EF , sincethe integral as a function of µ is constant for µ ∈ [bn, an+1], thus non-invertible. Sinceµ is continuous in a vicinity of T = 0, we see that the integral exactly equals N evenfor small finite temperatures. Hence, the second term on the right hand side must vanishasymptotically, √

π

2T

32V

αv e−

µ−bnT −αc e−

an+1−µT

∼ 0 , (17.19)

Page 90: Introduction to Solid State Physics - arXiv

90

which determines µ at small T to be

µ =bn + an+1

2+T

2ln(αvαc

)+ O

(T∞). (17.20)

In particular,

EF =bn + an+1

2. (17.21)

The Fermi energy of an insulator is in the middle of the band gap.By definition

∆ = an+1 − bn (17.22)

is the ‘width of the band gap’ or simply ‘the band gap’. Using (17.8), (17.18b) and(17.20)-(17.22) one straightforwardly obtains the expression

CV =αv + αc

2

√π

TV(∆

2

)2e−

∆2T(1 + O(T )

)(17.23)

for the low-temperature asymptotic behaviour of the specific heat of a band insulator.Here again a few concluding remarks are in order.

(i) The functional form of the low-T specific heat in (17.23) is called ‘thermally ac-tivated behaviour’. One says that the activation barrier equals half of the bandgap.

(ii) The size of the electronic contribution to the specific heat of an insulator depends inan extremal way on the band gap. Example:

∆ = 0,5 eV ⇒ e−∆2T ≈ 10−4 , (17.24a)

∆ = 3,0 eV ⇒ e−∆2T ≈ 10−24 (17.24b)

at T ≈ 300K. This makes the difference, as far as the electronic specific heat isconcerned, between insulators and ‘semi-conductors’.

(iii) For an insulator in the low-temperature regime the electrons contribute almostnothing to the specific heat. The specific heat of insulators is mostly determinedby the phonons. In conductors, on the other hand, the electrons do contribute tothe specific heat and even dominate it at low enough temperatures. For this reasonmetals have, in general, a larger heat capacity than insulators.

L18 Electrons in solids – the second quantized picture

We recall that the Hamiltonian (14.1) of the electrons in solids in adiabatic approximationis

Hel =

N∑j=1

1

2‖pj‖2 + VI(xj)

+

∑1≤j<k≤N

VC(xj − xk) , (18.1)

where VI(x) is the periodic potential of the ions in the crystal, VC(x) = 1/‖x‖ is theCoulomb potential, and N is the number of electrons we are taking into account.

Page 91: Introduction to Solid State Physics - arXiv

91

18.1 An auxiliary potential

Our aim in the following lectures is to proceed beyond the single-particle approximation.To begin with, we remark that we can introduce an auxiliary potential VA(x) withoutchanging too much the structure of the Hamiltonian. Let

V (x) = VI(x) + VA(x) . (18.2)

Then

Hel −N∑j=1

1

2‖pj‖2 + V (xj)

=

∑1≤j<k≤N

VC(xj − xk)−N∑j=1

VA(xj)

=∑

1≤j<k≤N

VC(xj − xk)−

VA(xj) + VA(xk)

N − 1︸ ︷︷ ︸=U(xj ,xk)

⇔ Hel =N∑j=1

1

2‖pj‖2 + V (xj)

+

∑1≤j<k≤N

U(xj ,xk) . (18.3)

Note that the choice of the auxiliary potential is entirely at our disposal. We may, forinstance, choose the mean field potential defined in section 14.2. If we neglect U weare applying a single-particle approximation. Good single-particle approximations areobtained for appropriate choices of VA. A single-particle approximation is good, if thetwo-particle matrix elements of U(x,y) calculated with eigenstates of the single-particleHamiltonian

h(x,p) =‖p‖2

2+ V (x) (18.4)

are small for single-particle energies close to the Fermi surface.

18.2 Bloch basis and Wannier basis

In the following calculation we will not fix the auxiliary potential VA. Our only explicitassumption is that it is periodic. Implicitly we shall also assume that it allows us to takescreening into account. Due to the periodicity the eigenfunctions of h are ‘Bloch functions’ϕαk labeled by a band index α ∈ N and a lattice momentum k ∈ BZ,

hϕαk(x) = εα(k)ϕαk(x) . (18.5)

We shall call the εα(k) the single-particle energies. The Bloch theorem implies

ϕαk(x) = ei〈k,x〉 uαk(x) (18.6)

with a lattice periodic function uαk. The set ϕαkα∈N,k∈BZ is a single-particle orthonor-mal basis of the electronic Hilbert space called the ‘Bloch basis’.

Define

φα(x) =1√L

∑k∈BZ

ϕαk(x) , (18.7)

Page 92: Introduction to Solid State Physics - arXiv

92

where L is the number of unit cells.* Then φα(x − Rj)|α ∈ N,Rj ∈ B is anotherorthonormal basis (prove it!) called the Wannier basis. φα(x) is called a Wannier function.The Wannier functions generalize the atomic orbitals in section 15.3.

Bloch basis and Wannier basis are connected via Fourier transformation,

1√L

L∑j=1

ei〈k,Rj〉 φα(x−Rj) =1

L

L∑j=1

∑p∈BZ

ei〈k,Rj〉 ei〈p,x−Rj〉 uαp(x−Rj)

=∑

p∈BZei〈p,x〉 uαp(x)

1

L

L∑j=1

ei〈k−p,Rj〉 = ϕαk(x) . (18.8)

Here we have used that the second sum on the right hand side of the second equation equalsLδk,p.

18.3 Hamiltonian in second quantization

Let c+αk,a the creation operator of a Bloch electron of spin a ∈ ↑, ↓. The operators

c+αj,a =

1√L

∑k∈BZ

e−i〈k,Rj〉 c+αk,a (18.9)

are then an alternative set of creation operators, creating electrons in Wannier orbitals. Thiscan be seen by writing down the corresponding field operators,

Ψ+a (x) =

∑αk

ϕ∗αk(x)c+αk,a =

∑αk

ϕ∗αk(x)1√L

L∑j=1

ei〈k,Rj〉 c+αj,a

=

L∑j=1

(1√L

∑αk

ϕ∗αk(x) ei〈k,Rj〉)c+αj,a =

∑α

L∑j=1

φ∗α(x−Rj)c+αj,a (18.10)

According to the general prescription (see lecture on QM) the Hamiltonian in ‘occupationnumber representation’ (‘second quantization’) can be written as

Hel =

∫d3xΨ+

a (x)hΨa(x) +1

2

∫d3x

∫d3y Ψ+

a (x)Ψ+b (y)U(x,y)Ψb(y)Ψa(x)

=∑α,β;i,j

∫d3x φ∗α(x−Ri)hφβ(x−Rj)︸ ︷︷ ︸

=δαβtαij

c+αi,acβj,a

+1

2

∑α,β,γ,δi,j,k,`

∫d3x

∫d3y φ∗α(x−Ri)φ

∗β(y −Rj)U(x,y)φγ(y −Rk)φδ(x−R`)︸ ︷︷ ︸

=Uαβγδijk`

×

× c+αi,ac

+βj,bcγk,bcδ`,a

=∑α;i,j

tαijc+αi,acαj,a +

1

2

∑α,β,γ,δi,j,k,`

Uαβγδijk` c+αi,ac

+βj,bcγk,bcδ`,a . (18.11)

*Note the change of notation!

Page 93: Introduction to Solid State Physics - arXiv

93

Here implicit summation over the spin indices a, b is implied. The tαij are called ‘tran-

sition matrix elements’ or ‘hopping matrix elements’, the Uαβγδijk` are called ‘interactionparameters’.

Note that:

(i) So far the Hamiltonian is only rewritten in ‘second quantization’. No other approxi-mation than the adiabatic approximation has been applied.

(ii) In this form it is the starting point of the ‘theory of strongly correlated electronsystems’.

(iii) An optimal choice of the Wannier functions (through an optimal choice of VA)minimizes the strength and the range of the interaction parameters.

(iv) Suppose the auxiliary potential VA can be chosen in such a way that the interactionparameters are always small. Then they can be neglected and we are in the realm ofband theory.

(v) If the Wannier functions can be constructed in such a way that they resemble atomicwave function in the sense that they are localized around the origin and decaysufficiently fast away from it, then the interaction parameters become short-range,and it may be justified to consider only on-site or near-neighbour contributions.

18.4 Exercise 16. Wannier functions in one dimension

The dispersion relation of a free non-relativistic particle of mass m is ε(k) = ~2k2/2m.Consider a lattice of N sites and physical length L under periodic boundary conditions.Then only N discrete wave vectors of the form k = 2πn/L, n ∈ Z, are possible inside theBrillouin zone. All wave vectors outside will be folded back to the Brillouin zone, i.e. newbands with dispersion ε(k + 2πm/a) develop, where a = L/N is the lattice constant andm ∈ Z is the band index. The corresponding Bloch functions are labeled in the followingway,

ϕm,k(x) =1√L

exp

[i(k +m

a)x

]. (18.12)

Construct the complementary Wannier basis φm(x−Ri), i = 1, . . . , N , in the thermo-dynamic limit (N,L→∞, L/N = a = const.). This requires to calculate

φm(x) =1√N

∑k∈BZ

ϕm,k(x) , (18.13)

where, in the thermodynamic limit, the summation over the first Brillouin zone can bereplaced by an integration between the zone boundaries ±π/a.

18.5 Exercise 17. Spinless Fermions on the lattice

Consider creation and annihilation operators c+m, cn,m,n = 1, . . . , L, of spinless Fermions.

They satisfy canonical anti-commutation relations

cm, cn = c+m, c

+n = 0 , cm, c+

n = δm,n . (18.14)

Page 94: Introduction to Solid State Physics - arXiv

94

Let |0〉 denote the Fock vacuum defined by cm|0〉 = 0, m = 1, . . . , L. Non-interactingFermions are described by the Hamiltonian

H =

L∑m,n=1

tmn c+mcn . (18.15)

Here the matrix t with matrix elements tmn is called the transition matrix. It is Hermitian,tmn∗ = tnm, by definition.

(i) Show that the canonical anti-commutation relations (18.14) are invariant undertransformations of the form

ck =L∑

m=1

Ukmcm , (18.16)

if the matrix U with matrix elements Ukm is unitary.

(ii) Show that a unitary transformation U exists which transforms the Hamiltonian(18.15) to the form

H =

L∑m=1

εmc+mcm . (18.17)

Explain why this transformation solves the eigenvalue problem H|ψ〉 = E|ψ〉.

(iii) We say that the transformation in (ii) diagonalizes H . Diagonalize the Hamiltonianwith transition matrix elements

tmn = −t0(δm,n+1 + δm,n−1) (18.18)

explicitly. Here the indices of the right hand side should be understood modulo L.

(iv) Also diagonalize the Hamiltonian with transition matrix elements

tmn =

0 if m = nπiL sin−1

((m−n)π

L

)else.

(18.19)

Hint: Use the canonical gauge transformation cn 7→ eiπn/2Lcn as well as the relation1/(e−iπ(m−n)/L − 1) = 1

L

∑L−1k=0 ke

−iπ(m−n)k/L. The latter can be proved bydifferentiating a geometric sum with respect to an appropriate parameter.

L19 The Hubbard model

19.1 Motivation and definition

The formalism of second quantization, in which the Hamiltonian of the electrons takesthe form (18.11), gives us a more intuitive access to the problem. If the Fermi surface islocated within a single conduction band with band index α, say, then the interaction betweendifferent bands can be neglected for small excitation energies, and we can suppress the bandindices (Greek indices) in (18.11). If, moreover, the ‘intra-atomic Coulomb interaction’

Page 95: Introduction to Solid State Physics - arXiv

95

Uiiii is dominant, then a single effective interaction parameter U , say, remains, and Hel

can be approximated by

H =∑i,j

tijc+iacja +

U

2

∑i

c+iac

+ibcibcia (19.1)

which defines the so-called (one-band) Hubbard model [7, 8].

(i) The Hubbard model is a ‘minimal extension’ of the band model in the sense thatas few as possible of the interaction parameters of the full Hamiltonian (18.11) aretaken into account.

(ii) In spite of its apparent simplicity the Hubbard model is a true many-body model. Ingeneral it is very hard to deal with by any of the means of modern theoretical physicsas e.g. perturbation theory, renormalization group analysis, quantum Monte-Carlomethods.

(iii) As simple it is intuitively, the Hubbard model is the starting point for extensions.Basically all models of ‘strongly correlated electrons’ (everything beyond the bandmodel) are extended Hubbard models which are either obtained by taking interactionparameters of wider range into account (e.g. nearest neighbours, next-to-nearestneighbours) or by taking into account a larger number of bands (e.g. two-bandHubbard model, three-band Hubbard model).

(iv) As it stands the Hubbard model is believed to give a realistic description of

• the electronic properties of solids with tight bands• band magnetism (iron, copper, nickel)• the interaction-induced metal-insulator transition (Mott transition)

(v) If the Hubbard Hamiltonian is supplied with periodic boundary conditions, thenumber of lattice sites (which is equal to the number of Wannier orbitals) is finite.As we are are also dealing with a finite number of states per site, the model hasa finite-dimensional space of states, and can be thought of as a ‘fully regularized’quantum field theory. Its Hamiltonian can be represented by a finite Hermitianmatrix. This makes the Hubbard model attractive for computer based approaches.

(vi) The 1d Hubbard model has the amazing feature of being integrable [10, 12]. Rathermuch is known about its elementary excitations and its thermodynamics [6].

19.2 Tight-binding approximation

The assumption that the Wannier functions are strongly localized in the vicinity of the‘lattice sites’ Rj is compatible with the restriction of the hopping amplitudes tij to nearestneighbours 〈ij〉 on the lattice, which is called the ‘tight-binding approximation’. If weintroduce the ‘density operators’ (local particle-number operators)

ni↑ = c+i↑ci↑ , ni↓ = c+

i↓ci↓ (19.2)

and apply the tight-binding approximation to (19.1) we obtain

H = −t∑〈ij〉

c+iacja + U

∑i

ni↑ni↓ . (19.3)

Page 96: Introduction to Solid State Physics - arXiv

96

Here we have assumed isotropic nearest-neighbour hopping of strength −t and the vanish-ing of the on-site energies tii, which for a homogeneous model can always be assumed,since in this case it is equivalent to a redefinition of the chemical potential. For theinteraction part we have calculated

c+iac

+ibcibcia = c+

i↑c+i↓ci↓ci↑ + c+

i↓c+i↑ci↑ci↓ = 2ni↑ni↓ . (19.4)

With the Hamiltonians (19.1) and (19.3) the terminology is not so sharp. The Hamil-tonian (19.3) is also called ‘the Hubbard Hamiltonian’. And, in fact, the term Hubbardmodel most commonly refers to the model described by (19.3).

19.3 Interpretation of the Hubbard Hamiltonian

Let us have a closer look at the Hubbard Hamiltonian. For simplicity we consider theone-dimensional model,

H = −tL∑j=1

(c+jacj+1a + c+

j+1acja)

+ UL∑j=1

nj↑nj↓ , (19.5)

subject to periodic boundary conditions, cL+1a = c1a. Its space of states will be denotedH(L). It is generated by filling electrons into Wannier states. According to the Pauliprinciple, every Wannier state may be unoccupied, occupied with one electron of spin ↑ or↓, or with two electrons with opposite spins, giving altogether dimH(L) = 4L states.

Let us construct the basis of Wannier states explicitly. For this purpose define the rowvectors x = (x1, . . . , xN ), a = (a1, . . . , aN ), where xj ∈ 1, . . . , L, ak ∈ ↑, ↓ forj, k = 1, . . . N ; N ∈ 1, . . . , 2L. The state

|x;a〉 = c+xN ,aN

. . . c+x1,a1|0〉 (19.6)

is a Wannier state representing N electrons at sites xj with spins aj . The set of all differentstates of this form,

BW =

|x;a〉 ∈ H(L)

∣∣∣∣ N = 0, . . . , 2Lxj+1 ≥ xj , aj+1 > aj if xj+1 = xj

, (19.7)

is a basis of Wannier states, since all such states are linear independent and since theirnumber is

2L∑N=0

(2L

N

)= 4L . (19.8)

The operators nj,↑, nj,↓ are the local particle number operators for electrons of spin↑ and ↓ at site j. Let us recall why this name is justified. Using the canonical anti-commutation relations of the Fermi operators and the fact that the cj,a annihilate the Fockvacuum |0〉 we conclude that

[nj,↑, c+k,b] = δjkδ↑bc

+k,b , nj,↑|0〉 = 0 , (19.9)

and therefore

nj,↑|x,a〉 =N∑k=1

δj,xkδ↑,ak |x,a〉 (19.10)

Page 97: Introduction to Solid State Physics - arXiv

97

and similarly for nj,↓. Thus, nj,a|x,a〉 = |x,a〉, if site j is occupied by an electron of spina, and zero elsewise.

A first interpretation of the Hubbard model can be obtained by considering separatelythe two contributions that make up the Hamiltonian (19.5). For t = 0 or U = 0 it can bediagonalized and understood by elementary means. For t = 0 the Hamiltonian reduces toH = UD, where

D =L∑j=1

nj↑nj↓ . (19.11)

Using (19.10) we can calculate the action of D on a state |x,a〉,

D|x,a〉 =

N∑k,l=1

δxk,xlδ↑,akδ↓,al |x,a〉

=∑

1≤k<l≤Nδxk,xl(δ↑,akδ↓,al + δ↓,akδ↑,al)|x,a〉

=∑

1≤k<l≤Nδxk,xl(δ↑,ak + δ↓,ak)(δ↑,al + δ↓,al)|x,a〉

=∑

1≤k<l≤Nδxk,xl |x,a〉 .

(19.12)

Here we used δ↑,akδ↓,ak = 0 in the second equation and the Pauli principle in the thirdequation. As we learn from (19.12) every state |x,a〉 is an eigenstate of the operator D.Thus, D is diagonal in the Wannier basis. The limit t → 0 of the Hubbard Hamiltonian(19.5) is called the atomic limit, because the eigenstate |x,a〉 describes electrons localizedat the sites x1, . . . , xN , which are identified with the loci of the atomic orbitals the electronsmay occupy.

The meaning of the operator D is evident from equation (19.12). D counts the numberof double-occupied sites in the state |x,a〉. The contribution of the term UD to the energyis non-negative for positive U and increases with the number of double-occupied sites.This can be viewed as on-site repulsion among the electrons. Negative U on the other hand,means on-site attraction. Hence, it is natural to refer to D as to the operator of the on-siteinteraction.

In the other extreme, when U = 0, the Hamiltonian (19.5) turns into

H0 = −tL∑j=1

(c+j,acj+1,a + c+

j+1,acj,a) . (19.13)

This is called the tight-binding Hamiltonian. Like every translation invariant one-bodyHamiltonian it can be diagonalized by discrete Fourier transformation. Let us define

c+k,a =

1√L

L∑j=1

eiφkj c+j,a , k = 0, . . . , L− 1, (19.14)

where φ = 2π/L. Then, by Fourier inversion

c+j,a =

1√L

L−1∑k=0

e−iφjk c+k,a , j = 1, . . . , L . (19.15)

Page 98: Introduction to Solid State Physics - arXiv

98

Equation (19.15) is readily verified by inserting (19.14) into the right hand side and usingthe geometric sum formula. Clearly, c+

k+L,a = c+k,a. Insertion of (19.15) into (19.13) leads

to

H0 = −2tL−1∑k=0

∑a=↑,↓

cos(φk)nk,a , (19.16)

where nk,a = c+k,ack,a.

The Fourier transformation leaves the canonical anti-commutation relations invariant,

cj,a, ck,b = c+j,a, c

+k,b = 0 , (19.17a)

cj,a, c+k,b = δjkδab . (19.17b)

A transformation with this property is called canonical. Applying (19.14) to the emptylattice state |0〉 (the Fock vacuum), we obtain

ck,a|0〉 = 0 , k = 0, . . . , L− 1 , a =↑, ↓ . (19.18)

Thus, acting with the creation operators c+k,a on the empty lattice |0〉we obtain an alternative

basis BB . Let us introduce the row vectors q = (q1, . . . , qN ) = (k1, . . . , kN )φ and thestates

|q,a〉 = c+kN ,aN

. . . c+k1,a1|0〉 . (19.19)

It can be shown that these states are eigenstates of a lattice momentum operator witheigenvalue

(∑Nj=1 qj

)mod 2π. The set

BB =

|q,a〉 ∈ H(L)

∣∣∣∣ N = 0, . . . , 2Lqj+1 ≥ qj , aj+1 > aj if qj+1 = qj

(19.20)

is a basis of H(L). This basis is sometimes called the Bloch basis. Electrons in Bloch states|q,a〉 are delocalized, but have definite momenta q1, . . . , qN .

By virtue of (19.17), the analogues of (19.9) and (19.10) are satisfied by nj,a and c+k,b.

It follows that

H0|q,a〉 = −2tN∑j=1

cos(qj)|q,a〉 . (19.21)

Thus, the tight-binding Hamiltonian H0 is diagonal in the Bloch basis. It describes non-interacting band electrons in a cosine-shaped band of width 4t.

The tight-binding Hamiltonian H0 and the operator D which counts the number ofdouble-occupied sites do not commute. Therefore the Hubbard Hamiltonian can neither bediagonal in the Bloch basis nor in the Wannier basis. The physics of the Hubbard modelmay be understood as arising from the competition between the two contributions, H0 andD, to the Hamiltonian (19.5). The tight-binding contribution H0 prefers to delocalize theelectrons, while the on-site interaction D favours localization. The ratio

u =U

4t(19.22)

is a measure for the relative contribution of both terms and is the intrinsic, dimensionlesscoupling constant of the Hubbard model.

Page 99: Introduction to Solid State Physics - arXiv

99

19.4 Exercise 18. Peierls phases

If interacting electrons (charge −e, mass m) in a periodic potential V (x) are exposed toan external electro-magnetic field, their one-particle Hamiltonian becomes

h =1

2m

∥∥∥p +e

cA(x, t)

∥∥∥2+ V (x)− eφ(x, t) , (19.23)

where A(x, t) and φ(x, t) are the vector and scalar potentials of the external field. In thelecture we considered the many-body Hamiltonian generated by h relative to the Wannierbasis. It was parameterized by hopping matrix elements tjk. Here we would like to findout how the external field modifies the tjk.

We start by fixing the gauge such that φ(x, t) = 0.

(i) Let λ(x, t) an arbitrary differentiable function. Using that pk = −i∂k verify thecommutator relation[

pk, e−ieλ(x,t)/c]= −e

c

∂λ(x, t)

∂xke−ieλ(x,t)/c . (19.24)

(ii) Denote by φ(x − Rj) the Wannier orbital at site Rj and recall that the hoppingmatrix elements for vanishing external fields were

tij =

∫d3x φ∗(x−Ri)

[‖p‖22m

+ V (x)

]φ(x−Rj) . (19.25)

Show that in the presence of the external field this has to be modified to become

tij =

∫d3x φ∗(x−Ri) e−ieλ/c

[1

2m

pk +

e

c

(Ak − ∂λ

∂xk

)2

+ V (x)

]eieλ/c

× φ(x−Rj) , (19.26)

where λ is still arbitrary.

(iii) Choosing now

λ(x, t) =

∫ x

x0

dykAk(y, t) (19.27)

for an arbitrary fixed point x0 and redefining φ(x−Rj) = eieλ(x,t)/c φ(x−Rj) weobtain

tij =

∫d3x φ∗(x−Ri)

[‖p‖22m

+ V (x)

]φ(x−Rj) . (19.28)

Discuss under which conditions the approximation φ(x−Rj) ≈ eieλ(Rj ,t)/c φ(x−Rj), the so-called Peierls substitution, should be valid.

(iv) Show that the Peierls substitution leads to a modification of the hopping part of themany body Hamiltonian according to the rule tij → tij eiλij , where

λij =e

c

∫ Rj

Ri

dykAk(y, t) . (19.29)

Page 100: Introduction to Solid State Physics - arXiv

100

L20 The strong-coupling limit

In this lecture we shall consider the limiting case, when the intra-atomic Coulomb in-teraction U of the Hubbard model is large compared to the band width t. We define

T =

L∑j,k=1

tjkc+j,ack,a . (20.1)

Here we only assume that tjk = t∗kj , guaranteeing Hermiticity of T , and that tjj = 0. Forfixed particle number the latter setting merely shifts the energy scale and, for this reason,does not imply any restriction to generality. As we shall see, however, this assumption willhave several technical advantages in the calculations below.

Using an appropriate definition of the tjk and an appropriate enumeration of the latticesites in (20.1), we may write the general Hubbard Hamiltonian (19.1) on any finite lattice,under any kind of boundary conditions and in any dimension in the form

H = T + UD . (20.2)

We shall assume that U > 0. This is natural, since positive U corresponds to therepulsion of electrons in the same Wannier orbital which is expected as a consequence oftheir mutual Coulomb repulsion. If |tjk| U we can consider T as a small perturbationof UD. As we have seen in (19.12), D counts the number of double-occupied sites. Thus,the eigenvalues of UD are 0, U, 2U, . . . , LU . Their number grows linearly with L, whilethe number of states in H(L) grows exponentially like 4L. The eigenvalues of UD aretherefore highly degenerate. Let us denote the projection operators onto the correspondingeigenspaces Hn by Pn, n = 0, 1, . . . , L. Then

H(L) = H0 ⊕H1 ⊕ · · · ⊕HL , (20.3)

and D has the spectral decomposition

D =L∑n=0

nPn . (20.4)

If the particle number N ≤ L, then the ground states of UD are in H0 which hasdimH0 = 3L.

20.1 Degenerate perturbation theory – a reminder

Consider a Hamiltonian H + λV on a Hilbert space H. Assume that H has the spectraldecomposition

H =∑n

EnPn (20.5)

with mutually distinct eigenvalues En and orthogonal projectors Pn (i.e. PnPm = δnmPn),such that dimPnH is not necessarily equal to one. We consider λV , λ ∈ R, as a smallperturbation of ‘strength λ’.

DefineRn =

∑m,m 6=n

PmEn − Em

. (20.6)

Page 101: Introduction to Solid State Physics - arXiv

101

We assume that the eigenvalues and eigenvectors of the perturbed Hamiltonian H+λV arecharacterized by the quantum numbers n of the unperturbed problem and sets of additionalquantum numbers νn ∈ 1, . . . ,dimPnH. In other words

(H + λV )|Ψn,νn〉 = εn,νn |Ψn,νn〉 (20.7)

in such a way that

limλ→0|Ψn,νn〉 = |Ψ(0)

n,νn〉 ∈ PnH , limλ→0

εn,νn = En . (20.8)

We rewrite (20.7) in the form

(En −H)|Ψn,νn〉 =(λV − (εn,νn − En︸ ︷︷ ︸

=∆n,νn

))|Ψn,νn〉 . (20.9)

Then

Rn(En −H)|Ψn,νn〉 = (1− Pn)|Ψn,νn〉 = Rn(λV −∆n,νn)|Ψn,νn〉 ,

⇔ |Ψn,νn〉 = Pn|Ψn,νn〉︸ ︷︷ ︸=|ϕn,νn 〉

+Rn(λV −∆n,νn)|Ψn,νn〉 ,

⇔ |Ψn,νn〉 =

∞∑k=0

[Rn(λV −∆n,νn)

]k|ϕn,νn〉 . (20.10)

Here the series converges if H is finite dimensional and |λ| is small enough. Otherwise wemay have to interpret it as an asymptotic series. Applying, on the other hand, Pn to (20.9)we obtain

PnV |Ψn,νn〉 =∆n,νn

λ|ϕn,νn〉 . (20.11)

Thus,∞∑k=0

PnV[Rn(λV −∆n,νn)

]k|ϕn,νn〉 =∆n,νn

λ|ϕn,νn〉 . (20.12)

This is a non-linear spectral problem on PnH describing the splitting of the energylevel En under the influence of the perturbation λV . The operator on the left hand side isan effective Hamiltonian on PnH. Up to second order (∆n,νn = λ∆

(1)n,νn +λ2∆

(2)n,νn + . . . )

it is given by

H2 = PnV Pn + PnV Rn(λV −∆n,νn)Pn = PnV Pn + λPnV RnV Pn

= PnV Pn + λ∑

m,m 6=n

PnV PmV PnEn − Em

. (20.13)

Thus, up to second order in λ equation (20.12) reduces to a linear spectral problem with aneffective Hamiltonian H2.

Remark. The corresponding eigenstates of the perturbed problem are obtained from(20.10), once the |ϕn,νn〉 are known.

Page 102: Introduction to Solid State Physics - arXiv

102

20.2 Application to the Hubbard model at strong coupling

We now apply (20.13) to (20.2). For this purpose we divide by U . Then H/U = D+T/U .Recall that D has eigenvalues 0, 1, . . . , L. Inserting this for n = 0 with V = T , λ = 1/Uinto (20.13) we obtain

H2 = P0TP0 −1

U

L∑m=1

P0TPmTP0

m. (20.14)

20.3 Explicit form of the projection operators

In order to express the projection operators Pm in terms of Fermi operators, we introducethe function

G(α) =

L∏j=1

(1− αnj↑nj↓) . (20.15)

Its action on the Wannier basis is (see (19.10))

G(α)|x;a〉 = (1− α)n|x;a〉 , (20.16)

where n is the number of double-occupied sites in |x;a〉. In particular,

G(1) =

L∏j=1

(1− nj↑nj↓) = P0 . (20.17)

Moreover,

(−1)k

k!∂kαG(α)

∣∣∣α=1|x;a〉 = |x;a〉

(nk

)(1− α)n−k

∣∣α=1

k ≤ n0 k > n

= δn,k|x;a〉 . (20.18)

It follows that

Pn =(−1)n

n!∂nαG(α)

∣∣∣α=1

, (20.19)

meaning thatG(α) is a generating function for the projection operators Pn, n = 0, 1, . . . , L.

20.4 Application toH2

We will now use the explicit construction of the projection operators to express H2 in termsof Fermions. First of all

P0nj↑nj↓ =

[ L∏k=1

(1− nk↑nk↓)]nj↑nj↓ = 0 = nj↑nj↓P0 , (20.20)

entailing that

G(α)TP0 =L∑

j,k=1

tjk

[ L∏`=1

(1− αn`↑n`↓)]c+j,ack,aP0

=

L∑j,k=1

tjk(1− αnj↑nj↓)c+j,ack,aP0 , (20.21)

Page 103: Introduction to Solid State Physics - arXiv

103

and further, using (20.18),

L∑m=1

PmTP0

m=

L∑j,k=1

tjknj↑nj↓c+j,ack,aP0 . (20.22)

Using the latter equation in (20.14) we arrive at

H2 = P0

L∑j,k=1

tjkc+j,ack,a −

1

U

L∑j,k,k′,`=1

tjktk′`c+j,ack,ank′↑nk′↓c

+k′,bc`,b

P0

= P0

L∑j,k=1

tjkc+j,ack,a −

1

U

L∑j,k,`=1

tjktk`c+j,ack,ank↑nk↓c

+k,bc`,b

P0 . (20.23)

Here we have used (20.20) and the fact that tjj = 0 in the second equation.H2 is an effective Hamiltonian describing the splitting of the lowest level (n = 0) of D

under the influence of the perturbation by T . SometimesH2 is called ‘the t-J Hamiltonian’.Using the canonical anti-commutation relations of the Fermi operators equation (20.23)can be further simplified. After a straightforward but slightly cumbersome calculation weobtain

H2 =L∑

j,k=1j 6=k

tjkc

+j,ack,a(1− nj) +

2|tjk|2U

(Sαj S

αk −

njnk4

)

− 1

U

L∑j,k,`=1j 6=k 6=`6=j

tjktk` nkc+j,ack,ac

+k,bc`,b(1− nj) . (20.24)

Note that the Hamiltonian leaves the space H0 invariant by construction. In (20.24) wehave introduced the notation

nj = nj↑ + nj↓ , (20.25a)

Sαj =1

2

∑a,b=↑,↓

(σα)abc

+j,acj,b , α = x, y, z , (20.25b)

for the particle density and spin density operators. The matrices σα are the well-knownPauli matrices.

20.5 Exercise 19. Strong coupling limit of the Hubbard model

Obtain (20.24) from (20.23)!

L21 Heisenberg model and Mott transition

21.1 Heisenberg Hamiltonian in the language of Fermi operators

In the previous lecture we have derived the strong coupling Hamiltonian H2 by consideringthe ‘hopping’ T as small perturbation to the atomic limit UD of the Hubbard Hamiltonian

Page 104: Introduction to Solid State Physics - arXiv

104

on the subspace H0 ⊂ H(L) with no double-occupied sites. The particle number operatoron H(L) is

N =

L∑j=1

nj . (21.1)

Since [N , Pn] = [N , T ] = 0 we see from (20.14) that H2 preserves the particle number,[N ,H2] = 0. This implies that H2 leaves the subspaces H0,N ⊂ H0, N = 0, . . . , L, offixed particle numbers N invariant. The case N = L is called the case of ‘half-filling’.The corresponding subspace H0,L is spanned by all states of the form

|a〉s = c+L,aL

c+L−1,aL−1

. . . c+1,a1|0〉 . (21.2)

These are states for which every site is occupied by exactly one electron. Thus,

(1− nj)|a〉s = 0 (21.3)

for j = 1, . . . , L and for all |a〉s ∈ H0,L.Hence, we can conclude with (20.24) and (21.3) that the action of H2 on the space

H0,L reduces to the action of the Hamiltonian

Hspin =∑j,k=1j 6=k

2|tjk|2U

(Sαj S

αk −

1

4

)(21.4)

which is called the (isotropic) Heisenberg Hamiltonian with exchange integrals

Jjk =2|tjk|2U

. (21.5)

If we start from the Hubbard model with nearest-neighbour hopping

tjk =

−t for nearest-neighbour sites0 else

, (21.6)

the resulting ‘spin Hamiltonian’ at half-filling is

Hspin = J∑〈jk〉

(Sαj S

αk −

1

4

), where J =

2t2

U. (21.7)

This is called the Heisenberg-Hamiltonian with nearest-neighbour exchange interaction.The Heisenberg model is the model for the antiferromagnetism of insulators (which isubiquitous in nature).

21.2 Heisenberg model in the language of spin operators

The space H0,L, spanned by the states |a〉s with aj ∈ ↑, ↓, j = 1, . . . , L, is 2L dimen-sional, hence isomorphic to

(C2)⊗L. This makes it possible to describe the action of Hspin

directly by certain matrices acting on(C2)⊗L.

The vectors space(C2)⊗L has the canonical basis vectors

|a〉 = ea1 ⊗ · · · ⊗ eaL , e↑ =

(1

0

), e↓ =

(0

1

). (21.8)

Page 105: Introduction to Solid State Physics - arXiv

105

We would like to identify these vectors with the vectors |a〉s. Then, on the one hand,

Sαj |a〉s =1

2

(σα)abc+L,aL

. . . c+j+1,aj+1

c+j,acj,bc

+j,aj︸ ︷︷ ︸

=c+j,ajc+j,acj,b+δ

bajc+j,a

. . . c+1,a1|0〉

=1

2

(σα)aaj|(a1, . . . , aj−1, a, aj+1, . . . , aL)〉s , (21.9)

while, on the other hand,

1

2

(σα)aaj|(a1, . . . , aj−1, a, aj+1, . . . , aL)〉

=1

2ea1 ⊗ · · · ⊗ eaj−1 ⊗

(σα)aajea︸ ︷︷ ︸

σαeaj

⊗eaj+1 ⊗ . . . eaL

=1

2

(I⊗(j−1)2 ⊗ σα ⊗ I⊗(L−j)

2

)︸ ︷︷ ︸

=Sαj

|a〉 . (21.10)

Thus, the isomorphism H0,L∼=(C2)⊗L induces the identification of operators

1

2I⊗(j−1)2 ⊗ σα ⊗ I⊗(L−j)

2 7→ 1

2

(σα)abc+j,acj,b . (21.11)

Since the operators σα, α = x, y, z, and I2 form a basis of End(C2), the operators

Sαj , j = 1, . . . , L, as defined in (21.10), together with the identity, generate a basis of

End(C2)⊗L. With the identification (21.11) we can interpret the Heisenberg model on

‘spin space’(C2)⊗L with Hamiltonian (21.7), where now the Sαj are the spin matrices

defined in (21.10).

Remark. Such an identification is not possible for the t-J model. The operator H2 actson H0 which contains H0,L as a proper subspace. We have e.g. Sαj |0〉 = 0 (where |0〉 isthe vacuum for Fermions).

21.3 Interpretation of the exchange interaction

Define

P =1

2

(1 + σα ⊗ σα

)=

1

11

1

. (21.12)

Then

Px⊗ y =

1

11

1

x1y1

x1y2

x2y1

x2y2

= y ⊗ x . (21.13)

This means that P is a permutation (transposition) matrix. In particular,

P | ↑↓〉 = | ↓↑〉 . (21.14)

Page 106: Introduction to Solid State Physics - arXiv

106

Define ‘symmetrizer’ and ‘antisymmetrizer’

P± =1

2(P ± 1) . (21.15)

They satisfy

(P±)2

=1

4(P ± 1)2 =

1

2(1± P ) = ±P± , (21.16a)

P+P− = P−P+ = 0 . (21.16b)

This means that ±P± are orthogonal projectors onto the symmetric and antisymmetricsubspaces of C2 ⊗ C2. The subspace P+C2 ⊗ C2 has a basis

Bt =

|1,1〉 = | ↑↑〉, |1,0〉 =

1√2

(| ↑↓〉+ | ↓↑〉

), |1,−1〉 = | ↓↓〉

, (21.17)

while the 1d subspace P−C2 ⊗ C2 is generated by

Bs =

|0,0〉 =

1√2

(| ↑↓〉 − | ↓↑〉

). (21.18)

These bases are called ‘spin triplet’ and ‘spin singlet’.Comparing (21.12), (21.15) and (21.7) we see that for L = 2

Hspin = JP− . (21.19)

It follows that Hspin has eigenvalue 0,−J . The corresponding eigenvectors are |1, s〉,s = 0,±1, and |0,0〉. If J > 0, then the singlet |0,0〉 is the ground state. This is calledantiferromagnetism, since

σz ⊗ σz|0,0〉 = −|0, 0〉 , ⇒ 〈0,0|σz ⊗ σz|0,0〉 = −1 . (21.20)

One says that the ground state has antiferromagnetic correlations.In the general case

Hspin = J∑〈jk〉

P−jk (21.21)

is a sum of projectors onto ‘local singlets’. An antiferromagnet (J > 0) may therefore becharacterized as a system which prefers local singlets. By contrast, a ferromagnet (J < 0)prefers triplets. Local singlets are incompatible with the global SU(2) invariance of theHamiltonian. For this reason the ground state of a macroscopic Heisenberg antiferromagnetis a complicated ‘many-body state’. On the other hand, the state

| ↑ . . . ↑〉 = e⊗L↑ (21.22)

may be seen as a tensor product of states from local triplets. It is annihilated by Hspin andis one of its ground states in the ferromagnetic case. The full ground state subspace can beconstructed using the global SU(2) symmetry.

Page 107: Introduction to Solid State Physics - arXiv

107

21.4 Mott transition

In an external electromagnetic field the hopping term in the Hubbard Hamiltonian ismodified by so-called Peierls phases

tjk → tjk eiλjk , λjk ∈ R . (21.23)

(cf. section 19.4). The Hubbard interaction U , on the other hand remains, unchanged. Itfollows that the corresponding Heisenberg model (strong-coupling limit at half-filling) isnot modified, since it depends only on |tjk|2. In other words, to leading order perturbationtheory the Hubbard model at half-filling does not couple to an external field. The systemis an insulator, although it has an odd number of electrons per unit cell, and therefore isa conductor at U = 0. This suggests that there might be an interaction induced metal-insulator transition (‘Mott transition’) somewhere in between, at a finite value Uc of theinteraction. It is believed that such a transition occurs indeed in any spatial dimension. Aproof only exists in d = 1, where Uc = 0.

21.5 Exercise 20: A short XXX chain

Consider the periodic Heisenberg Hamiltonian (XXX model) on a four-site 1d lattice

H = P12 + P23 + P34 + P41 . (21.24)

Here Pjk is the transposition, interchanging spin states on sites j and k of the chain. Recallthat the space of states of the model is the tensor product H = C2 ⊗ C2 ⊗ C2 ⊗ C2. Thevectors e↑ = ( 1

0 ) and e↓ = ( 01 ) form a basis of C2. Hence,

B =|a1a2a3a4〉 = ea1 ⊗ ea2 ⊗ ea3 ⊗ ea4 ∈ H

∣∣ a1, a2, a3, a4 =↑, ↓

(21.25)

is a basis of H.Denote the embeddings of the Pauli matrices into End(H) by σαj , j = 1,2,3,4, α =

x, y, z. Then the transpositions can be expressed as

Pjk =1

2

(id +σαj σ

αk

). (21.26)

The operator of the total spin has components Sα = 12(σα1 + σα2 + σα3 + σα4 ) which can

be used to define the ladder operators S± = Sx ± iSy. The shift operator for the chain oflength four is defined as

U = P12P23P34 . (21.27)

(i) Show that the operators H , S2, Sz and U can be simultaneously diagonalized. Forthis purpose verify the commutation relations

[H,Sα] = [U , Sα] = [H, U ] = 0 , [Sα, Sβ] = iεαβγSγ (21.28)

for α, β = x, y, z.

(ii) Construct a basis of common eigenvectors and obtain the corresponding eigenvaluesof the above operators. Sketch the spectrum of H . Hint: Use, for instance, thatU4 = id, and use the angular momentum algebra.

Page 108: Introduction to Solid State Physics - arXiv

108

L22 Linear response theory

In this lecture we study the response of a quantum many-body system, in thermal equi-librium with a heat bath of temperature T , to a small external perturbation. A solid isresponding, for instance, with a current to an external voltage, with a thermal current toa temperature gradient, or with a deformation to external mechanic stress. An examplewhich will be considered in some detail is the absorption of microwaves by a spin systemin a homogeneous magnetic field (ESR experiment). This will allow us to take a glimpseat our own work [4, 15].

22.1 Time evolution of the statistical operator

Consider a quantum system with Hamiltonian H possessing a discrete spectrum (En)∞n=0

with corresponding eigenstates |n〉∞n=0. At time t0 a time-dependent perturbation V (t) isadiabatically switched on. We are interested in the time evolution of the system, assumingit initially, at times t < t0, in an equilibrium state described by the statistical operator

ρ0 =1

Z

∞∑n=0

e−EnT |n〉〈n| (22.1)

of the canonical ensemble. Here T denotes the temperature and Z the canonical partitionfunction.

LetU(t) the time evolution operator of the perturbed system. It satisfies the Schrodingerequation

i∂tU(t) =(H + V (t)

)U(t) (22.2)

with initial condition U(t0) = id (note that we have set ~ = 1 as before). Under theinfluence of the perturbation the eigenstate state |n〉 evolves into |n, t〉 = U(t)|n〉, and thestatistical operator at time t becomes

ρ(t) =1

Z

∞∑n=0

e−EnT |n, t〉〈n, t| = U(t)ρ0U

−1(t) . (22.3)

We are interested in an approximation to ρ(t) linear in the strength of the perturbationV (t). In order to derive this approximation we define

R(t) = eiHt(ρ(t)− ρ0

)e−iHt , (22.4a)

W (t) = eiHt V (t) e−iHt . (22.4b)

Then

i∂tR(t) = i∂t eiHt U(t)ρ0

(eiHt U(t)

)−1

= [W (t), eiHt ρ(t) e−iHt] = [W (t), R(t) + ρ0] . (22.5)

Since R(t0) = 0 by construction, we obtain

R(t) = −i

∫ t

t0

dt′ [W (t′), R(t′) + ρ0] . (22.6)

Page 109: Introduction to Solid State Physics - arXiv

109

This Volterra equation is an appropriate starting point for a perturbation theory. Assuming|W (t)| to be small we conclude that

R(t) = −i

∫ t

t0

dt′ [W (t′), ρ0] + O(|W |2

), (22.7)

and therefore, to lowest order in W ,

ρ(t) = ρ0 − i e−iHt

∫ t

t0

dt′ [W (t′), ρ0] eiHt . (22.8)

This is the statistical operator in so-called Born approximation. In the following t0 will besent to −∞.

22.2 Time evolution of expectation values

Using (22.8) we can calculate the time evolution of the expectation value of an operator Aunder the influence of the perturbation. We shall denote the canonical ensemble average by〈 · 〉T = trρ0 · and write A(t) = eiHtA e−iHt for the Heisenberg time evolution of A.Then

δ〈A〉T = tr(ρ(t)− ρ0)A = −i

∫ t

−∞dt′ tr

[W (t′), ρ0] eiHtA e−iHt

= −i

∫ t

−∞dt′⟨[A(t),W (t′)]

⟩T

= −i

∫ t

−∞dt′⟨[A(t− t′), V (t′)]

⟩T. (22.9)

Here we have used the cyclic invariance of the trace in the third equation and the fact thatH commutes with ρ0 in the fourth equation.

A typical example of a perturbation, which will be relevant for our discussion below, isa classical time-dependent field hα(t) coupling linearly to operators Xα,

V (t) = hα(t)Xα . (22.10)

In this case

δ〈A〉T = −i

∫ t

−∞dt′ hα(t′)

⟨[A(t− t′), Xα]

⟩T. (22.11)

22.3 Absorption of energy

The absorbed energy per unit time is

dE

dt=

d

dttrρ(t)(H + V (t))

= −i tr[H + V (t), ρ(t)](H + V (t))+ trρ(t)V (t)

= 〈V (t)〉T + δ〈V (t)〉T . (22.12)

Here we used (22.2), (22.3) in the second equation and the cyclic invariance of the trace inthe third equation. Assuming that V (t) is of the form (22.10) and using (22.11) we obtain

dE

dt= hα(t)〈Xα〉T − i

∫ t

−∞dt′ hα(t)hβ(t′)

⟨[Xα(t− t′), Xβ]

⟩T. (22.13)

Page 110: Introduction to Solid State Physics - arXiv

110

22.4 Application to quantum spin chains

Let us now apply the above formalism to the Heisenberg-Ising (alias XXZ) spin chain in alongitudinal static magnetic field of strength h. The Hamiltonian of this model is definedas

H0 = J

L∑j=1

(sxj−1s

xj + syj−1s

yj + ∆szj−1s

zj

). (22.14)

Here we have switched from Pauli matrices σα to spin operators sα = σα/2. We shallassume periodic boundary conditions sα0 = sαL, α = x, y, z. The real parameter ∆ iscalled the anisotropy parameter. For ∆ = 1 the Hamiltonian H0 turns into the HeisenbergHamiltonian considered in lecture L21. Values of ∆ different from one may be neededfor a more accurate modeling of real magnetic materials and are due to a combination ofcrystal symmetry, spin-orbit interactions and dipole-dipole interactions.

We define the operator of the total spin as

S =

SxSySz

, Sα =1

2

L∑j=1

σα for α = x, y, z. (22.15)

If a spin system is exposed to a homogeneous magnetic field h a so-called ‘Zeemanterm’ −〈h,S〉 must be added to the Hamiltonian. Assuming a field in z-direction ourHamiltonian takes the form

H = H0 − hSz , (22.16)

Note that the z-direction is special in that [H0, Sz] = 0.

We perturb the spin chain by a circular polarized electro-magnetic wave propagating inz-direction. We assume that the wave length is large compared to the length of the spinchain† and idealize this assumption by setting the wave number k = 0. Then the magneticfield component of the wave is

h(t) = A

cos(ωt)− sin(ωt)

0

, A > 0 . (22.17)

It couples to the system by another Zeeman term

V (t) = hα(t)Sα . (22.18)

Using (22.13) we obtain

dE

dt= 〈Sα〉T hα(t)− i

∫ t

−∞dt′⟨[Sα(t− t′), Sβ]

⟩Thα(t)hβ(t′)

=A2ω

4

∫ ∞0

dt′

eiω(2t−t′)⟨[S+(t′), S+]⟩T− e−iω(2t−t′)⟨[S−(t′), S−]

⟩T

+ eiωt′⟨[S+(t′), S−]

⟩T− e−iωt′

⟨[S−(t′), S+]

⟩T

. (22.19)

Here we have used that

〈Sα〉T hα(t) = 〈Sz〉T hz(t) = 0 , (22.20)†The wavelength of microwaves is of the order of 10 cm.

Page 111: Introduction to Solid State Physics - arXiv

111

since 〈S±〉T = 〈Sx ± iSy〉T = ±〈[Sz, S±]〉T = 0 which holds, in turn, because Hcommutes with Sz . Under the integral we have used (22.17),⟨

[Sα(t− t′), Sβ]⟩Thα(t)hβ(t′)

= −A2ω⟨[sin(ωt)Sx(t− t′) + cos(ωt)Sy(t− t′), cos(ωt′)Sx − sin(ωt′)Sy]

⟩T

=iA2ω

4

⟨[eiωt S+(t− t′)− e−iωt S−(t− t′), eiωt′ S+ + e−iωt′ S−]

⟩T

=iA2ω

4

⟨eiω(t+t′)[S+(t− t′), S+]− e−iω(t+t′)[S−(t− t′), S−]

+ eiω(t−t′)[S+(t− t′), S−]− e−iω(t−t′)[S−(t− t′), S+]⟩T. (22.21)

The ability to absorb radiation is a material property. Hence, we generally expect theabsorbed energy per unit time to be proportional to the number of constituents of a physicalsystem and to diverge in the thermodynamic limit. In order to define a quantity that trulycharacterizes the material and is finite in the thermodynamic limit we should thereforenormalize by the average intensity A2 of the incident wave and by the number of latticesites L. Further averaging the normalized absorption rate over a half-period π/ω of theapplied field, we obtain the normalized absorbed intensity

I(ω, h) =ω

LA2π

∫ πω

0dtdE

dt

4L

∫ ∞0

dt

eiωt⟨[S+(t), S−]

⟩T

+ e−iωt⟨[S+, S−(t)]

⟩T

=

ω

4L

∫ ∞−∞

dt eiωt⟨[S+(t), S−]

⟩T. (22.22)

Introducing the function

χ′′+−(ω, h) =1

2L

∫ ∞−∞

dt eiωt⟨[S+(t), S−]

⟩T, (22.23)

the normalized absorbed intensity can be written as

I(ω, h) =ω

2χ′′+−(ω, h) . (22.24)

The function χ′′+−(ω, h) is called the (imaginary part of) the dynamic susceptibility. Thisfunction is a typical ‘response function’. Note that it is calculated as Fourier transform ofthe dynamical correlation function 〈[S+(t), S−]〉T .

L23 Microwave absorption by the Heisenberg-Ising chain

23.1 The isotropic chain

In the general case χ′′+− cannot be calculated. For ∆ = 1, however, the situation simplifiesdrastically.

[H,S] = −h[Sz,S] , (23.1)

Page 112: Introduction to Solid State Physics - arXiv

112

and the Heisenberg equation of motion for S can be solved,

S± = i[H,S±] = −ih[Sz, S±] = ∓ihS± , ⇒ S±(t) = e∓iht S± . (23.2)

Also Sz(t) = Sz . Hence, the total spin behaves as

S(t) =

cos(ht) sin(ht)− sin(ht) cos(ht)

1

S . (23.3)

It is precessing clockwise about the z axis.On the other hand, inserting (23.2) into (22.23) we obtain

χ′′+−(ω) =1

2L

∫ ∞−∞

dt ei(ω−h)t⟨[S+, S−]

⟩T

= 2πδ(ω − h)m(T, h) . (23.4)

where m(T, h) = 〈Sz〉T /L is the magnetization per lattice site. The correspondingnormalized absorbed intensity is

I(ω) = πδ(ω − h)hm(T, h) (23.5)

and is proportional to the magnetic energy hm(T, h) per lattice site. This case includesthe familiar paramagnetic resonance (Zeeman effect) for which the magnetization is knownmore explicitly, namely, m(T, h) = 1

2 th(h

2T

)for J = 0. In general, an exact calculation

of the magnetization of the isotropic Heisenberg chain at any finite temperature is notelementary and requires the machinery of Bethe Ansatz and quantum transfer matrix [9].

Comparing (22.17), (23.3) and (23.5) we interpret the absorption of energy as aresonance between the rotating field of the incident wave and the precessing total spin ofthe chain. Both are rotating clockwise with angular velocity ω = h. If we deviate from theisotropic point ∆ = 1 of the Hamiltonian (18.15) we expect that energy is transferred formthe ‘coherent motion of the total spin’ to ‘other modes’, causing a damping of the spinprecession and hence a shift and a broadening of the δ-function shaped spectral line (23.5).

23.2 Resonance shift and line width in the anisotropic case

At the present state of the art dynamical correlation functions, such as χ′′+−, cannot becalculated exactly, not even for models as simple as the XXZ chain. The interactioninduced shift of the resonance frequency and the width of the spectral line, on the otherhand, are more simple quantities which need less information to be calculated.

For every finite L and for every n ∈ N the integrals

In =

∫ ∞−∞

dω ωnI(ω) (23.6)

exist. Since I(ω) is everywhere non-negative and since I0 > 0 (see below), we mayinterpret I(ω)/I0 as a probability distribution with moments In/I0. I1/I0 is the mean

value of the distribution and(I2/I0 − (I1/I0)2

) 12 its variance. The quantities may be used

to define the ‘resonance frequency’ and the line width.Instead of the moments of the normalized absorption intensity we shall use the ‘shifted

moments’ of the dynamical susceptibility,

mn =

∫ ∞−∞

2πJn(ω − h)nχ′′+−(ω) . (23.7)

Page 113: Introduction to Solid State Physics - arXiv

113

As we shall see these quantities are more natural from a theoretical point of view as theycan be more easily calculated. Using the binomial formula we can relate them to themoments of the normalized intensity,

InI0

=

∫∞−∞ dω ωn+1χ′′+−(ω)∫∞−∞ dω ωχ′′+−(ω)

=

∑n+1k=0 h

kJn+1−kmn+1−kJm1 + hm0

. (23.8)

Let us now define

δω = 〈ω〉 − h =I1

I0− h = J

Jm2 + hm1

Jm1 + hm0. (23.9)

δω is the ‘resonance shift’, i.e., the deviation of the resonance frequency from the resonancefrequency in the isotropic case. Similarly,

∆ω2 = 〈ω2〉 − 〈ω〉2 =I2

I0− I2

1

I20

= J2Jm3 + hm2

Jm1 + hm0− δω2 (23.10)

is a measure of the line width. Hence, in order to calculate the resonance shift and the linewidth, we need to know the first four shifted moments m0, m1, m2, m3 of the dynamicsusceptibility χ′′+−.

23.3 The shifted moments of the dynamic susceptibility

In the following we shall employ the notation adX · = [X, · ] for the adjoint actionof an operator X . Then, using that adH = adH0 −h adSz , that [H0, S

z] = 0 and that[Sz, S+] = S+, we see that

S+(t) = eit adH S+ = eit adH0 e−iht adSz S+ = e−iht eit adH0 S+ . (23.11)

Thus,

χ′′+−(ω) =1

2L

∫ ∞−∞

dt ei(ω−h)t⟨[eit adH0 S+, S−]

⟩T

=1

2L

∫ ∞−∞

dt ei(ω−h)t⟨[S+, e−it adH0 S−]

⟩T. (23.12)

It follows that

(ω − h)nχ′′+−(ω) =(−i)n

2L

∫ ∞−∞

dt(∂nt ei(ω−h)t

)⟨[S+, e−it adH0 S−]

⟩T

=1

2L

∫ ∞−∞

dt ei(ω−h)t⟨[S+, adnH0

e−it adH0 S−]⟩T, (23.13)

entailing that

mn =1

2L

⟨[S+, adnH0/J

S−]⟩T. (23.14)

The latter formula shows that the moments mn are static correlation functions whoserange and complexity grows with growing n. The first few of them can be easily calculatedby hand. The most basic one is

m0 =1

2L

⟨[S+, S−]

⟩T

=1

L

⟨Sz⟩T

= m(T, h) , (23.15)

Page 114: Introduction to Solid State Physics - arXiv

114

which is the magnetization per lattice site. The subsequent moments vanish in the isotropicpoint ∆ = 1. It turns our that they are polynomials in

δ = ∆− 1 . (23.16)

Unlike the magnetization they do not have an immediate interpretation. Using (23.14) theycan be calculated one by one. We obtain, for instance,

m1 = δ〈s+1 s−2 − 2sz1s

z2〉T , (23.17a)

m2 =1

2δ2〈sz1 + 4sz1s

z2sz3 − 4sz1s

+2 s−3 〉T , (23.17b)

m3 =1

4δ2⟨2s+

1 s+2 s−3 s−4 + 4s+

1 s−2 s

+3 s−4 − 2s+

1 s−2 s−3 s

+4 − 8sz1s

z2s

+3 s−4 − 4sz1s

+2 s

z3s−4

+ 8sz1s+2 s−3 s

z4 − 4s+

1 s−2 − s+

1 s−3 + 8sz1s

z2sz3sz4 + 2sz1s

z3 − 4sz1s

z2

+ δ(8sz1s+2 s−3 s

z4 + 2s+

1 s−2 − 8sz1s

z2)⟩T. (23.17c)

The moments are certain combinations of static short-range correlation functions. Thisimplies, in particular, that they all exist in the thermodynamic limit. Substituting (23.17)in to (23.9) and (23.10) we have expressed our measures for the resonance shift and linewidth in terms of short-range static (rather than dynamical) correlation functions.

We close this subject with a number of comments.

(i) Short range static correlation functions of the XXZ chain can be calculated exactlyat all values of temperatures and magnetic fields.

(ii) Our formulae show that special combinations of short-range static correlation func-tions can be measured, at least in principle, by macroscopic experiments.

(iii) In practice such measurements are difficult due to limitations of the experimentalaccuracies and limitations to experimental techniques, or due to the fact, that δ intypical spin chain materials is too small.

(iv) It follows from the existence of the moments that the shape of the ESR absorptionlines cannot be Lorentzian, as is assumed by many experimentalists and in most ofthe more conventional theoretical approaches.

(v) Based on the expressions (23.15), (23.17) for the moments in terms of correlationfunctions we can prove and specify our claim that I(ω) is generally positive, evenfor vanishing magnetic field h. Using (23.15) and (23.17a) in I0 = π(Jm1 + hm0)and noting that 〈s+

1 s−2 〉T = 2〈sx1sx2〉T we obtain

I0 = π〈hsz1 + 2δJ(sx1sx2 − sz1sz2)〉T . (23.18)

Here the first term in the brackets is positive if h 6= 0 as the magnetization is positivefor positive h and negative for negative h. Note, however, that it becomes extremelysmall for small temperatures in the massive phase δ > 0. As for the second term,for small enough h and δ > −1 the neighbour correlators are negative, and the zz-correlations are weaker than the xx-correlations for negative δ and stronger than thexx-correlations for positive δ. Hence, the second term is positive even for vanishingh as long as δ is non-zero.

Page 115: Introduction to Solid State Physics - arXiv

115

(vi) The resonance shift and line width as determined by (23.9) and (23.10) show asimple scaling behaviour as functions of the exchange interaction J . Namely δω/Jand ∆ω/J depend on J only through the ratios T/J and h/J . This is true for thestatistical operator ρ, hence also for all spin correlation functions, and then by ourformulae (23.15), (23.17) for δω/J and ∆ω/J as well. Note also that δω/J and∆ω/J both vanish proportional to δ as we approach the isotropic point δ = 0.

23.4 Exercise 21: Shifted moments

Obtain m1 and m2 from (23.14).

Page 116: Introduction to Solid State Physics - arXiv

116

References

[1] F. Bloch, Uber die Quantenmechanik der Elektronen in Kristallgittern, Z. Phys. 52(1929), 555–600.

[2] M. Born and R. Oppenheimer, Zur Quantentheorie der Molekeln, Ann. d. Phys. 389(1927), 457–484.

[3] Max Born and Kun Huang, Dynamical theory of crystal lattices, The Clarendon Press,Oxford, 1954.

[4] M. Brockmann, F. Gohmann, M. Karbach, A. Klumper, and A. Weiße, Theory ofmicrowave absorption by the spin-1/2 Heisenberg-Ising magnet, Phys. Rev. Lett. 107(2011), 017202.

[5] P. Debye, Zur Theorie der spezifischen Warmen, Ann. d. Phys. 344 (1912), 789–839.

[6] F. H. L. Essler, H. Frahm, F. Gohmann, A. Klumper, and V. E. Korepin, The One-Dimensional Hubbard Model, Cambridge University Press, 2005.

[7] M. C. Gutzwiller, Effect of correlation on ferromagnetism of transition metals, Phys.Rev. Lett. 10 (1963), 159.

[8] J. Hubbard, Electron correlations in narrow energy bands, Proc. R. Soc. (London) A276 (1963), 238.

[9] A. Klumper, Thermodynamics of the anisotropic spin-1/2 Heisenberg chain andrelated quantum chains, Z. Phys. B 91 (1993), 507.

[10] E. H. Lieb and F. Y. Wu, Absence of Mott transition in an exact solution of theshort-range, one-band model in one dimension, Phys. Rev. Lett. 20 (1968), 1445,Erratum: ibid. 21 (1968) 192.

[11] N. D. Mermin, A short simple evaluation of expressions of the Debye-Waller form, J.Math. Phys. 7 (1966), 1038.

[12] B. S. Shastry, Exact integrability of the one-dimensional Hubbard-model, Phys. Rev.Lett. 56 (1986), 2453.

[13] A. Sommerfeld, Zur Elektronentheorie der Metalle auf Grund der Fermischen Statis-tik, Z. Phys. 47 (1928), 1.

[14] L. van Hove, The occurrence of singularities in the elastic frequency distribution of acrystal, Phys. Rev. 89 (1953), 1189.

[15] J. Zeisner, M. Brockmann, S. Zimmermann, A. Weiße, M. Thede, E. Ressouche,K. Yu. Povarov, A. Zheludev, A. Klumper, B. Buchner, V. Kataev, and F. Gohmann,Anisotropic magnetic interactions and spin dynamics in the spin-chain compoundCu(py)2Br2: An experimental and theoretical study, Phys. Rev. B 96 (2017), 024429.