Anda di halaman 1dari 173

Geometry and Topology

in Electronic Structure Theory

Raffaele Resta

Notes subject to ongoing editing


This version run through LATEX on 10Mar16 at 18:14
Contents

1 Introduction 1
1.1 About the present Notes . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.1 Notice to the reader . . . . . . . . . . . . . . . . . . . . . 1
1.1.2 Why writing these Notes? . . . . . . . . . . . . . . . . . . 1
1.2 What topology is about . . . . . . . . . . . . . . . . . . . . . . . 2
1.2.1 Gauss-Bonnet theorem . . . . . . . . . . . . . . . . . . . . 3
1.2.2 Euler characteristic . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Electronic wavefunctions . . . . . . . . . . . . . . . . . . . . . . . 4
1.4 Units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.5 Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.6 Gauge and flux . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.6.1 Classical mechanics . . . . . . . . . . . . . . . . . . . . . . 6
1.6.2 Quantum mechanics, open boundary conditions . . . . . . 7
1.6.3 Quantum mechanics, periodic boundary conditions . . . . 7
1.6.4 Example: Free particle in 1d . . . . . . . . . . . . . . . . 7
1.6.5 Flux and flux quantum . . . . . . . . . . . . . . . . . . . 8

2 Early discoveries 10
2.1 The Aharonov-Bohm effect: A paradox? . . . . . . . . . . . . . . 10
2.2 Conical intersections in molecules . . . . . . . . . . . . . . . . . . 11
2.3 Quantization of the surface charge . . . . . . . . . . . . . . . . . 14
2.4 Integer quantum Hall effect . . . . . . . . . . . . . . . . . . . . . 15
2.4.1 Classical theory (Drude-Zener) . . . . . . . . . . . . . . . 15
2.4.2 Landau levels . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.4.3 The experiment . . . . . . . . . . . . . . . . . . . . . . . . 17
2.4.4 Early theoretical interpretation . . . . . . . . . . . . . . . 18

3 Berryology 21
3.1 Phases and distances . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.2 Berry phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.3 Connection and curvature . . . . . . . . . . . . . . . . . . . . . . 24
3.4 Chern number . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.5 Metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.6 Sum over states . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.7 Time-reversal and inversion symmetries . . . . . . . . . . . . . . 27
3.8 NonAbelian case . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.9 Bloch orbitals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

i
4 Manifestations of the Berry phase 32
4.1 A toy-model Hamiltonian . . . . . . . . . . . . . . . . . . . . . . 32
4.1.1 Connection and curvature . . . . . . . . . . . . . . . . . . 32
4.1.2 Chern number . . . . . . . . . . . . . . . . . . . . . . . . 33
4.1.3 Berry phase . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.1.4 Numerical considerations . . . . . . . . . . . . . . . . . . 33
4.2 Early discoveries reinterpreted . . . . . . . . . . . . . . . . . . . . 35
4.2.1 Aharonov-Bohm effect . . . . . . . . . . . . . . . . . . . . 35
4.2.2 Molecular Aharonov-Bohm effect . . . . . . . . . . . . . . 36
4.2.3 Digression: the Z2 invariant . . . . . . . . . . . . . . . . . 37
4.2.4 Integer quantum Hall effect (TKNN invariant) . . . . . . 38
4.2.5 Classical limit of TKNN . . . . . . . . . . . . . . . . . . . 40
4.3 Adiabatic approximation in a magnetic field . . . . . . . . . . . . 41
4.4 Anomalous Hall effect . . . . . . . . . . . . . . . . . . . . . . . . 43
4.5 Semiclassical transport . . . . . . . . . . . . . . . . . . . . . . . . 44
4.5.1 Textbook equations of motion . . . . . . . . . . . . . . . . 45
4.5.2 Modern equations of motion . . . . . . . . . . . . . . . . . 45
4.5.3 Equations of motion in symplectic form . . . . . . . . . . 46
4.5.4 Geometrical correction to the density of states . . . . . . 47
4.5.5 Outstanding consequences of the modified density of states 48
4.6 Quantum transport . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.6.1 Transport by a single state . . . . . . . . . . . . . . . . . 48
4.6.2 Current carried by filled bands . . . . . . . . . . . . . . . 49
4.6.3 Quantization of charge transport . . . . . . . . . . . . . . 50

5 Macroscopic polarization 51
5.1 Polarization and electric field . . . . . . . . . . . . . . . . . . . . 51
5.2 Polarization itself vs. polarization difference . . . . . . . . . . 52
5.3 Independent electrons . . . . . . . . . . . . . . . . . . . . . . . . 54
5.3.1 The King-Smith and Vanderbilt formula . . . . . . . . . . 54
5.3.2 The quantum of polarization . . . . . . . . . . . . . . . . 55
5.3.3 Wannier functions . . . . . . . . . . . . . . . . . . . . . . 56
5.3.4 The surface charge theorem . . . . . . . . . . . . . . . . . 57
5.3.5 Noncrystalline systems: The single-point Berry phase . . 59
5.4 Correlated wavefunctions . . . . . . . . . . . . . . . . . . . . . . 60
5.4.1 Single-point Berry phase again . . . . . . . . . . . . . . . 60
5.4.2 Kohn-Sham polarization vs. real polarization . . . . . . . 61
5.5 Polarization as a Z2 topological invariant . . . . . . . . . . . . . 62

6 Quantum metric and the theory of the insulating state 64


6.1 Nongeometrical theories of the insulating state . . . . . . . . . . 64
6.2 Metric-curvature tensor . . . . . . . . . . . . . . . . . . . . . . . 65
6.2.1 Open boundary conditions . . . . . . . . . . . . . . . . . . 65
6.2.2 Periodic boundary conditions . . . . . . . . . . . . . . . . 66
6.2.3 Sum over states again . . . . . . . . . . . . . . . . . . . . 67
6.2.4 Localization tensor (a.k.a. cumulant moment) . . . . . . . 67
6.2.5 Discrete formula (single point) . . . . . . . . . . . . . . . 68
6.3 Localization vs. conductivity . . . . . . . . . . . . . . . . . . . . 69
6.3.1 Linear response . . . . . . . . . . . . . . . . . . . . . . . . 69
6.3.2 Conductivity in the scalar potential gauge . . . . . . . . . 70

ii
6.3.3 Sum rules . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
6.3.4 Screened vs. unscreened field . . . . . . . . . . . . . . . . 72
6.4 Localization in the insulating state . . . . . . . . . . . . . . . . . 72
6.4.1 Independent electrons . . . . . . . . . . . . . . . . . . . . 73
6.4.2 Band insulators and band metals . . . . . . . . . . . . . . 75
6.5 Wannier and hermaphrodite orbitals . . . . . . . . . . . . . . . . 76
6.5.1 Wannier orbitals . . . . . . . . . . . . . . . . . . . . . . . 76
6.5.2 Hermaphrodite orbitals . . . . . . . . . . . . . . . . . . . 77
6.5.3 Maximally localized Wannier orbitals . . . . . . . . . . . 78
6.6 More about conductivity in metals . . . . . . . . . . . . . . . . . 79
6.6.1 Drude phenomenological conductivity . . . . . . . . . . . 79
6.6.2 Conductivity in the vector potential gauge . . . . . . . . . 80
6.7 Localization in different kinds of insulators . . . . . . . . . . . . 82
6.7.1 Small molecules . . . . . . . . . . . . . . . . . . . . . . . . 82
6.7.2 Band insulators . . . . . . . . . . . . . . . . . . . . . . . . 83
6.7.3 Correlated (Mott) insulators . . . . . . . . . . . . . . . . 84
6.7.4 Disordered (Anderson) insulators . . . . . . . . . . . . . . 86

7 Topological invariants and topological insulators 88


7.1 Chern invariant . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
7.1.1 Computing the Chern number . . . . . . . . . . . . . . . 89
7.1.2 Chern number as the curvature for unit area . . . . . . . 89
7.1.3 Chern number and hermaphrodite orbitals . . . . . . . . . 91
7.1.4 Lowest Landau level . . . . . . . . . . . . . . . . . . . . . 92
7.2 Quantum Hall insulators: Correlated wavefunction . . . . . . . . 93
7.3 Topological insulators . . . . . . . . . . . . . . . . . . . . . . . . 95
7.3.1 Haldanium . . . . . . . . . . . . . . . . . . . . . . . . . . 95
7.3.2 Time-reversal symmetric topological insulators . . . . . . 96
7.4 Digression: Bulk-boundary correspondence . . . . . . . . . . . . 97
7.5 Topological order in r space . . . . . . . . . . . . . . . . . . . . . 98

8 Orbital magnetization 100


8.1 Magnetization and magnetic field . . . . . . . . . . . . . . . . . . 101
8.2 Magnetization itself . . . . . . . . . . . . . . . . . . . . . . . . 102
8.2.1 The operator P r Q . . . . . . . . . . . . . . . . . . . . . . 103
8.2.2 Magnetization in k space . . . . . . . . . . . . . . . . . . 104
8.2.3 Magnetization in r space . . . . . . . . . . . . . . . . . . 106

9 Chern-Simons geometric phase 108


9.1 Polarization revisited . . . . . . . . . . . . . . . . . . . . . . . . . 109
9.2 Chern number revisited . . . . . . . . . . . . . . . . . . . . . . . 109
9.3 Axion angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
9.4 Z2 topological insulators . . . . . . . . . . . . . . . . . . . . . . . 111
9.5 Numerical considerations . . . . . . . . . . . . . . . . . . . . . . 112

A Magnetoelectrics (basic features) 114


A.1 Generalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
A.2 B vs. H fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
A.3 Multiferroics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
A.4 Linear magnetoelectrics . . . . . . . . . . . . . . . . . . . . . . . 116

iii
A.5 Parsing the magnetoelectric effect . . . . . . . . . . . . . . . . . . 116
A.6 Axion term . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

B Fundamentals of piezoelectricity 118


B.1 Generalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
B.2 First order properties . . . . . . . . . . . . . . . . . . . . . . . . 119
B.3 Second order properties . . . . . . . . . . . . . . . . . . . . . . . 119
B.4 Open circuit vs. closed circuit . . . . . . . . . . . . . . . . . . . . 120
B.5 Piezoelectricity from first principles . . . . . . . . . . . . . . . . . 121

C Linear-response theory 123


C.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
C.2 Causality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
C.3 Kramers-Kronig relationships . . . . . . . . . . . . . . . . . . . . 124
C.4 Fluctuation-dissipation theorem (classical) . . . . . . . . . . . . . 125
C.5 Finite-temperature polarizability . . . . . . . . . . . . . . . . . . 127
C.5.1 Zero-field boundary conditions . . . . . . . . . . . . . . . 128
C.5.2 Reaction field . . . . . . . . . . . . . . . . . . . . . . . . . 129
C.6 Quantum vs. classical . . . . . . . . . . . . . . . . . . . . . . . . 130
C.7 Fluctuation-dissipation theorem (quantum) . . . . . . . . . . . . 131

D Advanced lattice dynamics 134


D.1 Huangs phenomenological theory . . . . . . . . . . . . . . . . . . 134
D.2 Infrared absorption . . . . . . . . . . . . . . . . . . . . . . . . . . 136
D.3 Dissipation and lifetime . . . . . . . . . . . . . . . . . . . . . . . 137
D.4 Dynamical matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 139
D.5 Coulomb forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
D.6 A simple approach to alkali metals . . . . . . . . . . . . . . . . . 142
D.7 Ionic crystals: rigid-ion model . . . . . . . . . . . . . . . . . . . . 143
D.8 Zone-center modes: general case . . . . . . . . . . . . . . . . . . 145
D.8.1 Free energy and equations of motion . . . . . . . . . . . . 146
D.8.2 Microscopic meaning of the Born charges . . . . . . . . . 147
D.8.3 Cochran-Cowley formula . . . . . . . . . . . . . . . . . . . 150
D.9 High-symmetry cases . . . . . . . . . . . . . . . . . . . . . . . . . 152
D.10 Finite temperature . . . . . . . . . . . . . . . . . . . . . . . . . . 154
D.11 Anharmonic systems . . . . . . . . . . . . . . . . . . . . . . . . . 155

Bibliography 157

iv
Chapter 1

Introduction

1.1 About the present Notes


1.1.1 Notice to the reader
This version of the Notes, posted at http://www-dft.ts.infn.it/~resta/ is
not final, and probably will never be such. On the contrary, these Notes are
likely to be continuously edited, updated, and augmented. Subsequent versions
will replace the present one at the same URL.
It is easy to guess these Notes to be plagued by many typos and even more
serious errors. I kindly ask any reader spotting such occurrences to notify me
at resta@democritos.it.
The present Notes include some Appendices whose topics are not geometri-
cal or topological, although some of their ingredients are calculated by means
of Berry phases. However, the Appendix topics are regularly included in the
course for which the present Notes are planned.

1.1.2 Why writing these Notes?


The occasion, or I should better say the pretext, for starting to these Notes
is a graduate course delivered at SISSA (International School for Advanced
Studies, Trieste) in 2012 and covering a partand a part onlyof the topics
dealt with here. Further editing continues towards an advanced undergraduate
course delivered at the University of Trieste since March 2013 onwards.
The need for some support material for such courses is hardly compelling,
since many very good review papers exist, whose union covers about the whole
of the present topics, all of them state-of-the-art at the time of their publication.
Some of these reviews are authored or coauthored by me [1, 2, 3, 4, 5, 6, 7, 8, 9],
and many more by outstanding colleagues. Of these, I quote only the most
recent ones; a non exhaustive list is [10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20].
The writing of a set of Lecture Notes in due form appears as a heavy burden,
diverting for a significant amount of time the author from his day-to-day research
work, already also diverted by administration and teaching. Instead, I used
above the word pretext, because these Notes are mostly written for my own
sake. In fact I regard such burden as an important occasion for a pause of
reflection, devoted to critical rethinking about the subject in its wholeness. In

1
Figure 1.1: The hallmark of topology,
as in many popular presentations. Hun-
dreds of figures like this, and even some
very perspicuous videos, can be down-
loaded from the internet. The two
closed surfaces (two-dimensional com-
pact manifolds) have the same topo-
logical invariant g = 1, which measures
the number of handles.

the present case, the pause of reflection was made much easier by extended stays
far from Trieste: at CECAM in Lausanne, and at DIPC in San Sebastian.
I wrote Lecture Notes in four other occasions, last time in 2000 [21]. I
know by experience that the task of writing in uniform notations and in a
logical sequence many results scattered in the literature leads me to scrutinize
from a novel viewpoint not only other peoples work, but even my own. Here
LATEX plays a major role; soon after its appearance in the early 1980s my way
of thinking has changed dramatically. While previously Ilike everybody
reasoned by jotting formulas on a sheet of paper, nowadays I cannot reason
clearly unless the formulas are neatly typeset in LATEX.
The very fact of writing many different known results altogether has the
beneficial effect of making quite evident several links, hitherto unsuspected.
The perspective view gained while writing the present Notes (and typesetting
them in LATEX) will doubtless influence the future course of my research activity,
as in fact already happened in the previous occasions.

1.2 What topology is about


Topology is defined as a branch of mathematics that describes properties which
remain unchanged under smooth deformations; such properties are usually la-
belled by integer numbers, named topological invariants. The concepts and
tools belonging to topology are continuity and connectivity, open and closed
sets, neighborhoods, and the like.
Differentiability, or even a metric structure, are not needed; theorems are
proved under very general hypotheses, and are therefore very powerful, being
applicable to very diverse frameworks. The tradeoff is that proofs, and even def-
initions, look clumsy and obscure to readers with the mathematical background
of a typical condensed matter physicist. The good news is that the topological
properties most relevant for electronic structure theory can be formulated in the
more familiar language of differential geometry.
Many introductions to topology start with the statement that, to a topolo-
gist, a coffee cup and a doughnut are the same thing, as in Fig. 1.1. Intuitively,
the common feature of the two objects is the presence of one, and only one,
handle. The mathematical definition of handle is coming soon.

2
Figure 1.2: A sphere of radius R, and
its tangent plane in a generic point. The
Gaussian curvature in this trivial case is
= 1/R2 .

1.2.1 Gauss-Bonnet theorem


We start with the simplest example, a sphere, and a tangent plane at a given
point. In a local system of Cartesian coordinates on the plane the equation of
the sphere is
p x2 + y 2
z = R R2 x2 y 2 , (1.1)
2R
and the Hessian matrix is
2 z 2z
! !
1/R 0
x2 xy
H= 2 z 2z
= . (1.2)
yx y 2 0 1/R

The Gaussian curvature is by definition the determinant of the Hessian at the


tangency point. It is obviously constant and equal to 1/R2 at any point of the
sphere; notice that the orientation of the z axis (either inwards or outwards) is
irrelevant. The integral of over the whole closed surface is 4.
Next we consider a smooth (i.e. twice differentiable) closed surface of arbi-
trary shape: the Gaussian curvature is defined as the determinant of the Hessian
at the tangent plane, similarly to what we did for the sphere:
2 2
!
z z
x2 xy
= det 2 z 2z
. (1.3)
yx y 2

In general, can be positive, negative (at a saddle point), or zero (e.g. for
a plane or a cylinder). The Gauss-Bonnet theorem states that for any closed
smooth surface
1
Z
d = 2(1 g), (1.4)
2 S
where g is a nonnegative integer, called the genus of the surface. Surfaces
which can be continuously deformed into each other (i.e. homeomorphic)
have the same genus. For the sphere and any surface homeomorphic to it g = 0;
both the coffee cup and the doughnut, Fig. 1.1 have g = 1; a double-handle
cup has g = 2. The genus is thus the mathematical definition for the number
of handles.

3
1.2.2 Euler characteristic
We have considered smooth surfaces so far, but topological invariants are
based on the more general condition of continuity, andto the delight of
mathematiciansunsurprisingly can be defined even for pathological surfaces
(manifolds in topology-speak). The simplest non smooth case addresses poly-
hedra, where the Gaussian curvature is either zero (on the faces) or singular (at
vertices and edges).
The Euler characteristic is defined as = V E + F , where V is the num-
ber of vertices, E is the number of edges, and F is the number of faces. If we
address the set of regular polyhedra (tetrahedron, cube, octahedron, dodeca-
hedron, icosahedron) it is easily verified that = 2. All these surfaces can be
continuously deformed into (are homeomeorphic to) each other, and into a
sphere. In fact there is a one-to-one relationship between the Euler character-
istic and the genus: = 2(1 g). Polyhedra can also have 6= 2, like the
doughnut-shaped one shown in Fig. 1.4.

1.3 Electronic wavefunctions


In the domain of electronic structure, the typical object addressed via geomet-
rical and/or topological concepts is the electronic ground state of some system.
Whenever an observable effect has the nature of a topological invariant, i.e.
it is an integer number, two remarkable features occur. (1) The observable
is measurable in principle with infinite precision (109 is actually attained for
the quantum Hall effect). (2) The observable is very robust under even strong
variations of the sample conditions; a very disruptive perturbation is needed to
switch from one integer to another. Topology concerns mostly insulators: in
this case the disruptive perturbation amounts to crossing a metallic state.
These Notes are entirely devoted to physical properties having a topological
and/or geometrical character. I am not sure of always using the right seman-
tics. Loosely speaking, I would use the term topological for something which
is quantized, and geometrical for something which is not. The framework
and the mathematical tools are often the same for quantized and nonquantized
quantities, the former frequently occurring as special cases of the latter.
The Berry phase is the typical geometrical quantity which is not quantized,
although it can be quantized in high-symmetry cases. The macroscopic po-

Figure 1.3: Some surfaces and their genus

4
Figure 1.4: A doughnut shaped polyhe-
dron. This surface has Euler character-
istic = 0 or, equivalently, genus g = 1.

larization of a solid is a Berry phase, and is obviously (from an experimental


viewpoint) a nonquantized observable. Nonetheless, there are aspects of the
modern theory of polarization that I would define topological. The same ap-
plies to other geometrical properties considered in this Notes. It is reassuring
that even other authors often use geometrical and topological as synony-
mous, and that an illustrious author like Michael Berry confesses an original
mistake about the semantics [22].
Finally, a few words about the many calculations cited and sometimes briefly
outlined here. Unless otherwise stated, the term first-principle calculations,
when referred to a condensed matter system, means density functional calcula-
tions; independent-electron eigenfunctions and eigenvalues are the Kohn-Sham
(KS) ones. Despite these Notes mostly address a computational physics reader-
ship, no technical details are given (basis sets, pseudopotentials, functionals...);
they are obviously detailed in the original literature, while the focus here is on
the physical properties.

1.4 Units
We use Gaussian electromagnetic units throughout: these have the advantage
(at variance with SI units) that electric and magnetic fields have the same dimen-
sions. Furthermore, the nasty 0 and 0 disappear; SI formulas are converted
by setting 40 = 1 and 4/0 = 1.
For a single particle, the Newton equation of motion and the Hamiltonian
read, respectively  
dv 1
M =f =Q E+ vB , (1.5)
dt c
 2
1 Q
H= p A(r) + Q(r). (1.6)
2M c
Generally, Gaussian electromagnetic units are associated to mechanical cgs
units, but this is no means necessary. In electronic structure theory, it is ex-
pedient to associate Gaussian electromagnetic units with atomic units (a.u.),
defined as e2 = 1, me = 1, ~ = 1. The unit of energy is the hartree (1 Ha = 2
Ry = 27.21 eV). In the present Notes the electron charge is e, with e > 0; this
sign choice agrees with most (but not all) the recent literature. For instance, the
very popular review of Ref [1] adopts the symbol e for the algebraic electron
charge (e < 0).
The speed of light in a.u. is c = 137. This immediately hints at why the
largest atomic number Z in the periodic table is Z 100: in fact the core
electrons have (in a.u.) energies of the order of Z 2 , hence velocities of the order
of Z.

5
1.5 Symbols
I am faced here with two contrasting issues: adopting the symbols most currently
used in the literature, and adopting different symbols for different objects. This
proved to be near to impossible in a work of the present kind, if baroque symbols
are to be ruled out. For instance along the present Notes I do use A, A, A, A, A ,
all with a different meaning. Similarly, I use P, P , P, P, P. Despite this, I
found unavoidable to usein different Chaptersthe same symbol for different
objects. For instance, depending on the context, the symbol P may indicate
a projector or, otherwise, a one-dimensional electrical polarization. Therefore
caution is in order when extrapolating a given symbol from its own context.

1.6 Gauge and flux


We consider here a simple exercise which plays the role of a very important
paradigm; it illustrates basic concepts and results which are going to reappear
several times all along the present Notes.
We address the single-particle Hamiltonian
1 e
H= (p + A)2 + V (r), (1.7)
2m c
where the vector potential A is independent of space and time. It is usually
said that A is a pure gauge, meaning with this that it does not affect the fields:
1 A
B = A 0, E= 0. (1.8)
c t

1.6.1 Classical mechanics


Let us first adopt a classical viewpoint. The Hamilton equation of motions are
H
p = = V (r) (1.9)
r
H 1 e
r = = (p + A). (1.10)
p m c
From these we get
e
p = mr A, (1.11)
c
which leads to the Newton equation of motion

mr = V (r). (1.12)

The bottom line looks quite obvious: a pure gauge has no effect. A basic tenet
of classical mechanics is that the equations of motion can always be directly
expressed in terms of the forces (i.e. the fields), while the potentialsscalar
and vectorare auxiliary quantities, devoid of physical meaning.

6
1.6.2 Quantum mechanics, open boundary conditions
Next we switch to quantum mechanics. It is expedient to rewrite Eq. (1.7) as
1 e
H() = (p + ~)2 + V (r), = A, (1.13)
2m c~
where , having the dimensions of an inverse length, will be referred to as
twist in the following. The Schrodinger equation is
H()|n ()i = n ()|n ()i. (1.14)
The eigenvectors and eigenvalues of the Schrodinger equation depend on the
boundary conditions assumed.
The so-called open boundary conditions (OBCs) require that the bound
eigenstates are square-integrable over R3 . Let |n (0)i be a nondegenerate
eigenstate of the untwisted Hamiltonian within OBCs. Then the state
eir |n (0)i obviously obeys OBCs as well, and also obeys Eq. (1.14) with
a -independent eigenvalue. Therefore it coincides with the n-th eigenstate
|n ()i of the twisted Hamiltonian; notice that this eigenstate is arbitrary by
a -dependent phase factor.
We conclude that a pure gauge within OBCs affects the wavefunction, but
does not affect any of the observable quantities, such as expectation values,
density, and current. We spell this, in jargon, by saying that the twist is
easily gauged away within OBCs.

1.6.3 Quantum mechanics, periodic boundary conditions


We assume periodic boundary conditions (PBCs) over a cubic box of side L, i.e.
we require the eigenstates of Eq. (1.14) to be Born-von-Karman periodic with
period L over x, y, and z at any given . Each Cartesian coordinate is therefore
equivalent to an angle, e.g. x = 2x/L.
If |n (0)i is an eigenstate of the untwisted Hamiltonian within PBCs, then
the state eir |n (0)i obeys Eq. (1.14) with a -independent eigenvalue, but
for a general it does not obey PBCs, and therefore in general does not coincide
with the genuine eigenstate |n ()i. Within PBCs the spectrum of Eq. (1.14)
depends on the twist in a nontrivial way.
If |n ()i is an eigenstate of Eq. (1.14) within PBCs with eigenvalue n (),
then the auxiliary state |n ()i = eir |n ()i obeys the untwisted ( = 0)
Schrodinger equation, and quasi-periodical (a.k.a. twisted or skewed)
boundary conditions: at any two opposite faces of the cube the wavefunction
differs by a -dependent phase factor.
In other words the problem can be formulated in two equivalent ways: either
the Hamiltonian is -dependent, as in Eq. (1.14), and the boundary conditions
are -independent; or the Hamiltonian is -independent but the boundary con-
ditions are twisted in a -dependent way.

1.6.4 Example: Free particle in 1d


For the sake of simplicity, we consider Eq. (1.14) in 1d, and with V 0:
2
~2

d
i + |n ()i = n ()|n ()i. (1.15)
2m dx

7
The eigenfunctions within PBCs and the spectrum are
2n
hx|n ()i ei L x , n Z, (1.16)
2
~2

2n
n () = + , (1.17)
2m L
where the nontrivial -dependence is perspicuous. The velocity operator can be
written as
1 H
v= , (1.18)
~
and the Hellmann-Feynman theorem yields
1 dn ()
hn |v|n i = . (1.19)
~ d
We have introduced PBCs as a basic framework of condensed matter physics.
Many concepts (like the Bloch vector or the Fermi surface) make sense only
within PBCs. But we also may regard this problem as if the electrons were
confined to a circular rail of circumference L, as in Fig. 1.5. There is no field
(electric or magnetic) on the rail, but a constant vector potential of intensity
A = c~/e is present along the rail; eigenvectors and eigenfunctions depend on
its value.

1.6.5 Flux and flux quantum


The constant vector potential A on the circular rail corresponds to a magnetic
flux = LA threading the surface encircled by the rail, in a region not vis-
ited by the electronic system; it has been appropriately called by some authors
inaccessible flux.
We further observe that the spectrum, Eq. (1.17), is periodic in with period
2/L; alternatively, it is periodic in the flux with period 0 = 2~c/e = hc/e,
the elementary flux quantum. In cgs units hc/e = 4.135107 gauss cm2 , while
in SI units 0 = h/e = 4.136 1015 Wb. Notice also that, in the framework
of superconductivity, the same symbol 0 indicates one half of this (it refers to
electron pairs).
We stress that only the fractional part of the flux affects the results in a
nontrivial way. This is perspicuous if we recast Eq. (1.17) as
2  2
~2

2
n () = n+ . (1.20)
2m L 0

Figure 1.5: The electron motion is con-


fined to a circular rail. A constant vec-
tor potential A = c~/e along the rail,
as in Eq. (1.15), corresponds to vanish-
ing fields (electric and magnetic), yet
the spectrum depends on the inaccessi-
ble flux threading the surface encircled
by the rail.

8
The flux breaks time-reversal symmetry ( ), and the spectrum is non-
degenerate, except when = 0 or = 0 /2, the latter also called flux. In
these two cases (and in these cases only) the eigenfunctions can be chosen as
real.
When the flux is varied with time, an emf is induced along the loop. Using
Eq. (1.19), the current is
dn
I = ev = c . (1.21)
d
This result is remarkable: it holds even in presence of a potential V (x), and
generalizes straightforwardly to N noninteracting electrons. It will be used in
the discussion of the quantum Hall effect: see Eq. (2.17) below.

9
Chapter 2

Early discoveries

2.1 The Aharonov-Bohm effect: A paradox?


The Aharonov-Bohm effect is the paradigm for a measurable effect induced by
an inaccessible flux. We anticipate that in many other phenomena such flux may
be purely geometrical or topological, without any relationship to a genuine
magnetic field: this is e.g., the case considered in the next Section. It is only
in the Aharonov-Bohm effect that one addresses indeed the inaccessible flux of
a magnetic field, as present e.g. inside a solenoid. An interference experiment
detects the presence of the flux even when the electronic motion is confined in
the region outside the solenoid, where the magnetic field is zero. This seems
paradoxical: something which happens in a region not visited by the quantum
particle may affect some observable properties. Indeed, the founding fathers of
quantum mechanics (in the 1920s) failed to notice such peculiar feature. It only
surfaced more than 30 years afterwards in the milestone paper by Aharonov and
Bohm [23], appeared in 1959, whose abstract states verbatim ...contrary to the
conclusions of classical mechanics, there exist effects of potential on charged
particles, even in the regions where all the fields (and therefore the forces on the
particles) vanish.
The paper was shocking, and its conclusions were challenged by several au-
thors; nonetheless experimental validations appeared as early as 1960 [25, 26].
The main message of Ref. [23] is at the basis of many subsequent developments

Figure 2.1: The Aharonov-Bohm interference experiment (From Ref. [24])

10
in electronic structure theory, many of them illustrated below in the present
Notes. The Aharonov-Bohm effect is also at the root of the commercial SQUID
technology [27].
It is remarkable that R. P. Feynman included the Aharonov-Bohm effect
in his legendary lectures, delivered to the sophomore class at Caltech during
the 1962-63 academic year [24]. In the final sentence about this topic, Feynman
says: ...E and B are slowly disappearing from the modern expression of physical
laws; they are being replaced by A and .
It is also remarkable and shameful that a paper bearing the title Nonex-
istence of the Aharonov-Bohm effect [28] was published as late as 1978. All
challenges disappeared with the publication in 1984 of the celebrated paper by
Michael Berry (now Sir Michael Berry [29]), where the eponymous phase made
its first appearance [30].

2.2 Conical intersections in molecules


At the time Berry wrote his famous paper, only two occurrences of a geometrical
phase (called Berry phase soon afterwards) in quantum mechanics were known to
him: the Aharonov-Bohm effect and a somewhat exotic phenomenon occurring
in molecular physics. Even the latter was known since the late 1950s [31, 32],
and appropriately rebaptized in the late 1970s as molecular Aharonov-Bohm
effect [33, 34]. In the subsequent years Berry phases were discovered in many
branches of physics.
The smallest molecular system where the molecular Aharonov-Bohm effect is
possible is a trimer, having three internal coordinates (e.g. the three internuclear
distances), and the simplest trimers are of course the homonuclear ones, where
symmetry plays a major role. I give a simple outline for this particular system:
a dynamical Jahn-Teller effect, bearing the conventional symmetry label E .
We focus on a trimer of monovalent atoms, e.g. H3 or Na3 , and we assume
an independent-electron picture in the Born-Oppenheimer approximation. We
start with the molecule in the equilateral configurations, Fig. 2.2. Two of the
valence electrons occupy a totalsymmetric orbital, while the unpaired electron
occupies the next available one, which has E symmetry and is doubly degener-
ate. In a simple tight-binding (alias minimal-basis LCAO) scheme, a possible
basis in the two-dimensional manifold is:
1 1
|1i = ( |Bi|Ci ) ; |2i = ( 2|Ai|Bi|Ci ) , (2.1)
2 6
where A,B,C are atomic labels (as in the figure). This choice deserves an impor-
tant comment. We are adopting OBCs, as appropriate for an isolated molecule,
and the Hamiltonian is invariant under time reversal (no magnetic field, no
spin-orbit interaction). These two conditions guarantee that the orbitals may
always be chosen as real. They may, but they dont need: it may instead be
convenient to choose a complex basis in the same two-dimensional degenerate
manifold.

Figure 2.2: A homonuclear trimer in its equilateral configu-


ration.

11
Figure 2.3: A schematic representation of a (counterclockwise) pseudorota-
tion, where subsequent snapshots differ by 2/3. The corresponding electronic
ground states in the tight-binding approximation, are also shown.

When we distort the molecule from its equilateral configuration, the doublet
is linearly split: one of the two components is energetically favored, the molecule
undergoes Jahn-Teller distortion, and the electronic ground state in the Born-
Oppenheimer approximation becomes nondegenerate.
Next we analyze the motion of the nuclei. There are three linearly inde-
pendent normal modes for the small oscillations of the internal coordinates. Of
course, in absence of a Jahn-Teller effect, the equilateral configuration is the
equilibrium one. One mode is totalsymmetric, and cannot split the electronic
levels. The remaining modes are degenerate, having in fact E symmetry, and
couple to the electronic doublet, originating in fact the dynamical Jahn-Teller
effect. The notation E means indeed that an E vibrational mode is coupled
to an E electronic state: conventionally, one uses upper case letters as symmetry
labels for the vibrational states, and lower case ones for the electronic states.
The adiabatic electronic ground state follows the nuclear motion. For a
cyclic pseudorotation, shown in Fig. 2.3, the Hamiltonian is periodical, but the
electronic wavefunction is antiperiodical. The total wavefunction in the Born-
Oppenheimer approximation factors into the electronic one times the nuclear
one. Given that the total wavefunction must be single-valued, even the nuclear
wavefunction must be quantized using antiperiodical boundary conditions, and
this affects the pseudorotation spectrum in a measurable way.
This feature has to do with the peculiar shape of the Born-Oppenheimer
surface, shown in Fig. 2.4. If we adopt a two-dimensional Cartesian normal
coordinate = (1 , 2 ), the ionic displacements are parametrized as:
xA = 1
yA = 2
xB = 21 1 + 3
2 2
yB = 3
2 1
21 2
xC = 12 1 3
2 2 yC = 3 1
2 1 2 2

Figure 2.4: The Born-Oppenheimer sur-


face of the Jahn-Teller split doublet: a
double-valued function with a conical
intersection (a.k.a. diabolical point).
The potential minimum is a circle of
radius min centered at the degeneracy
point.

12
The meaning of this coordinate choice is transparent with reference to Fig 2.3:
when the atom A is displaced by , the displacements of B and C are of equal
magnitude ||, but pointing in directions rotated by 2/3 and 4/3, re-
spectively. If we neglect Jahn-Teller coupling beyond linear order, no potential
energy is associated to a motion at constant ||, which is indeed a free pseudoro-
tation (or a rotation wave), also schematized in the succession of snapshots
in Fig. 2.3.
In absence of Jahn-Teller coupling, the surface would simply be a parabola,
everywhere doubly degenerate. The linear Jahn-Teller splitting is function of
||, hence to linear order the electronic eigenvalues are:

E () ||2 const ||. (2.2)

This double-valued function is displayed in Fig. 2.4 and has a conical intersection
at the origin. The double cone is also called a diabolo (after a spinning toy of
the same shape), so the degeneracy points are also called diabolical. The
lowest sheet E () has a circular valley of radius min , where a classical particle
travels freely (if nonlinear Jahn-Teller coupling is neglected). Nothing exotic
happens if the nuclear motion can be considered as classic; but when we quantize
the nuclear degrees of freedom, antiperiodical boundary conditions have to be
imposed for the cyclic motion, as said above.
A simple approximation for the rotovibrational levels is thus:
1
E(u, j) = (u + ) 0 + Aj 2 , (2.3)
2
corresponding to an oscillation of frequency 0 and quantum number u, and
a two-dimensional internal rotation with rotor constant A. The antiperiodical
boundary conditions imply half-odd-integer values for the quantum number j;
notice that the ground state (j = 1/2) is twofold degenerate.
The pseudorotation term in the spectrum can be compared to Eq. (1.20);
the moment of inertia in the prefactor becomes obviously a nuclear rotor, but
the spectrum is the same if we identify the inaccessible flux with half a flux
quantum 0 (a.k.a. flux).
There is no magnetic field in this problem; the flux is purely topological
and can be regarded as an obstruction: the nuclear path cannot be contracted
without crossing a degeneracy point. It is remarkable that the topological nature
of this problem was clearly stated as early as 1963much earlier than topology
became fashionable in electronic structureby Herzberg and Longuet-Higgins
[32], who say verbatim: ...a conically self-intersecting potential surface has a
different topological character from a pair of distinct surfaces which happen to
meet at a point. Indeed, if an electronic wave function changes sign when we
move round a closed loop in configuration space, we can conclude that somewhere
inside the loop there must be a singular point at which the wave function is
degenerate.
In modern jargon, we would say that the cases = 0 and = 0 /2 are topo-
logically distinct; owing to time-reversal invariance, other flux values are ruled
out. We anticipate that the present problem is revisited in order to introduce
the Z2 topological invariant in Sec. 4.2.3.
The present paradigm also illustrates the robustness of topological proper-
ties against smooth deformations. For instance, here we have addressed the

13
ultrasimple tight-binding model, but the ground wavefunction can be contin-
uously deformed to the exact correlated wavefunction: topology-wise, the two
wavefunctions are essentially the same object, insofar as the conical intersection
is present. Notice also that at the conical intersection the Born-Oppenheimer
approximation breaks down.
One could also address more general closed paths, according to their winding
number round the obstruction. Only paths having the same winding number
can be continuously deformed into each other: they are homotopic.

2.3 Quantization of the surface charge


The pioneering selfconsistent calculations of the electronic structure of surfaces,
performed at IBM (Yorktown Heights) and at Bell Labs in the mid 1970s,
pointed out the occurrence of quantization of charge at insulating surfaces.
After an early paper by V. Heine in 1966 [35], the theorem made its appearance
in a 1974 paper by Appelbaum and Hamann [36]. Other papers addressed the
issue in the 1970s [37, 38], but the topological explanation came much later; it
will be discussed below, Sec. 5.3.4.
The quantization of surface charge may appear counterintuitive, if one sticks
at the idea that a solid is an array of classical charges (ions), as many people
still do. Possibly because of its counterintuitive content, this important theorem
is surprisingly ignored even by well known specialists in surface physics. The
extreme of such ignorance occurs in a recently published invited review paper
which I abstain from quotingabout polar surfaces. The theorem is even more
ignored in quantum chemistry, where it addresses end charges in linear polymers.
Electrons are quantum particles, and classical ideas may prove wrong. Solids
are not assemblies of ions; they are assemblies of atoms, having ionic character
only because neighboring atoms have a different electronegativity [39]. At the
surface, one has to look at what happens to the bonds.
A simple statement of the theorem is the following. If the bulk of the crystal
is centrosymmetric, and if the surface is insulating, then the charge per surface
cell may only be an integer or half integer; the surface charge can be nonquan-
tized only if the bulk is noncentrosymmetric, or if the surface is metallic.
Quite often, the actual quantized value is zero because of energy consider-
ations; therefore even polar surfaces are (counterintuitively) neutral under the
above two essential hypotheses, which I stress again: the bulk is centrosymmet-
ric, and the surface is insulating. The microscopic mechanism can be understood
as an intrinsic surface-state neutralization [39]; however, topology guarantees
quantization independently of microscopic details.
We nowadays regard bulk-surface correspondence in many phenomena as
one of the hallmarks of geometry and topology in condensed matter physics. In
modern jargon, I would say that the surface charge of insulators is topologically
protected. More about the bulk-boundary correspondence will be said in Sec.
7.4.

14
2.4 Integer quantum Hall effect
2.4.1 Classical theory (Drude-Zener)
We consider any 2d system, in the setup shown in Fig. 2.5. If dissipation is
accounted for by a single relaxation time , the Newton equation of motion for
a single carrier of mass m and charge e, is
   
dv 1 1
m + v = e E + v B . (2.4)
dt c

Setting dv/dt = 0 we get the steady-state solution:


 
e 1
v= E+ vB . (2.5)
m c
eB
In terms of the cyclotron frequency c = mc the solution with Ey = 0 is
e
vx = Ex c vy
m
vy = c vx . (2.6)

If n is the carrier density, the current is j = ne v:

ne2
jx = Ex c jy
m
jy = c jx . (2.7)

in zero B field we retrieve the standard Drude (diagonal) conductivity:

ne2
jx = 0 Ex , 0 = , (2.8)
m
while for B 6= 0 the conductivity tensor is:
0
jx = Ex = xx Ex
1 + (c )2
c 0
jy = Ex = yx Ex . (2.9)
1 + (c )2

Inversion of the conductivity tensor


xx yx
xx = 2 + 2
xy = 2 + 2
(2.10)
xx yx xx yx

Figure 2.5: Hall effect in a 2d system.


The E field is applied along x, while the
B field is along z. The system is shorted
in the y direction; the current j has both
longitudinal (x) and transverse, a.k.a.
Hall (y) components.

15
provides a remarkably simple expression for the longitudinal and transverse
resistivity
m mc 1
xx = 1/0 = 2 , xy = = B. (2.11)
ne ne2 nec
The Hall resistivity is therefore linear in B and independent of both mass and
relaxation time; more accurately, since we may consider even carriers of positive
charge e (holes), its sign does depend on the carrier charge. Notice also that
in the nondissipative regime ( 1/c ) both xx and xx vanish.
If we write n as N/A (number of carriers per unit area), then
1
xy = B= , (2.12)
nec N ec
where = AB is the magnetic flux through area A. Although we are still at a
purely classical level, it is instructive to multiply and divide xy by h. We may
thus identically write
1 h N 0
xy = , = . (2.13)
e2
Here 0 = hc/e is the flux quantum, as defined above. The dimensionless
quantity , called the filling factor, equals the ratio between the number of
electrons N and the number of flux quanta /0 . Eq. (2.13) expresses the
transverse resistivity in terms of the natural resistance unit h/e2 . Since 1990
this is a new metrology standard, accurate to more than nine figures: 1 klitzing
= h/e2 = 25812.807557(18) ohm.
In 2d resistance and resistivity have the same dimensions, and coincide in
the transverse case. We write therefore the Hall resistance as
1 h
Vy /Ix = RH Rxy = . (2.14)
e2
Upon obvious dimensionality arguments, even in the quantum case the Hall
resistance can be written in this way; but then the concentration- and B-
dependence of are expected to be very different from the simple monothonical
form of Eq. (2.13).

2.4.2 Landau levels


In quantum mechanics, the Schrodinger equation for an electron in 2d subject
to a perpendicular B field (and in a flat potential) can be exactly dealt with,
both in the Landau gauge and in the central gauge. The spectrum is discrete
n = (n + 12 )~c . We define the magnetic length as = (~c/eB)1/2 ; it diverges
in the zero-field limit, and is of the order of 100 angstrom in a typical quantum
Hall experiment. In the Landau gauge (Ax = By, Ay = 0) the eigenfunctions
with energy n are
nk (x, y) eikx n (y 2 k), (2.15)
where n (y) are harmonic oscillator eigenfunctions with frequency c . Each LL
is infinitely degenerate (one eigenfunction for each k). For a system of area A,
the number of states in each level is N = A/(22 ); this has a simple form in
terms of the magnetic flux through area A: N = /0 .
If we now consider N noninteracting electrons, the lowest LL is completely
filled when N = N ; more generally, one expects a periodicity in the filling factor
= N/N = N /0 , whenever crosses an integer value, in most physical
properties.

16
Figure 2.6: The original figure from von
Klitzing et al. Ref. [40]. The gate volt-
age Vg was supposed to control the car-
rier density. Instead, the Hall resistance
is quantized and insensitive to Vg over
a large interval; over the same interval,
the longitudinal resistance vanishes.

2.4.3 The experiment


The Hall resistance of a noninteracting 2d electron gas has been computed
quantum-mechanically by Ando in 1974 [41]. The result, when expressed as in
Eq. (2.14) showed indeed oscillations in with integer period. The experiment,
performed by von Klitzing and collaborators in 1980 [40], provided qualitatively
different and very surprising results, shown in Fig. 2.6. The discovery of the
quantum Hall effect triggered a revolution with far reaching consequences in
electronic structure theory at large; Klaus von Klitzing was awarded the Nobel
prize in 1985.
In the original experiment, the 2d electrons were confined by a MOSFET
(metal-oxide-semi-conductor field-effect transistor); later, higher mobilities were
obtained at semiconductor heterojunctions. Fig. 2.6 shows a very robust

Figure 2.7: A modern realization of the integer quantum Hall effect

17
plateau, where Rxx = 0 and Rxy = 6453.3 0.1, corresponding to the fill-
ing factor = 4. The accuracy in the quantized Rxy value is clearly far beyond
the experimental control of the carrier concentration and of the B field uni-
formity over the sample. A novel, qualitatively different, state of matter was
discovered. In modern jargon, the plateaus are topologically protected.
A modern realization of the integer quantum Hall effect is shown in Fig.
2.7, where xx and xy are plotted as a function of the magnetic field. The
plateau quantization is accurate to nine figures. The 2d electron gas is typically
confined at a GaAs/GaAlAs heterojunction. The = 1 value is achieved above
10 tesla; at low field (high ) the system becomes dissipative (xx > 0), while
the classical linear behavior of xy is recovered; the slope depends on electron
concentration n.

2.4.4 Early theoretical interpretation


The breakthrough paper, by Laughlin, appeared as early as 1981 [42]. This
is a remarkably concise paper (two pages) which, in retrospect, is based on
topological arguments. One key ingredient of the theory is disorder: in fact, the
quantum Hall effect becomes less spectacular for very clean samples, while
some dirtyness enhances the effect.
Laughlin devised a Gedankenexperiment based on the setup shown in
Fig. 2.8. The corresponding 2d Schrodinger Hamiltonian in the Landau gauge
is  
1  e 2 2
H() = px + Ax + py + eEy + V (x, y), (2.16)
2m c
where E is the field across the ribbon, and V is an arbitrary substrate potential.
The addition of a constant vector potential along x, Ax Ax +A, in Eq. (2.16)
corresponds to threading a flux = L A through the cylinder; we use the
symbol , not to be confused with the real magnetic flux normal to the
surface.
Similarly to what discussed in Sec. 1.6, the eigenvalues acquire a de-
pendence. According to Eq. (1.21), if n () is the n-th eigenvalue the current
transported by the corresponding eigenstate is Ix = c n ()/. For an
independent-particle system with N carriers the current is thus

U ()
Ix = c , (2.17)

Figure 2.8: Geometry for Laughlins


Gedankenexperiment. A 2d channel is
bent into a loop of circumference L, and
a magnetic B field of constant magni-
tude pierces the cylinder normal to the
surface. A current I Ix circles the
loop; VH Vy is the Hall voltage. The
loop may be threaded by the inaccessi-
ble flux .

18
Figure 2.9: The density of states for a 2d system of noninteracting carriers. (a)
Clean system, with zero substrate potential. (b) Actual sample, in presence of
substrate disorder and impurities.

where U () is the total energy of the system. Implicitly, we are assuming a


dissipationless system.
The expression in Eq. (2.17) for the current is remarkably simple, general,
and robust: it does not depend on the substrate potential V (x, y), nor the
number N of carriers, and not even on their mass m. But for a disordered
potential the eigenstates come in two kinds: localized and extended. The latter
ones are phase-coherent round the loop, while the former are exponentially
localized for L . The localized states are insensitive to the flux insertion
(like the OBCs eigenstates in Sec. 1.6), and the whole current is carried by
the delocalized ones. Therefore Eq. (2.17) provides a nonzero result insofar
at least one of the occupied eigenstates in the disordered sample is extended,
i.e. phase-coherent round the loop; besides this, the number and nature of the
current-carrying states is irrelevant. It is therefore crucial to address the nature
of the single-particle eigenstates in a quantum Hall sample. For a clean sample
(flat substrate) the LLs are sharp, all eigenstates are extended (in the Landau
gauge), and the density of states is a series of delta functions, shown in Fig. 2.9
(a); the weight of each delta is /0 . In presence of disorder, the deltas broaden
into alternating bands of localized and extended states, as sketched in Fig. 2.9
(b).
The electron fluid is in the quantum Hall regime whenever the Fermi level
falls in a region of localized states. Therefore xx = 0 (the fluid is a quantum
Hall insulator), and xx = 0 (transport is dissipationless).
We now imagine to adiabatically increase the vector potential by an amount
A = 0 /L, where 0 is a flux quantum: all of the current-carrying states are
mapped back into themselves, while the localized ones are unaffected. Hence the
ground state has the same energy; nonetheless Eq. (2.17) implies U ( + 0 )
U () 0 Ix /c 6= 0. This is only possible if an integer number of electrons is
transferred from one cylinder edge to the other, each of them contributing the
energy eVy . If we call such integer number, the relationship is then
0 1 h
0 Ix /c = eVy ; RH = Vy /Ix = = . (2.18)
ce e2
The flux acts therefore as a charge pump; the pump cycle is one flux quantum
0 .
Ideally, the sample ground state can be continuously deformed from dirty
to clean. Insofar as the Fermi level stays is in a region of nonconducting states,

19
the (topological) integer cannot change, even if the number of current carry-
ing states does obviously change. The identification of with the number of
filled LLs comes from the clean-sample limit, which is exactly soluble. Setting
V (x, y) = 0 the eigenfunctions of Eq. (2.16) are

cE
nk (x, y) = eikx n (y y0 ), y0 = 2 k . (2.19)
c B
For a finite L, the allowed ks are integer multiples of 2/L and the correspond-
ing centers y0 are spaced by 22 /L = L0 /. Threading a flux shifts y0
linearly in ; when equals one flux quantum each eigenfunction goes over to
the next. Therefore one carrier is shifted for each n; the integer index measures
therefore the number of occupied LLs. Similar arguments can be reformulated
in different gauges and in different geometries [43, 44].

20
Chapter 3

Berryology

Berryology is used as synonymous for geometry in nonrelativistic quantum me-


chanics. Topological (i.e. quantized) quantities are defined via geometrical
quantities, analogously to what we did above in Sec. 1.2. But many important
quantities (notably the Berry phase) are merely geometrical [22].
Let us assume that a generic time-independent quantum Hamiltonian has a
parametric dependence. The Schrodinger equation is

H()|()i = E()|()i, (3.1)

where the d-dimensional real parameter is defined in a suitable domain of Rd :


a 2d has been chosen for display in Fig. 3.1. In most of this Section we discuss
the most general case, and therefore we do not specify which quantum system is
described by this Hamiltonian, nor what the physical meaning of the parameter
is.
In the subsequent Chapters |()i will be identified with either a single-
particle wavefunction (a.k.a. orbital) or a many-electron wavefunction. As for
the parameter , it may be a nuclear coordinate, a phase angle, a magnetic
flux, a Bloch vector, a momentum, and more: it could therefore have various
dimensions. Sometimes, the parameter is called the slow variable, while the
electronic coordinates are the fast variable. At the end of this Chapter, Sec.
3.9, we address the special case where the parameter is identified with a Bloch
vector.
The state vectors |()i are all supposed to be normalized and to reside
in the same Hilbert space: this amounts to saying that the wavefunctions are
supposed to obey -independent boundary conditions. We focus on the ground
state |0 ()i, and we assume it to be nondegenerate for in some domain of
Rd .
Any quantum mechanical state vector is arbitrary by a constant phase fac-
tor. Here we refer to choosing this phase as to the choice of the gauge. The
semantic is a bit ambiguous. In presence of magnetic fields, we may change the
magnetic gauge: this changes the Hamiltonian and the eigenfunctions. Once
the magnetic gaugehence the Hamiltonianis fixed, we still remain with the
phase arbitrariness referred to above. All measurable quantities (e.g. expecta-
tion values) are obviously gauge-invariant (in both senses), but the reverse is
also true: all gauge-invariant properties areat least in principlemeasurable.
It is expedient to define the ground-state projector (a.k.a. density matrix)

21
and its complement, i.e.
P () = |0 ()ih0 ()|; Q() = 1 P (). (3.2)
Both P () and Q() are gauge-invariant (for a fixed Hamiltonian).

3.1 Phases and distances


We define the phase difference between the ground eigenstates at two different
points in the most natural way:
h0 ( 1 )|0 ( 2 )i
ei12 = ; (3.3)
|h0 ( 1 )|0 ( 2 )i|
12 = Im log h0 ( 1 )|0 ( 2 )i . (3.4)
For any given choice of the two states, Eqs. (3.3) and (3.4) provide a 12
which is unique modulo 2, except in the very special case that the states are
orthogonal. However, it is also clear that such 12 is gauge-dependent and
cannot have, by itself, any physical meaning.
The distance between quantum states has been defined by Bures [45] as:
2
D12 = 1 |h0 ( 1 )|0 ( 2 )i|2 . (3.5)
Such distance fulfills the familiar axioms from calculus textbooks; it vanishes
when the two states physically coincide (i.e. independently of the phase factor),
and is maximum (equal to one) when the states are orthogonal. At variance
with 12 defined above, the Bures distance is gauge-invariant, and can be
explicitly expressed in terms of ground-state projectors
2
D12 = 1 Tr {P ( 1 )P ( 2 )}, (3.6)
where Tr is the trace over the Hilbert space.

3.2 Berry phase


We have already observed that the phase difference 12 between any two states
is gauge-dependent and cannot have any physical meaning. Matters are quite
different when we consider the total phase difference along a closed path which
joins several points in a given order, as shown in Fig. 3.2:
= 12 + 23 + 34 + 41
= Im log h0 ( 1 )|0 ( 2 )ih0 ( 2 )|0 ( 3 )i
h0 ( 3 )|0 ( 4 )ih0 ( 4 )|0 ( 1 )i. (3.7)

Figure 3.1: State vectors in the two-


dimensional -space. The phase dif-
ference between two of them is de-
h ( )| ( )i
fined as ei12 = |h0 (1 )|0 (2 )i| ,
0 1 0 2
2
and their distance as D12 = 1
2
|h0 ( 1 )|0 ( 2 )i| .

22
Figure 3.2: A closed path joining four
states in -space.

It is now clear that all the gauge-arbitrary phases cancel in pairs, such as to
make the overall phase a gaugeinvariant quantity. The above simpleminded
algebra leads to a result of overwhelming physical importance: in fact, a gauge
invariant quantity is potentially a physical observable. In essence, this is the
revolutionary message of Berrys celebrated paper, appeared in 1984 [30, 46].
Next we consider a smooth closed curve C in the parameter domain, such
as in Fig. 3.3, and we discretize it with a set of points on it. Using Eq. (3.3),
we write the phase difference between any two contiguous points as

h0 ()|0 (+)i
ei = . (3.8)
|h0 ()|0 (+)i|

If we further assume that the gauge is so chosen that the phase varies in a
differentiable way along the path, then from Eq. (3.8) we get to leading order
in :
i h0 ()| 0 ()i . (3.9)
In the limiting case of a set of points which becomes dense on the continuous
path, the total phase difference converges to a circuit integral:
M
X I
= s,s+1 A() d, (3.10)
s=1 C

where A() is called the Berry connection:

A() = i h0 ()| 0 ()i. (3.11)

Figure 3.3: A smooth closed curve C in


-space, and its discretization.

23
Since the state vectors are assumed to be normalized at any , the connection
is real; we can therefore equivalently write

A() = Im h0 ()| 0 ()i. (3.12)

A number of manifestations of the Berry phase occurring in molecular and


condensed matter phenomena will be discussed in the present Notes. Many
reviews papers; here we quote Refs. [46, 47, 48, 3, 13].
At this point it is also worth emphasizing that in computational physics there
are no derivatives. The ground state 0 () is generally found by diagonalizing
a matrix on a finite set of points, and the phase (i.e. the gauge) is chosen by
the diagonalization routine; the phase is therefore nonsmooth and possibly even
random. The discrete form in Eq. (3.11) at finite M is the one universally used
in numerical work; it is clearly unaffected by any erratic phase factor.

3.3 Connection and curvature


The loop integral of the Berry connection (i.e. the Berry phase ) is non trivial
in two cases: either the curl of A() is nonzero, or the curl is zero but the curve
C is not in a simply connected domain. In the former case, we can invoke Stokes
theorem; the formulation is very simple when is a 3d parameter. If C is the
boundary of a surface (i.e. C ), and the curl of A() is regular on ,
then Stokes theorem reads
I Z
= A() d = () n d, (3.13)

where is the Berry curvature, defined as

() = A() = Im h 0 ()| | 0 ()i


= ih 0 ()| | 0 ()i, (3.14)

with the usual meaning of the cross product between three-component bra and
ket states. Equation (3.13) may be spelled out by saying that the curvature is
the Berry phase per unit area of .
For d 6= 3 the Berry curvature is conveniently written as the d d antisym-
metric matrix
() = 2 Im h 0 ()| 0 ()i; (3.15)
Greek subscripts are Cartesian coordinates throughout, and = / . The
Stokes theorem can still be applied, generalizing Eq. (3.13) to
1
Z
= d d (). (3.16)
2
The Berry connection is also known as gauge potential, and the Berry
curvature as gauge field [48]. It is worth pointing out that the former is
gauge-dependent, while the latter is gauge-invariant and therefore corresponds
in general to a measurable quantity, even before any integration. The two quan-
tities play (in -space) a similar role as the vector potential and the magnetic
field in elementary magnetostatics: A(r) is gauge-dependent, nonmeasurable;
B(r) = r A(r) is gauge-invariant, measurable.

24
The Berry phase , defined as the integral over a closed curve C of the
connection, is gauge invariant only modulo 2. This indeterminacy is resolved
by Eqs. (3.13) and (3.16) whenever the curve C is the boundary of a surface
where the curvature is regular. In fact, the curvature is gauge-invariant and
has no modulo 2 indeterminacy.

3.4 Chern number


The rhs of Eqs. (3.13) and (3.16) is the flux of the Berry curvature on the surface
; such flux remains meaningful even on a closed surface (e.g. a sphere or a
torus), in which case is the empty set. The key result is that such an integral
is quantized. Here we limit ourselves to 3d, where we identify with the sphere
S 2 (Fig. 3.4):
1
Z
() n d = C1 ; (3.17)
2 S 2
C1 is an integer Z, called Chern number of the first class.
The proof is based on a similar algebra as for Diracs theory of the magnetic
monopole [47, 49]. The theorem goes sometimes under the name of Gauss-
Bonnet-Chern theorem; the analogy with Eq. (1.4) is perspicuous. A specific
example is dealt with in detail in Sec. 4.1.
The curvature is regular (and divergence-free) on the closed surface S 2 ; the
lhs of Eq. (3.17) is the flux of () across S 2 . The integrand () is the curl of
the connection A(); the latter in general cannot be defined as a single-valued
function globally on S 2 , but only on patches of it [47, 49]. To fix the ideas,
suppose that () is singular at = 0, and that S 2 is the spherical surface
centered at the origin (Fig. 3.4). We cut this surface at the equator z = 0 and
we consider the flux across the two open surfaces:
Z Z Z
() n d = () n d + () n d. (3.18)
S2 S+ S

We notice that S+ = S = C, but the surface normals n have opposite


orientations. From Stokes theorem, Eq. (3.13), we get:
Z I
() n d = A () d (3.19)
S C
Z I I
() n d = A+ () d A () d . (3.20)
S2 C C

The two upper and lower Berry connections A () may only differ by a gauge
transformation; the rhs of Eq. (3.20) is the difference of two Berry phases on

Figure 3.4: A sphere cut at the equator


in two hemispheres.

25
the same path and is necessarily a multiple of 2. This concludes the proof of
Eq. (3.17).
We emphasize that the Chern number is a robust topological invariant of
the wavefunction, and is at the origin of observable effects. The Chern number
made its first appearance in electronic structure in 1982, in the famous TKNN
paper about the quantum Hall effect [50] (Sec. 4.2.4). Nowadays more complex
topological invariants are in fashion, and they characterize a completely novel
class of insulators, called topological insulators [10, 11, 15, 51, 52, 53, 54, 55,
56].

3.5 Metric
Starting from Eq. (3.5), the infinitesimal distance is
d
X
2
D ,+d = g ()d d , (3.21)
,=1

where the metric tensor is easily shown to be

g () = Re h 0 ()| 0 ()i
h 0 ()|0 ()ih0 ()| 0 ()i
= Re h 0 ()|Q()| 0 ()i; (3.22)

the projector Q() is the same as defined in Eq. (3.2). This quantum metric
tensor was first proposed by Provost and Vallee in 1980 [57].
At this point we may compare Eq. (3.22) to Eq. (3.15), noticing that the
insertion of Q() is irrelevant in the latter, i.e.

() = 2 Im h 0 ()|Q()| 0 ()i. (3.23)

It is therefore clear that g and are, apart for a trivial 2 factor, the real
(symmetric) and the imaginary (antisymmetric) parts of the same tensor, which
we are going to call F in the following:

F () = h 0 ()|Q()| 0 ()i. (3.24)

The metric-curvature tensor F is gauge-invariant. A compact equivalent ex-


pression is
F () = Tr { P ()Q() P ()}, (3.25)
manifestly gauge-invariant and Hermitian.

3.6 Sum over states


The -derivatives entering many of the previous equations, e.g. Eq. (3.24), can
be expressed starting from perturbation theory:

|0 ( + )i |0 ()i (3.26)
X hn ()| [ H( + ) H() ] |0 ()i
|n ()i ;
E0 () En ()
n6=0

26
X hn ()| H()|0 ()i
| 0 ()i = |n ()i . (3.27)
E0 () En ()
n6=0

These seemingly obvious and innocent formulae need some caveat. It is clear
that inserting Eq. (3.27) into the Berry connection, Eqs. (3.11) and (3.12), we
get a vanishing result for any . This happens because the simple expression
of Eq. (3.26) corresponds to a very specific gauge choice (called the parallel-
transport gauge [3]); multiplying the rhs by a -dependent phase factor is le-
gitimate, and must not modify any physical result, while the Berry connection
is instead affected. Nonetheless, since our F () is a gauge-invariant quantity,
we may safely evaluate it in any gauge, including the parallel-transport gauge,
implicit in Eq. (3.27). The result is
F () (3.28)
X h0 ()| H()|n ()ihn ()| H()|0 ()i
= .
[E0 () En ()]2
n6=0

This expression shows explicitly that both the curvature and the metric are
ill defined and singular wherever the ground state is degenerate with the first
excited state. Indeed, this is the main reason why the domain may happen not
to be simply connected.

3.7 Time-reversal and inversion symmetries


According to Eq. (3.25) the ground-state projector uniquely determines the
curvature
() = 2 Im Tr { P ()Q() P ()}. (3.29)
It is therefore expedient to analyze the symmetries of the ground-state projector
P (), which coincide with the symmetries of the Hamiltonian; these in turn
depend on the nature of the parameter . We only address spinless electrons
(no spin-orbit coupling).
When is even under time reversal (like e.g. a nuclear coordinate), then
time-reversal invariance implies that both H() and P () are real for any , and
Eq. (3.29) warrants that the curvature is everywhere vanishing. The Berry phase
can be nonzero (modulo 2) only if the curve C loops around a singularity or,
more generally, it does not lie in a simply connected domain; the only allowed
values are = 0 or = .
When instead is odd under time-reversal (like e.g. a momentum) then
time-reversal symmetry requires P () = P (), therefore
() = (). (3.30)
The Berry phase along an inversion-symmetric path vanishes; the Chern number
vanishes as well.
Next we switch to inversion symmetry. When evaluating any -dependent
matrix-element (or trace), inversion of the coordinates at fixed is equivalent to
keeping the coordinates fixed and inverting . This statement holds whether
is a coordinate or a momentum; both are in fact odd under inversion. Therefore
inversion symmetry implies P () = P (), and
() = (). (3.31)

27
If both time-reversal and inversion symmetry are present, then the Berry curva-
ture is everywhere vanishing. The Berry phase can be only zero or ; the latter
case requires a domain which is not simply connected, as above.
Crucial to the above arguments is the fact that the double derivative appear-
ing in Eq. (3.29) are even under either time-reversal or inversion. Summarizing
the symmetry results, for the case where is a momentum: a non vanishing
Chern number can only occur in absence of time-reversal symmetry, but may
occur even in inversion-symmetric cases.

3.8 NonAbelian case


Very soon after the appearance of Berrys milestone paper, Wilczek and Zee [58]
introduced a many-state generalization. Suppose we do not focus on a single
state in the Hilbert space (say the ground state), while we are interested instead
in the behavior of n different states altogether, as a function of . The original
motivation was the possibility of degeneracy amongst the n lowest eigenstates
|j ()i of a given Hamiltonian at some points of the path, and they formulated
the theory under the hypothesis that these lowest states are never degenerate
with the n+1-th. Strictly speaking, an Hamiltonian is unnecessary: one only
needs to unambiguously identify an n-dimensional (n fixed) manifold of states,
as function of the parameter . A gauge transformation is then a n n unitary
matrix; in the Abelian case, this matrix is a diagonal one, with unitary elements.
The gauge invariant quantity which defines the manifold is the n-dimensional
projector
Xn
P () = |j ()ihj ()|. (3.32)
j=1

Within loss of generality, we may address the case which is most interesting in
electronic structure. We assume that the n states |j ()i are spin orbitals, and
that |()i is the many-body wavefunction built as their Slater determinant:
1
()i = |1 ()2 () . . . N ()|. (3.33)
N!
We can therefore apply some of the results of the previous Sections. The many-
body Berry phase is given in Eqs. (3.10) and (3.11), which we prefer to rewrite
here in the form  I 
ei = exp i A() d , (3.34)
C
where A is the connection of the many-body wavefunction.
In the nonAbelian case the connection generalizes to a vector of n n Her-
mitian matrices
Ajj () = Im hj ()| j ()i, (3.35)
and the Berry phase factor of Eq. (3.34) generalizes to the unitary matrix
 I 
ei = P exp i A () d . (3.36)
C

Here P is a path-ordering operator, owing to the noncommuting nature of


the A matrices at different in the nonAbelian case. The P operator has a

28
precise meaning when the series expansion of the exponential is considered. The
discretization of Eq. (3.36) is rather straightforward and will not be discussed
here [21].
While the single-state phase factor ei is gauge-invariant, the Wilczek-Zee
unitary matrix ei is only gauge-covariant, i.e. a gauge transformation yields
a matrix unitarily equivalent to ei ; the gauge can be fixed by choosing the
vectors |j ()i at one point of the path. The eigenvalues j of the Hermitian
matrix , defined modulo 2, are gauge-invariant and in principle individually
observable. It is easily proved that their sum, i.e. the trace of , coincides with
the Berry phase of the many-body wavefunction, Eq. (3.34) [21].
In order to write the metric-curvature tensor in the nonAbelian case, we
start writing the many-body Abelian metric-curvature tensor of the Slater-
determinant wavefunction, Eq. (3.24), as

F () = h 0 ()| 0 ()i h 0 ()|P ()| 0 ()i. (3.37)

Next, we need now to explicitate the Cartesian indices , at the same time as
the matrix indices j, j . The nonAbelian metric-curvature tensor is the general-
ized form of Eq. (3.25), i.e.

F,jj () = h j ()| j ()i h j ()|P ()| j ()i, (3.38)

where now P () is the single-particle projector; even this matrix is gauge-


covariant. If we rewrite the Cartesian components of Eq. (3.35) as A,jj () =
Im hj ()| j ()i, the nonAbelian curvature becomes

,jj () = A,jj () A,jj () i[A (), A ()]jj . (3.39)

With respect to Eq. (3.15) notice the presence of an extra term, which vanishes
in the Abelian case. At fixed jj Eq. (3.39) is clearly antisymmetric in the
Cartesian indices, while at fixed it is an Hermitian n n matrix.
The trace over the j index of the (nonAbelian) metric-curvature tensor equals
the corresponding (Abelian) metric-curvature tensor of the many-body ground
state, Eq. (3.25):
n
X
F,jj ()
j=1
Xn n
X
= h j ()| j ()i h j ()|j ()ihj ()| j ()i
j=1 jj =1
= F (). (3.40)

The proof of the last line of Eq. (3.40) is tedious, but straightforward.

3.9 Bloch orbitals


We have remained very general so far. The case where the parameter coincides
with the Bloch vector k bears a particular relevance in the context of the present
Notes. In the framework of first-principle calculations for crystalline systems,
the Bloch states |jk i are the KS orbitals.

29
The domain where the k parameter varies (the reciprocal cell or, equivalently,
the Brillouin zone, BZ) has the geometry of a torus in 1d, 2d, 3d. The whole
BZ is a closed surface; we denote as G a reciprocal vector.
The Bloch orbital of the j-th band is |jk i = eikr |ujk i, where |ujk i are
BZ-periodical, and are eigenfunctions of the Hamiltonian Hk = eikr Heikr .
While the |jk i at different ks are mutually orthogonal, the |ujk i live instead
in the same Hilbert space (they are all BZ-periodical).
The physical meaning of all the mathematical quantities introduced next
will be discussed in the following Chapters, and not anticipated in the present
one.
The Berry connection of the j-th orbital is
Aj (k) = ihujk |k ujk i. (3.41)
The relative phases at different ks are arbitrary. Whenever possible, it is cus-
tomary to enforce the so-called periodic gauge |jk+G i = |jk i, which implies
|ujk+G i = eiGr |ujk i. (3.42)
We stress, however, that in topologically nontrivial crystals it is generally im-
possible to adopt a periodic gauge.
The interesting closed paths C on the torus are lines across the reciprocal
cell, from one face to the opposite one. For an insulator with n occupied bands
the Berry phase is, according to the previous section,
Xn Z n Z
X
=i Aj (k) dk = i dk hujk |k ujk i. (3.43)
j=1 C j=1 C

This Berry phase depends on the choice of the origin in the crystal cell. For
centrosymmetric crystals, if the origin is at a center of inversion symmetry, the
only allowed values are = 0 and = (modulo 2).
In case of band crossings the definition of individual bands is ambiguous,
but the Berry phase in Eq. (3.43) is not affected by such ambiguity. More
generally, the results of the previous Section show that Eq. (3.43) is invariant
by a nonAbelian gauge transformation, i.e. by mixing of the occupied orbitals
between themselves by an arbitrary unitary matrix Ujj (k). The mixed orbitals
are no longer Hamiltonian eigenstates; any gauge where instead the |ujk i are
eigenstates will be called Hamiltonian gauge.
The discretization of Eq. (3.43) is nowadays implemented in most electronic
structure codes [59, 60, 61, 62] in order to compute the macroscopic polarization
of crystalline dielectrics (Sec. 5.3). Such discretization is based on the following
result: if |(k)i is the Slater determinant of the n occupied |ujk i, then
h(k1 )|(k2 )i = det S(k1 , k2 ), (3.44)
where S(k1 , k2 ) is the n n overlap matrix
Sjj (k1 , k2 ) = hujk1 |uj k2 i. (3.45)
We discretize with M+1 points on the line, where kM+1 = k1 +G; it is understood
that |ujkM+1 i = eiGr |ujk1 i. The discretized formula is then
Z
= i dk h(k)|k (k)i Im log M s=1 h(ks )|(ks+1 )i
C
= Im log det M
s=1 S(ks , ks+1 ). (3.46)

30
Notice that this is numerically gauge invariant, i.e. it does not depend on
how the diagonalization routine chooses the phases and/or the ordering of the
eigenvectors.
The metric-curvature tensor of n occupied bands obtains straightforwardly
from Eq. (3.40):
n
X n
X
F (k) = h ujk | ujk i h ujk |uj k ihuj k | ujk i. (3.47)
j=1 jj =1

This is usually integrated over the BZ; being gauge invariant, it carries in prin-
ciple physical meaning even before any integration. Indeed, the k-dependent
(single-band) curvature enters the theory of semiclassical transport in crystalline
solids [63, 64, 13], Sec. 4.5.
Eq. (3.47) is the trace of the nonAbelian metric-curvature whose expression
is
n
X
F,jj (k) = h ujk | uj k i h ujk |ujk ihujk | uj k i
j=1

X
= h ujk |usk ihusk | uj k i. (3.48)
s=n+1

The nonAbelian metric and curvature are


1
g,jj = [ F,jj (k) + F,jj (k) ]
2
,jj = i[ F,jj (k) F,jj (k) ]. (3.49)

The many-band curvature obtains from either Eq. (3.47) or Eq. (3.49) as
n
X
= 2 Im h ujk | ujk i. (3.50)
j=1

31
Chapter 4

Manifestations of the Berry phase

4.1 A toy-model Hamiltonian


Our toy model here is a two-level spinless Hamiltonian, of the form

H() = ~
= (sin cos x + sin sin y + cos z ), (4.1)

where are the three Pauli matrices. The spectrum is non degenerate for
6= 0, and the lowest eigenvalue is . Upon symmetry arguments, we can
already guess the curvature to be isotropic.

4.1.1 Connection and curvature


The lowest eigenvector is

sin 2 ei
 
|(, )i = .
cos 2

This corresponds to a specific gauge choice; the eigenvector can be multiplied by


an overall (, )-dependent phase factor. The Berry connection and curvature
are

A = ih| i = 0

A = ih| i = sin2
2
1
= A A = sin . (4.2)
2
The curvature is gauge-invariant, while the connection is gauge-dependent.
Within our gauge choice the connection displays a vortex at the south pole
( = ); other gauges yield the singularity at a different point, but a singu-
larity is unavoidable. It is impossible to find a gauge which is smooth on the
whole closed surface, and a nonsingular connection; the singularityoften called
obstructioncan be moved but not removed. The algebra is the same as for
Diracs theory of the magnetic monopole [47, 49]: the degeneracy at the origin
is the monopole, and the singularity is the Dirac string.

32
Figure 4.1: A closed curve C on the surface of the
sphere, and the solid angle spanned by it.

4.1.2 Chern number


The domain of the parameters (, ) is a rectangle, which indeed has the topol-
ogy of a torus: a closed surface. Integrating the Berry curvature therein we
get
1 1
Z
dd sin = 1, (4.3)
2 S 2 2
i.e. the Chern number of the lowest eigenstate in this problem. This integer
measures the strength of the singularity (magnetic monopole), which resides in
a site inaccessible to the quantum system. The highest eigenstate has C1 = 1,
since the total Chern number is zero.
This simple example illustrates well the meaning of a topological invariant of
the quantum mechanical ground state. The Chern number is in fact very robust
under continuous deformations of the surface and of the Hamiltonian: its value
is always one insofar as one (and only one) degeneracy point is included in the
closed surface.

4.1.3 Berry phase


Suppose now we evaluate the Berry phase over any closed curve C on the sphere
(Fig. 4.1) I
= A() d. (4.4)
C
Owing to Stokes theorem, the Berry phase for this toy model problem clearly
equals the solid angle spanned by the curve, divided by 2. The inherent 4
arbitrariness in the solid angle leads to the well known 2 arbitrariness in the
Berry phase: for instance at the equator = modulo 2. If we cut the sphere
in two hemispheres, as in Sec. 4.1.2 (see Fig. 3.4), the difference of the boundary
Berry phases equals 0 modulo 2, i.e. 2 times an integer, as it must be. But in
order to tell which integer (the actual Chern number) the two connections are
useless: one has to integrate the curvature, as in Eq. (4.3). Despite this feature,
in numerical work Chern numbers are typically computed via Berry phases, as
described in the next Section.

4.1.4 Numerical considerations


This exactly soluble example also provides the occasion for illustrating the stan-
dard computational approach to Chern numbers. Suppose we discretize the
(, ) domain with a rectangular mesh, and that we diagonalize the Hamil-
tonian at the points of the mesh. The gauge at any point is chosen by the

33
Figure 4.2: Discretization of the domain [0, ]
[0, 2]; the Hamiltonian is not diagonalized at the
empty circles, only at the black ones, thus enforcing
toroidal topology.

diagonalization routine and is thus erratic; we only enforce the toroidal topol-
ogy by requiring that the phases at the opposite edges of the rectangle are the
same: Fig. 4.2.
Then for each small rectangle we compute the discrete Berry phase as in
Eq. (3.7), i.e.

= Im log h(, )|( + , )ih( + , )|( + , + )i


h( + , + )|(, + )ih(, + )|(, )i. (4.5)

In this simple, analytically soluble, case we know the exact value: the Berry
phase on the contour of each rectangle is = 12 [cos()cos(+)] modulo
2. The four-point Berry phase, Eq. (4.5), provides an approximated value for
this. The Chern number is the integral over the domain, and is therefore equal
to the sum of all the phases computed as in Eq. (4.5) and covering the whole
domain.
How do we then get rid of the modulo 2 indeterminacy in Eq. (4.5)? The
size of is small, and each contribution to the sum is also small (pro-
portional to ), although Eq. (4.5) is in principle arbitrary modulo 2. It
should be now pretty clear that the right solution is in choosing the Im log
branch with values in [, ].
As said above, each four-point Berry phase provides only an approximation
to the corresponding analytical integration; it is thus natural to guess that
a discrete computation of the Chern number yields approximately an integer,
coinciding with the exact value only in the limit of a dense mesh. Instead, this
is not the case; even a coarse mesh provides exactly an integer. Indeed, the role
of the mesh is to select which integer: for our toy model either C1 = 0 (if the
singularity is outside the surface), or C1 1 (for the lower and higher state, if
the singularity is inside the surface). This selection is a virtue of the branch
prescription discussed above for the multivalued function Im log.
In order to show that the discretized Chern number is exactly integer we
notice that when we sum over the small rectangles the contributions from their
inner sides cancel in pairs. Therefore the total phase must be equal (mod 2)
to the discrete phase computed on the outer contour of the whole domain. We
already observed that the states at the opposite edges must be the same, with
the same phases (see Fig. 4.2), hence the sum on the edges yields an overlap
product real and positive. This proves that the total (discrete) phase is exactly
zero mod 2.
Last but not least: where is the obstruction? In the continuous formulation,
any gauge choice yields a singularity at a single (, ) point. In the discrete

34
formulation, there is no way to locate the singularity: in some sense, the sin-
gularity is everywhere since the gauge is in principle erratic at all mesh points.
While in mathematicians topology the invariant is robust for continuous defor-
mations of the Hamiltonian and/or of the surface, our computational topology
is robust even for discontinuous deformations and, furthermore, even for a very
coarse mesh. Four points on the sphere, which define a tetrahedron, would be
enough. I regard all this as a manifestation of the unreasonable effectiveness
of topology.

4.2 Early discoveries reinterpreted


4.2.1 Aharonov-Bohm effect
Here we reformulate the Aharonov-Bohm effect as a special case of a Berry
phase. Suppose we have an electron in a box (infinite potential well) centered
at the origin. We take the ground wavefunction as real, and we write it as (r).
The time-independent Schrodinger equation is:
 2 
p
+ V (r) (r) = E(r). (4.6)
2m

Displacing the center of the box at position R changes the Hamiltonian to

p2
H(R) = + V (r R) : (4.7)
2m
we will identify the parameter with the box position R. Because of transla-
tional invariance, the R-dependence of the state vectors is

hr|(R)i = (r R), (4.8)

while the eigenvalue is R-independent.


Suppose now that a magnetic field is switched on somewhere in space. Then
the Hamiltonian becomes
1 h e i2
H(R) = p + A(r) + V (r R) , (4.9)
2m c

quantum
system

Figure 4.3: A particle in a box, trans-


ported round a solenoid

35
where A is the vector potential and e is the electron charge. It can be easily
verified that a solution of the Schrodinger equation can be formally written in
the form:
ie r
 Z 
hr|(R)i = exp A(r ) dr (r R). (4.10)
~c R
But such a solution in general is not a single-valued function of r, since the
phase factor depends on the path. Therefore we restrict ourselves to a less
general case, where the magnetic field is generated by a solenoid: the B field
is nonzero only within a given cylinder, and we dont allow our box to overlap
this cylinder by suitably restricting the domain of R. This situation is sketched
in Fig. 4.3. With such a choice the wavefunction, Eq. (4.10), is a single valued
function of r for any fixed R, and is therefore an honest ground wavefunction.
As for the dependence on R, Eq. (4.10) only guarantees local single-valuedness,
since the domain is not simply connected: when the system is transported on a
closed path winding once round the solenoid, the electron wavefunction picks up
a Berrys phase. This phase difference can be actually detected in interference
experiments.
The Berry connection of the problem is
e
A(R) = ih(R)|R (R)i = ih(R)|R (R)i A(R), (4.11)
~c
where the first term vanishes: (r) is real. Therefore in the present case the
Berry connection is proportional to the ordinary vector potential. A gauge
transformation in the quantum-mechanical sense also coincides with an electro-
magnetic gauge transformation, which changes A while leaving B invariant. In
fact, in this example B is essentially the Berry curvature. The Berry phase is
e 2
I I
= A(R) dR = A(R) dR, (4.12)
~c C 0 C
where 0 is the flux quantum. Therefore measures the flux of the magnetic field
across the interior of the solenoid, a space region not accessed by the quantum
system: above, we have called it inaccessible flux. Only the fractional part of
the flux has physical meaning.

4.2.2 Molecular Aharonov-Bohm effect


Here we identify the slow coordinate with a d-dimensional nuclear coordi-
nate, and the state vector |()i with the electronic ground-state wavefunction
in the Born-Oppenheimer approximation.
We start from the complete Hamiltonian H of an isolated molecular system,
and we explicitly separate the nuclear kinetic energy:
d
1 X
H(, [x]) = M1
p p + H(, [x]), (4.13)
2
,=1

where [x] indicates the electronic degrees of freedom collectively, p =


i~ / is the canonical momentum conjugated to , and the inverse mass
matrix M1 in general may be a function of , but not of the momenta.
The Born-Oppenheimer approximation starts by writing the eigenfunctions
of Eq. (4.13) in the Schrodinger representation as the product h [x] |()i ().

36
Our aim is obtaining an effective Schrodinger equation for the nuclear wave-
function (), where the electronic degrees of freedom have been integrated
out. We start considering the effect of the canonical nuclear momentum p on
the product ansatz:

p |()i () = i~ |()i () i~ | ()i (). (4.14)

We then multiply by the electronic eigenbra h()| on the left, thus integrating
over the electronic degrees of freedom. We get the effective nuclear momentum
acting on as:
() = [ p i~h()| ()i ] () = [ p ~A() ] (), (4.15)

where we easily recognize the Berry connection. The momentum is the kine-
matical (also called covariant, or mechanical) momentum, to be distinguished
from p = i~ , which is the canonical momentum.
Whenever the time scales of nuclear and electronic motions are well sepa-
rated the coupling between different electronic states can be neglected, and the
adiabatic approximations allows to treat the slow variable in H(, [x]) as a
classical parameter. The electronic eigenvalue E() of a given state (e.g. the
ground state) plays therefore the role of a (scalar) potential for nuclear motion,
whose effective Hamiltonian acting on () is then:
d
1 X
Heff = M1
+ E(). (4.16)
2
,=1

In the molecular physics literature the extra term in Eq. (4.15) is seldom
mentioned, and is identified with p. The reason is that for a time-reversal-
invariant Hamiltonian, and in absence of spin-orbit interaction, the wave func-
tion can always be taken as real. This corresponds to the parallel transport
gauge, and the Berry connection vanishes at all ; the tradeoff is thatin some
casesthe electronic wave function is not single valued along a closed path:
see Fig. 2.3. The alternative approach, due to Mead and Truhlar [33, 65], is
to choose a different gauge, where the electronic wave function is single valued
and complex. The Berry phase is gauge invariant (modulo 2); the values al-
lowed by time-reversal symmetry are 0 and ; the two cases are experimentally
distinguishable, as previously shown in Sec. 2.2.
We stress that, whenever the ionic motion is purely classical and governed
by Newtons equation, the vector-potential-like term in Eq. (4.15) is irrelevant:
the corresponding curvature (magnetic-field-like) is in fact identically vanishing
along the nuclear trajectory on the Born-Oppenheimer surface. We anticipate
that the case where a genuine magnetic field is presentand the Hamiltonian
is no longer time-reversal-invariantis qualitatively different in this respect, see
Sec. 4.3 below.

4.2.3 Digression: the Z2 invariant


The topological nature of the molecular Aharonov-Bohm effect has already been
pointed out in Sec. 2.2. Here we want to recast the finding in modern topological
language. The group Z2 is the additive group of the integers modulo 2: it has
only two elements, which can be labeled as 0 and 1, or as even and odd.

37
Owing to time-reversal invariance, the electronic Berry phase in a molecule
on any closed path in configuration spacecan only be either 0 or mod 2,
hence the Z2 nature of this invariant is obvious. We focus on the simplest
possible case, dealt with in Sec. 2.2: a two-level system with (or without)
a conical intersection. We identify our closed surface with the closed loop:
a one-dimensional torus or sphere. In absence of spin-orbit interaction,
a gauge which respects time-reversal symmetry requires the electronic wave
function to be everywhere real. In the nontrivial case (odd, i.e. = ) the
connection cannot be smooth on the whole surface, and necessarily has an
obstruction at a point on the path: see Fig. 2.3. The obstruction can be
moved but not removed, and is due to the monopole at the conical intersection,
in a region not visited by the molecular system.
However, it is a simple exercise to devise a smooth gauge throughout the
path: it is enough to adopt complex electronic wavefunctions. In modern jargon,
the obstruction is removed at the price of breaking time-reversal symmetry.
The topological invariant is easily retrieved from the discrete approach: only
three Hamiltonian diagonalizations on the path are needed. With reference to
Fig. 2.3 we define
1 1 1
|1 i = (|Bi |Ci), |2 i = (|Bi |Ai), |3 i = (|Ci |Ai). (4.17)
2 2 2
Independently of any arbitrary gauge factor, whether it respects time-reversal
symmetry or otherwise, we always have
1
h1 |2 ih2 |3 ih3 |1 i = , (4.18)
8
i.e. = mod 2: the ground state is Z2 odd.
The Z2 topological invariant is extremely robust against smooth deforma-
tions of the Hamiltonian. For instance, here we have addressed the ultrasim-
ple tight-binding model, but the ground wavefunction can be continuously
deformed to the exact correlated wavefunction: topology-wise, the two wave-
functions are essentially the same object: they can be regarded as analogous to
the proverbial doughnut and coffee cup, Figs. 1.1 and 1.3.
Only two conditions are essential to conserve the Z2 invariant: (1) the defor-
mation of the Hamiltonian must conserve time-reversal symmetry, and (2) the
number of conical intersections inside the path must remain the same. Notice
that at the conical intersection the Born-Oppenheimer approximation breaks
down. If (2) is violated, then a singularity has crossed the path during the con-
tinuous transformation: the gap between the ground state and the first excited
state did not remain open.

4.2.4 Integer quantum Hall effect (TKNN invariant)


The famous TKNN (Thouless, Kohmoto, Nightingale, and den Nijs) paper ap-
peared in 1982 [50] and marks the very first occurrence of a topological invariant
(the first Chern number) in electronic structure. The outline provided here is
inspired by Kohmoto [66]. In the quantum Hall context, first Chern number is
synonymous of TKNN invariant.
We consider a two-dimensional independent-electron system in a lattice-
periodical potential, and subject to a perpendicular B field. The Hamiltonian

38
is not translationally invariant, but one can address the magnetic translation
group. We choose a large enough supercell, such that the magnetic flux is
commensurate (i.e. an integer number of flux quanta 0 thread the supercell):
in this case a continuous k vector can be defined in the magnetic Brillouin zone.
As in Sec. 3.9,we define |jk i = eikr |ujk i and Hk = eikr Heikr ; the latter
takes here the form
1 h e i2
Hk = p + ~k + A(r) + V (r), (4.19)
2m c
where V is the substrate potential. The velocity can be expressed as
1
v= k Hk , (4.20)
~
a formula often recurring in the present Notes in various forms, see e.g.
Eqs. (1.18) and (2.17).
The Kubo formula for transverse conductivity is [67]
X 1 Z hujk |vx |uj k ihuj k |vy |ujk i
xy = 2~e2 Im 2
dk (4.21)
(2) jk <F <j k (jk j k )2
jj

e2 X 1 Z hujk |x Hk |uj k ihuj k |y Hk |ujk i


= 2 Im 2
dk .
~
(2) jk <F <j k (jk j k )2
jj

If we now consider the case where the Fermi level lies in a gap, with n filled
bands (Landau levels in a flat potential), Eq. (4.21) becomes the BZ integral
n
e2 1 hujk |x Hk |uj k ihuj k |y Hk |ujk i
Z X X
xy = 2 Im dk . (4.22)
~ (2)2 BZ j=1
(jk j k )2
j =n+1

The integrand is just a simple generalization of the sum-over-states formula of


Eq. (3.28). Using the same arguments as in Ch. 3 it is rather straightforward
to arrive at the identity
n
X X hujk |x Hk |uj k ihuj k |y Hk |ujk i
Im
j=1
(jk j k )2
j =n+1
n
X 1
= Im hx ujk |y ujk i = xy (k), (4.23)
j=1
2

where the many-band Berry curvature, Eq. (3.50), appears. Since the BZ is a
torus, the BZ integral of the curvature equals 2 times an integer, the (first)
Chern number C. The milestone TKNN discovery is that Hall conductivity is
a Chern number when expressed in klitzing1 :

e2
xy = C. (4.24)
h
Notice that the sign choices are not uniform across the literature.
Conductivity is a property of the excitations of the system, as it is perspic-
uous in the Kubo formula above. The Chern number, instead, is a ground state

39
property. The identity relating them belongs to the general class of fluctuation-
dissipation theorems, although this looks like an oxymoron, the Hall conduc-
tivity being here dissipationless. The interpretation of the Chern number as
a ground-state quantum fluctuation will be elaborated somewhere else in the
present Notes.
The topological nature of the observable explains its extreme robustness un-
der variations of magnetic field, carrier density, substrate disorder, and more.
The topological invariant C is identified with the filling using the same argu-
ments as in Sec. 2.4.4; the integer can only be varied by crosssing a conducting
state. It is expedient to change the sign, thus identifying = C. Since in the
quantum-Hall regime the conductivity tensor is off-diagonal, the Hall resistance
(or resistivity) is
1 h
xy = . (4.25)
e2
One final observation is in order: we have not counted a factor of two for double
spin occupation, for a good reason: at the high B fields needed to observe
quantization the Zeeman separation between the two spin states is much larger
than the Landau level separation. Therefore the electrons entering the quantum
Hall effect are spin polarized (or spinless). We quote here the fact that in
graphene instead the experimental conditions for the quantum Hall effect are
different, and the factor of two has to be restored.

4.2.5 Classical limit of TKNN


By virtue of topology, in the quantized (high field) regime the Hall resistivity
xy is insensitive to carrier density n: this was indeed the revolutionary finding
of von Klitzing and collaborators (see Fig. 2.6 and its caption). In the classical
(low field) regime, instead, xy is linear in 1/n. We reproduce here the Drude-
Zener formula of Eq. (2.11):
m B mc
xx = , xy = = , (4.26)
ne2 nec ne2
eB
where c = mc .
As discussed in Sec. 2.4.2 the degeneracy of each Landau level in a flat
potential is /0 , where is the magnetic flux through the sample. The density
of states is
X
D() = [ (n + 1/2)~c ], (4.27)
0 n
shown in Fig. 2.9(a). When a disordered substrate is present D() becomes
instead of the kind shown in Fig. 2.9(b). The density of states is conserved in
average: whenever the chemical potential is much larger than ~c the number
of occupied states will be the same in the clean and in the disordered case. The
density in the ~c limit is therefore
2 B 2 B
Z
n= d D() = . (4.28)
A 0 0 ~c 0
Notice that includes now the spin factor, since the Zeeman splitting vanishes
in the B 0 limit. Expressing now Eq. (4.25) in terms of n we have
B h B
xy = 2
= , (4.29)
n0 e nec

40
Figure 4.4: A modern realization of the integer quantum Hall effect. The dashed
line is the corresponding Drude-Zener transverse conductivity.

which is indeed the classical formula of Eq. (4.26).


At this point we are ready to address the classical limit in the experimental
data of Fig. 2.7, reproduced here as Fig. 4.4. In this sample at the conditions
of the actual experiment the classical-quantum transition occurs somewhere
between 1 and 2 tesla: below this B value the system becomes dissipative (xx 6=
0), and the plateaus in xy disappear. At B 1T we have xy xx ; the Drude-
Zener formulas of Eq. (4.26) suggest therefore that the effective relaxation time
for this sample is 1/c, evaluated at B 1 2 tesla.
In Fig. 4.4 I have indicated (dashed line) the classical Drude-Zener trans-
verse conductivity for this sample: it is linear in B and coincides with the exact
conductivity for integer values of the filling. It is also realized that the slope of
the dashed line is larger than the experimental one in the low-B regime. The
reason is that the effective carrier density in the high-B regime is one-half the
one in the low-B regime: electrons are spin-polarized in the former case, and
unpolarized in the latter.

4.3 Adiabatic approximation in a magnetic field


The general problem of the nuclear motionboth classical and quantum
in presence of an external magnetic field has been first solved in 1988 by
Schmelcher, Cederbaum, and Meyer [68]. It is remarkable that such a funda-
mental problem was solved so late, and that even today the relevant literature
is ignored by textbooks and little cited. The solution is a manifestation of geo-
metrical effects in electronic wavefunctions [69], which appears in a spectacular
way even when the nuclear motion is addressed at the classical level.
When a genuine magnetic field, generated by some external source, acts on
the molecular system, the Hamiltonian of Eq. (4.13) is modified by the addition
of a vector potential term in the kinetic energies of both the nuclei and the
electrons. Proceeding as in the zero field case, one writes an ansatz wavefunction
and arrives at the effective Hamiltonian for the ionic motion, Eq. (4.16), where

41
an extra term must be added to the kinematical momentum of Eq. (4.15).
There are thus two vector potentials in the effective nuclear Hamiltonian: a
geometric one, and a genuinely magnetic one.
However, with respect to the zero-field case, there is a qualitative differ-
ence whose importance is overwhelming. Since the electronic Hamiltonian is no
longer invariant under time-reversal, the electronic wavefunction is necessarily
complex, and the curvature is in general nonzero. No singularity is needed to
produce geometrical effects on the nuclear motion; the Berry phase will be in
general nonzero on any path in the space of nuclear coordinates.
Suppose we are interested into the nuclear motion at the purely classical
level. The Hamiltonian of Eq. (4.16)whose kinematical momentum includes
now the two different vector potentialsyields the Hamilton equations of mo-
tion, which can be transformed into the Newton equations of motion: within the
latter, the effects of the vector potentials appear in terms of fields, in the form
of Lorentz forces. The curl of the magnetic vector potential obviously yields
the magnetic field due to the external source; the curl of the geometric vector
potential (Berry curvature) yields an additional magnetic-like field which is
nonzero even on the classical trajectory of the nuclei. We stress that this is at
variance with the zero-field case, where the Berry phase had no effect on the
ionic motion at the classical level, and could only be detected when quantizing
the ionic degrees of freedom.
Within a nave BornOppenheimer approximationwhere Berry phases are
neglectedthe magnetic field acts on the nuclei as it they were naked charges:
a proper treatment must instead account for electronic screening: this is pro-
vided by the geometric vector potential. Surprisingly, there are very few calcu-
lations of the effect: it is pretty clear, however, that the geometric term is no
small correction.
For pedagogical purposes we consider the case of a hydrogen atom, hence we
identify the electronic degrees of freedom [x], used in Sec. 4.2.2, with a single
coordinate r, and the parameter with the nuclear coordinate R. If the atom
is subject only to a magnetic fieldand neglecting irrelevant spin-dependent
termsthe complete (nonrelativistic) Hamiltonian H and the electronic Hamil-
tonian H are i2
1 h e
H(R, r) = p A(R) + H(R, r); (4.30)
2M c
1 h e i 2 e2
H(R, r) = i~r + A(r) . (4.31)
2m c |r R|
As explained above, the nuclear kinematical momentum of Eq. (4.15) becomes
e
= p ~A(R) A(R) (4.32)
c
The case of a constant B field can be dealt with analytically. We choose the
central gauge A(r) = 21 B r. If (r) is the exact ground eigenfunction when
the proton sits at R = 0, the eigenfunction at a generic R is:
ie
hr|(R)i = e 2~c rBR (r R), (4.33)
with an R-independent eigenvalue. The Berry connection is clearly
e e
A(R) = ih(R)|R (R)i = h(R)|B r|(R)i = BR
2~c 2~c
e
= A(R), (4.34)
~c

42
since the R-derivative of (r R) does not contribute. Replacing Eq. (4.34)
into Eq. (4.32) we find = p, as it must be: the nucleus travels at constant
speed, and is not deflected by a Lorentz force.
Remarkably, the magnetic-like field due to the Berry phase isin this sim-
ple exampleexactly opposite to the external magnetic field, thus providing the
complete screening which is physically expected. In less trivial situations, the
screening affects significantly the molecular vibrations and the classical nuclear
motion in general. The case of H2 has been investigated in 2007 by Ceresoli et
al. [70].

4.4 Anomalous Hall effect


Edwin H. Hall discovered the eponymous effect, nowadays in all solid state
textbooks, in 1879. Shortly later, in 1881, he discovered that the effect in
ferromagnetic materials is anomalous. While in nonmagnetic materials the
Hall resistivity is linear in the magnetic field, in ferromagnetic ones it saturates
to a value roughly proportional to the macroscopic magnetization.
In 1954 Karplus and Luttinger [71] provided a theory of the anomalous Hall
effect (AHE) which, with hindsight, was indeed geometrical. The understanding
of AHE, however, remained controversial for another 40 years and more. One of
the reasons for this state of affairs is that the intrinsic geometrical contribution
is partly obscured by extrinsic contributions (jargon: skew scattering and side
jumps), not easily disentangled in the experimental data. A comprehensive
review appeared in 2010 [12]. Here we only outline the intrinsic theory, based
on geometrical concepts, which owes to a couple of papers appeared in 2002, by
Jungwith, Niu, and MacDonald [72], and by Onoda and Nagaosa [73].
The starting point is the Kubo formula of Eq. (4.21), which we rewrite in
dimension d as
X 1 Z hujk |vx |uj k ihuj k |vy |ujk i
xy = 2~e2 Im d
dk . (4.35)

(2) jk <F <j k (jk j k )2
jj

While in Sec. 4.2.4 this was integrated in the BZ, for a metal the k integral is
limited to the volume inside the Fermi surface. Nonetheless, the transformation
of the integrand, from a sum over states into a curvature, proceeds in the same
way. The geometrical contribution to the AHE is therefore proportional to the
integral of the Berry curvature within the Fermi surface. To allow for band
crossings, we write the intrinsic AHE conductivity in dimension d as
Z
e2 1 X
xy = 2 Im dk hx ujk |y ujk i; (4.36)
~ (2)d j=1 jk <F

the analogy to Eqs. (4.22) and (4.23) is self evident. But this analogy is partly
misleading: at variance with the quantum Hall case, no macroscopic B field is
present here, and the u orbitals in Eq. (4.36) are the (periodic parts of) genuine
Bloch orbitals. It is expedient to rewrite the 2d and 3d cases in the form:
Z
e2 1 X
xy = dk j,xy (k) 2d; (4.37)
h 2 j=1 jk <F

43
Z
e2 1 X
xy = dk j,xy (k) 3d. (4.38)
h (2)2 j=1 jk <F

In the 2d case we have expressed the Hall conductivity in klitzing1 times a


dimensionless number, which becomes an integer (Chern number) for an insu-
lating system. In 3d we get a more compact expression using the vector form
of the Berry curvature:
Z
e2 1 X
= K , K= dk j (k). (4.39)
2h 2 j=1 jk <F

The vector K, sometimes called the Chern vector, has the dimension of an
inverse length; in the insulating case K is quantized to a reciprocal lattice vector,
and takes the name of Chern invariant (Sec. 7.1).
Given that the Fermi surface is symmetrical under k k, the symmetry
considerations of Sec. 3.7 show that Eq. (4.36) can be nonzero only if time-
reversal symmetry is broken, while inversion symmetry is irrelevant. The typical
case studies are the ferromagnetic metals, whose ground state breaks indeed
time-reversal symmetry in absence of a macroscopic B field.
First-principle calculations were performed for Ni, Cu, and Fe, as well as
or for some oxides. The intrinsic geometric contribution appears to be the
dominant one. These calculations also pointed out the crucial role played by
avoided crossings of the bands near the Fermi surface, which induce a very spiky
behavior of the Berry curvature in the BZ. More than 106 k points where used
in Ref. [74] in order to perform the integration in Eq. (4.36); a more efficient
strategy was devised later [75].
A noninteracting (e.g. KS) many-electron system is a trivial example of
a Fermi liquid. Haldane [76] pointed out that the very basic tenet of Lan-
daus Fermi-liquid theory is that charge transport involves only quasiparticles
with energies within kB T of the Fermi level. This is apparently at odds with
Eq. (4.36), which is an integration over the whole occupied Fermi sea. The two
viewpoints can be reconciled, essentially via an integration by parts [76]. Even
this alternative form has been implemented in first-principle calculations [77].

4.5 Semiclassical transport


The semiclassical theory of Bloch electron dynamics plays a fundamental role in
the physics of metals and semiconductors, and is a typical textbook topic [78].
At the classical level, particle transport is considered to take place in phase
space, whose points are labeled by r and p, and is described by the distribution
function f (r, p) which satisfy the Boltzmann equations. Electrons in crystalline
materials may fit into this framework if we identify the particles a wave packets
constructed from the Bloch orbitals. The semiclassical theory addresses there-
fore the motion of a wave packet built as a superposition of Bloch states from
the n-th band Z
|W i = dk a(k, t)|nk i, (4.40)
BZ
where the envelope function is well localized in k-space. Because of this, it
is delocalized in r space; we assume, however, that its center of mass is well

44
defined. Owing to this, we may define the wave vector k and the center r of the
wave packet as
Z
k = dk k |a(k , t)|2 ; r = hW |r|W i. (4.41)
BZ

4.5.1 Textbook equations of motion


In absence of collisions, the equations of motion (EoM) reported in textbooks
and routinely used in device engineering are
1 k
r =
~ k
1
~k = e (E + r B), (4.42)
c
where k is the band structure of the relevant band, and E and B are the
perturbing fields, assumed weak and slowly varying in space and time.
As emphasized in Ref. [78], the derivation of Eq. (4.42), despite the formal
simplicity of the result, is a formidable task. The early derivations date from
the 1930s; the problem was reconsidered several times in the literature, by Slater
[79], Luttinger [80], and Zak [81] among others.

4.5.2 Modern equations of motion


The modern analysis of semiclassical transport owes to a couple of papers by
Q. Niu and coworkers [63, 64] (see also Ref. [13]). They start observing that a
wave packet generally possesses two kinds of motion: the center-of-mass motion
and the self-rotation around its center. Owing to the latter, the wave packet is
endowed with an orbital magnetic moment, whose expression is
ie
m(k) = hk uk | (Hk k ) |k uk i. (4.43)
2~c
Because of this, the band structure acquires a B-dependent term

k k = k m(k) B; (4.44)

Furthermore, the canonical momentum ~k has to be replaced with the kinetic


momentum, which includes the (geometrical) vector potential. In the Newton-
like EoM, its contribution is reminiscent of a Lorentz force in reciprocal space.
Eq. (4.42) must then be replaced with

1 k
r = k (k)
~ k
1
~k = e (E + r B), (4.45)
c
where (k) is the Berry curvature of the relevant band, having the dimensions
of a squared length. Notice that the curvature is nonzero even in presence of
time-reversal symmetry, provided that the crystal is noncentrosymmetric.

45
4.5.3 Equations of motion in symplectic form
A more accurate analysis performed in 2003 by Panati, Spohn, and Teufel [82]
proves that Eq. (4.45) is not yet the ultimate form for the semiclassical EoM;
it is only correct if B is constant in space. Furthermore, the equations which
supersede Eq. (4.45) are more symmetrical and elegant.
One starts defining

H (r, k) = k e(r) m(k) B(r), (4.46)

where (r) is the potential of the electric field E(r), and m(k) is the wave
packets moment, Eq. (4.43). Then the EoM are

1 H (r, k)
r = k (k)
~ k
1 H (r, k) e
k = r B(r), . (4.47)
~ r ~c
In order to realize the relationship of Eq. (4.47) to the standard symplectic form
of Hamiltons equations, it is expedient to adopt atomic units, and furthermore
to redefine B by including the 1/c factor: we get

r = p H (r, p) p (p)
p = r H (r, p) r B(r). (4.48)

Notice that the translational momentum is coupled to B via the explicit Lorentz
force r B(r), unlike in the standard Hamiltonian formulation where such
coupling would appear via a vector-potential-dependent H . It is easy to prove
that Eq. (4.48) conserves the energy:

dH
= r H r + p H p = 0 (4.49)
dt
In the simple case where and B are identically vanishing, H becomes the
standard classical Hamiltonian and Eq. (4.48) are the corresponding Hamiltons
equations. If I is the 3 3 identity, the standard 6 6 symplectic matrix is
 
0 I
= , (4.50)
I 0

and the symplectic form of Eq. (4.48) is


   
r r H
= . (4.51)
p p H

Figure 4.5: A wave packet car-


ries quasi momentum ~k, but
also an orbital moment and
a corresponding magnetic mo-
ment m(k)

46
When either B or (or both) are nonzero, one recasts identically Eq. (4.48) in
the similar form    
r r H
= , (4.52)
p p H
where is a deformed symplectic matrix defined as
~
 
B(r) I
= ~ , (4.53)
I (p)
~
and the 3 3 antisymmetric matrices B(r) ~
and (p) are

0 Bz By 0 z y
B~ = Bz 0 Bx , ~ = z 0 x . (4.54)
By Bx 0 y x 0
~ and the
Finally, we notice that if we set to zero both the Berry curvature
moment m(k) in Eq. (4.46), then Eq. (4.52) provides an alternative form for
Hamiltons equations, where the B field appears in the symplectic matrix and
no vector potential appears in the Hamiltonian H . This form is well known in
classical mechanics, and goes under the name of gauge-invariant Hamiltonian
formulation.

4.5.4 Geometrical correction to the density of states


In a remarkable 2005 paper by Xiao, Shi, and Niu [83] it was pointed out
that Eq. (4.45), in presence of a nonzero B field, violates Liouvilles theorem.
This means that the volume element V = rp changes in time during the
evolution of the system; it is possible, however to remedy this shortcoming in
an elegant way.
We are not following the original paper; instead, we start from the more
elegant symplectic formulation above. The time evolution of the volume element
is
1 V
= r r + p p
V t
= r [ p H p (p) ] + p [ r H r B(r) ]
= r [ p (p) ] p [ r B(r) ]. (4.55)
After a somewhat lengthy calculation, one arrives at a very simple result,
whichmost notablyis a total derivative
1 V d
= log [ 1 + B(r) (p) ], (4.56)
V t dt
and is clearly nonzero whenever B and are non constant in phase space. It is
worth mentioning that Eq. (4.56) bears a simple relationship to the deformed
symplectic matrix, Eq. (4.53): in fact
q
1 + B(r) (p) = det (r, p). (4.57)

In order to arrive at Eqs. (4.56) and (4.57) two special features of the matrix
(r, p) seem to be essential: its elements are either functions of r or p, but not
of both; the divergences of B(r) and (p) are both zero.

47
Within ordinary statistical mechanics, the density of states in phase space is
1/hd (in dimension d). The number of states in volume V = rp is V /h3 ,
andowing to Liouvilles theoremthis number remains constant during time
evolution. Here instead, such number remains constant only if we appropriately
modify the density of states. For the sake of clarity we restore Gauss units; the
modification needed is
1 1 2
d [ 1 + B(r) (p/~) ], (4.58)
hd h 0
where 0 is the flux quantum.

4.5.5 Outstanding consequences of the modified density


of states
After Ref. [83] appeared, it become immediately clear that the geometrical
correction to the density of states, Eq. (4.58), goes well beyond the scope of
semiclassical approximation. To my present taste, a convincing proof of the
universality of Eq. (4.58) has not yet appeared; nonetheless its consequences are
unavoidable. Statistical mechanics paraphernalia, such as partition functions
and the like, must adopt Eq. (4.58).
As a consequence of the modified density of states, the Fermi volume of a
metal changes when a macroscopic B field is switched on at constant electron
density. If instead we keep the chemical potential constant, then the electron
density n depends on B. At zero temperature (for each spin channel)
 
1 2
Z
n = dk 1 + B (k) ( k )
(2)d BZ 0
 
n 1
Z
= dk (k) ( k ) (4.59)
B (2)d1 0 BZ

The latter assumes a perspicuous meaning in 2d, when is in a gap:


 
n 1 C1 1
Z
= dk (k) = = xy . (4.60)
B 20 BZ 0 ec

For a quantum Hall system, this goes under the name of Streda formula [84],
and had been first derived in 1982 in a very different way.

4.6 Quantum transport


4.6.1 Transport by a single state
We are going to study here the current induced by an adiabatic change of the
potential, or more generally of the Hamiltonian, in the single-particle case. We
indicate as n (t) the adiabatic instantaneous eigenstates, and with (t) the time
evolution of the ground state. In order to get rid of the dynamical phase, it is
better to deal with density matrices

(t) = |(t)ih(t)| = |0 (t)ih0 (t)| + (t). (4.61)

48
The velocity of this state is

v(t) = Tr {(t) v} (4.62)


X
= h0 (t)| v |0 (t)i + h0 (t)| (t) |n (t)ihn (t)| v |0 (t)i.
n

Since the adiabatic density matrix commutes with H(t), the time evolution
is
d
[H(t), (t)] = i~(t) i~ |0 (t)ih0 (t)|
dt
= i~ ( |0 (t)ih0 (t)| + |0 (t)ih0 (t)| ), (4.63)

where we are neglecting a term of higher order in the adiabaticity parameter.


We now take the matrix elements between h0 (t)| and |n (t)i:

(E0 En )h0 (t)| (t) |n (t)i = i~ ( hn (t)|0 (t)i + h0 (t)|n (t)i ). (4.64)

The term with n = 0 in the rhs vanishes because of norm conservation; replace-
ment into Eq. (4.62) yields

v(t) = h0 (t)| v |0 (t)i


" #
X h0 (t)|n (t)ihn (t)| v |0 (t)i
+i~ c.c. . (4.65)
E0 En
n6=0

The first term on the rhs is zero in the special case whereat all timesthe
Hamiltonian H(t) is time-reversal invariant and the state is nondegenerate.

4.6.2 Current carried by filled bands


We now exploit the previous result for a system of noninteracting electrons in
the case where the Hamiltonian H(t) is lattice periodical and the ground state
is insulating; this means that the gap remains finite at all t. It will be enough
to consider the simple case of just one filled band, with band index zero; the
current carried by each Bloch orbital |0k i = eikr |u0k i is
" #
X h0k |jk ihjk | v |0k i
vk (t) = h0k | v |0k i + i~ c.c.
0k jk
j6=0
X  hu0k |ujk ihujk | v |u0k i 
= hu0k | v |u0k i + i~ c.c. , (4.66)
0k jk
j6=0

where the t dependence of the rhs is now implicit. We then adopt the usual
formula for the velocity, Eqs. (1.18) and (4.20), and the analog of Eq. (3.27):

1 1X hujk | v |u0k i
v= k Hk , |k u0k i = |ujk i (4.67)
~ ~ 0k jk
j6=0

1
vk (t) = k 0k + i ( hu0k |k u0k i hk u0k |u0k i ). (4.68)
~

49
The first term on the rhs integrates to zero over the BZ, while the second is
clearly a Berry curvature component in the four-dimensional k, t domain. The
current density carried by a filled band in dimension d is
ie
Z
j(t) = dk ( hu0k |k u0k i hk u0k |u0k i ); (4.69)
(2)d BZ
the Bloch states are normalized to one in the crystal cell (as everywhere in the
present Notes).

4.6.3 Quantization of charge transport


Let us consider the special case of a simple cubic crystal with lattice constant
a. The transported charge in time T in the z direction across one cell is
Z T Z T Z /a Z /a
2 ea2
Q = dt Iz (t) = a dt jz (t) = dkx dky
0 0 (2)2 /a /a
Z T Z /a
i
dt dkz ( hu0k |z u0k i hz u0k |u0k i ). (4.70)
2 0 /a

If the time evolution of the Hamiltonian is cyclic H(T ) = H(0), then the second
line in Eq. (4.70) is clearly a Chern number C (in the kz , t variables) and is
integer. Notice also that C is dimensionless, and therefore does not depend on
how fast the Hamiltonian varies with time; ideally, the adiabatic regime means
T .
We arrive therefore at the outstanding result
Z T
Q= dt Iz (t) = e integer (4.71)
0

first proved by Thouless in 1983 [85]; it holds of course for any dimension d.
Let me restate the theorem: if the Hamiltonian is changed adiabatically in such
a way that it returns to its starting value in time T , the transported charge
in an infinite periodic system is quantized provided that the system remains
insulating at all times. A cycle pumps an integer number of elementary charges
across the system.
Among the examples which realize a Thouless pump, the original paper
suggests a sliding charge-density wave. A more outstanding manifestation of
quantized charge transport was pointed out shortly afterwards by Pendry and
Hodges [86]: Faradays laws of of electrolysis (1832). The mass/charge transfer
ratio shows that charge is always transported in units of e per ion, to the extent
that electrolytic cells are used as standards of current. If a given ion sits at one
electrode at t = 0, and if it drifts to the other one at t = T , the Hamiltonian
can be considered as cyclic, whence charge quantization follows. However, at
intermediate t values the charge belonging to a given ion is definitely non
quantized, and arbitrarily defined: for a review of the possible definitions, see
Refs. [87, 88].
Thouless quantization of charge transport [85], discussed above, also has
profound relationships to later advances: namely, to the topological explanation
of the quantization of surface charge (discussed in Sec. 5.3.4), and to the modern
theory of polarization (discussed in Sec. 5.3).

50
Chapter 5

Macroscopic polarization

The macroscopic polarization P is a fundamental concept that all undergradu-


ates learn about in elementary courses [89, 90]. In view of this, it is truly ex-
traordinary that until the early 1990s there was no generally accepted formula
for P in condensed matter, even as a matter of principles. P is an intensive
vector quantities that intuitively carries the meaning of dipole per unit volume.
Most textbooks [78, 91] provide a flawed definition of P, not implementable in
practical computations [92].
A genuine change of paradigm was initiated by a couple of important pa-
pers [93, 94], after which the major development was introduced by King-Smith
and Vanderbilt in 1992 (paper published in 1993 [95]). Other important ad-
vances occurred during the 1990s [96, 97] and the so-called modern theory of
polarizationit is at a mature stage since more than a decade; several reviews
have appeared in the literature over the years [1, 2, 3, 5, 6, 7, 8, 20].
Aiming at a computational physics readership, it is worth emphasizing that
most ab-initio electronic-structure codes on the market, for dealing with ei-
ther crystalline or noncrystalline materials, implement the modern theory of
polarization as a standard option. A nonexaustive list includes abinit [59],
crystal [60], quantum-espresso [62], siesta [98], vasp [61], and cpmd [99];
the code tutorials often include an outline of polarization theory. Its imple-
mentations have been instrumental since more than a decade in the study of
ferroelectric and piezoelectric materials [100, 101, 102].
The basic concepts of the modern theory of polarization also start reaching
a few textbooks [103, 104], though very slowly; most of them are still plagued
with erroneous concepts and statements.

5.1 Polarization and electric field


The modern theory of polarization, at least in its original form, only addresses
the polarization P in a null macroscopic E field; in this case P can be nonzero
only if the medium breaks inversion symmetry. This applies to noncentrosym-
metric crystals, and more generally to cases where inversion symmetry is broken
by some perturbation (as e.g. a frozen phonon, or piezoelectric strain).
It must be realized that, insofar as we address an infinite system with no
boundaries, the E field is quite arbitrary. The microscopic charge density is
neutral in average and lattice periodical; the value of E is just an arbitrary

51
z

+ + + + + +

Figure 5.1: Macroscopic polarization P in a slab normal to z, for a vanishing


external field E(ext) . Left: When P is normal to the slab, a depolarizing field
E = 4P is present inside the slab, and charges at its surface, with areal
density surface = P n Right: When P is parallel to the slab, no depolarizing
field and no surface charge is present.

boundary condition for the integration of Poissons equation. The usual choice
(performed within all electronic-structure codes) is to impose a lattice-periodical
Coulomb potential, i.e. E = 0. Imposing a given nonzero value of E is equally
legitimate (in insulators), although technically more difficult [105, 106]).
When addressing a finite sample with boundaries, the E field is in princi-
ple measurable inside the material, without reference to what happens at the
sample boundary; this is not the case of D. In fact, E obtains by averaging
over a macroscopic length scale the microscopic electric field E(micro) (r), which
fluctuates at the atomic scale [90]. In a macroscopically homogeneous system
the macroscopic field E is constant, and in crystalline materials it coincides with
the cell average of E(micro) (r). A lattice-periodical potential enforces E = 0; for
a supercell calculation, this applies to the field average over the supercell, while
in different regions there can be a nonzero macroscopic field.
As explained so far, there is no need of addressing finite samples and external
vs. internal fields from a theoreticians viewpoint. Nonetheless a brief digression
is in order, given that experiments are performed over finite samples, often in
external fields. Suppose a finite macroscopic sample is inserted in a constant
external field E(ext) : the microscopic field E(micro) (r) coincides with E(ext) far
away from the sample, while it is different inside because of screening effects.
If we choose an homogeneous sample of ellipsoidal shape, then the macroscopic
average of E(micro)(r), i.e. the macroscopic screened field E, is constant in
the bulk of the sample. The shape effects are embedded in the depolarization
coefficients [89]: the simplest case is the extremely oblate ellipsoid, i.e. a slab of
a macroscopically homogeneous dielectric; more details are given in Ref. [8]. For
the slab geometry in a vanishing external field E(ext) the internal field E vanishes
when P is parallel to the slab (transverse polarization), while E = 4P is the
depolarization field when P is normal to the slab (longitudinal polarization):
see Fig. 5.1.

5.2 Polarization itself vs. polarization differ-


ence
Novel ideas about macroscopic polarization emerged in the early 1990s [93, 94];
these led to the modern theory, based on a Berry phase, which was founded
by King-Smith and Vanderbilt soon afterwards [95]. At its foundation, the

52
Figure 5.2: Top panel: The 14-
atom BeO supercell in a ver-
tical plane through the BeO
bonds; the wurtzite (W) and
zincblende (ZB) stackings are
perspicuous. Bottom panel:
Macroscopic averages of the va-
lence electron density (solid)
and of the electrostatic poten-
tial (dotted).

modern theory was limited to a crystalline system in an independent-electron


framework (either KS or Hartree-Fock). Later, the theory was extended to
correlated and/or disordered systems [96, 97].
The first calculation ever of spontaneous polarization was published in
1990 [93]. The case study was BeO: it has the simplest structure where in-
version symmetry is absent (i.e. wurtzite), and furthermore its constituents
are first-row atoms. The idea was to address the macroscopic polarization of
a slab of finite thickness, with faces normal to the c axis, embedding it in an
ad hoc medium which (i) has no bulk polarization for symmetry reasons, and
(2) does not produce any geometrical or chemical perturbation at the interface.
The optimal choice is a fictitious BeO in the zincblende structure. Because of
obvious reasons, the system is periodically replicated in a supercell geometry
(Fig. 5.2, top panel). The selfconsistent calculation shows well localized interface
charges, of opposite sign and equal magnitudes at the two nonequivalent inter-
faces (Fig. 5.2, bottom panel). The interface charge is related to the difference
in polarization between the two materials: interface = P n. The computer
experiment provides the value of interface , and since P vanishes by symmetry
in the zincblende slab, one thus obtains the bulk value of P in the wurtzite
material. Notice that here P is a longitudinal polarization, in a depolarizing
field.
It must be emphasized that the quantity really measured in this computer
experiment is P, not the polarization P itself. After Ref [93] was published,
a study of the experimental literature showed thatcontrary to an incorrect
widespread beliefno experimental value of P in any wurtzite material ex-
ists: only estimates are available. Ref. [93] marks, as said above, a change of
paradigm: polarization must be defined by means of differences, and the con-
cept of polarization itself must be abandoned. With hindsight, it is nowadays
pretty clear that the problem, as well as its solution, exists already at the clas-
sical level: this is sketched in Fig. 5.3. Most textbooks are missing this very
basic fact.
The modern theory avoids addressing the absolute polarization of a given
equilibrium state, quite in agreement with the experiments, which invariably
measure polarization differences. Instead, the theory addresses differences in
polarization between two states of the material that can be connected by an

53
Figure 5.3: A 1d solid with infinite
length. Different choices of the unit cell
give different P values: (a), (b). On the
other hand, the change of polarization
P does not depend on the choice of
the unit cell (c).

adiabatic switching process. The time-dependent Hamiltonian is assumed to


remain insulating at all times, and the polarization difference is then defined
[94] as the time-integrated transient macroscopic current that flows through the
insulating sample during the switching process:
Z t
P = P(t) P(0) = dt j(t). (5.1)
0

In the adiabatic limit t and j(t) 0, while P stays finite. Ad-


dressing currents (instead of charges) explains the occurrence of phases of the
wavefunctions (instead of square moduli) in the modern theory. Eventually the
time integration in Eq. (5.1) will be eliminated, leading to a two-point formula
involving only the initial and final states.

5.3 Independent electrons


5.3.1 The King-Smith and Vanderbilt formula
For a crystalline system of independent electrons the expression for the tran-
sient current occurring in Eq. (5.1) is precisely the same as previously derived
for quantum transport, Eq. (4.69). Therefore for one band (index 0) and sin-
gle occupancy in dimension d the electronic contribution to the polarization
difference is
Z t Z
ie
P = dt dk ( hu0k |k u0k i hk u0k |u0k i ); (5.2)
(2)d 0 BZ

the (classical) nuclear contribution must be added separately. We remind that


crystal-cell neutrality is essential. Notice also that, given the occurrence of
Bloch states, the Hamiltonian is lattice periodical at all t: this implicitly means
that in Eq. (5.2) P is evaluated at E = 0.
It is now expedient to introduce a dimensionless adiabatic time , with
|ujk (t)i = |ujk ((t))i, (0) = 0, (1) = t. Eq. (5.2) becomes then, for n
doubly occupied bands (index j = 1, n) in 3d:
Z 1
P = P(1) P(0) = d P()
0
n Z
2ie X
P = dk ( h ujk |k ujk i hk ujk | ujk i ). (5.3)
(2)3 j=1 BZ

54
It is essential that the gap does not close, i.e. the system remain insulating, for
all values.
The expression in Eq. (5.3) can be integrated with respect to to obtain
n Z
2ie X
P() = dk hujk |k ujk i : (5.4)
(2)3 j=1 BZ

this is the (by now famous) King-Smith and Vanderbilt formula [95], yielding
the polarization of the final state minus the polarization of the initial state,
Eq. (5.3). To understand the meaning of the k integral in 3d we take the simple
example of a simple cubic lattice of constant a, similarly to Eq. (4.70):
n Z
" Z #
2e X /a
Z /a /a
Pz () = dkx dky i dkz hujk |kz ujk i , (5.5)
(2)3 j=1 /a /a /a

where the square parenthesis highlights the Berry-phase, to be compared to


Eq. (3.43).
In first-principle implementations, the Berry phase is discretized as in
Eqs. (3.45) and (3.46), and the remaining 2d k-integral is discretized in the
trivial way. The first calculation ever of the spontaneous polarization of a
ferroelectric material (KNbO3 ) appeared in 1993 [107], and agreed within 10%
with the measured values. As said above, the Berry-phase formula is nowadays
implemented in most first-principle codes.

5.3.2 The quantum of polarization


Given that every phase is defined modulo 2, all of the two-point formulas for
P in terms of Berry phases are arbitrary modulo a polarization quantum.
This is the tradeoff one has to pay when switching from the curvature formula,
Eq. (5.3)where no such arbitrariness existsto the two-point King-Smith and
Vanderbilt formula, where only the connection occurs. The actual arbitrariness
of P in 3d is 2eR/Vcell, where R is a lattice vector and Vcell is the cell volume
(the 2 factor owes to double band occupancy). A similar arbitrariness of an
integer times eR/Vcell occurs for the classical nuclear contribution to polariza-
tion.
The quantum arbitrariness is rarely a problem in practice. In most cases, the
change in P that can be induced by a perturbation, such as a small sublattice
displacement, is insufficient to cause P to change by a large fraction of the
quantum. Where exceptions exist the ambiguity can be resolved by subdividing
the adiabatic path into several shorter intervals, for each of which the change
in P is unambiguous for practical purposes. Additional problems may occur in
the discretized version of the Berry-phase formula; this is discussed e.g. in Ref.
[8] and, in more detail, in Ref. [7].
Here we stress that the quantum ambiguity is an essential aspect of the
theory. For example, for the case of a closed cyclic adiabatic evolution of the
system, in which the parameter values = 0 and = 1 label the same physical
state of the system, we retrieve the quantization of charge transport, discussed
in Sec. 4.6, and governed by a Chern number.

55
5.3.3 Wannier functions
The KS (or Hartee-Fock) ground state is a Slater determinant of doubly occupied
orbitals; any unitary transformation of the occupied states among themselves
leaves the determinantal wavefunction invariant (apart for an irrelevant phase
factor), and hence it leaves invariant any KS ground-state property.
For an insulating crystal the Bloch KS orbitals of completely occupied bands
can be transformed to localized Wannier orbitals (or functions) WFs. This is
known since 1937 [108], but for many years the WFs have been mostly used as a
formal tool; they became a popular topic in computational electronic structure
only after the 1997 seminal work of Marzari and Vanderbilt [109]. A compre-
hensive review appeared as Ref. [19], and a public-domain implementation is in
wannier90 [110]. If the crystal is metallic, the WFs can still be technically
useful [111], but it must be emphasized that the ground state cannot be writ-
ten as a Slater determinant of localized orbitals of any kind, as a matter of
principle [112].
The transformation of the Berry phase formula in terms of WFs provides an
alternative, and perhaps more intuitive, viewpoint. The formal transformation
was known since the 1950s [113], although the physical meaning of the formalism
was not understood until the advent of the modern theory of polarization.
The unitary transformation which defines the WF wjR (r), labeled by band
j and unit cell R, within our normalization is
Vcell
Z
|wjR i = dk eikR |jk i . (5.6)
(2)3 BZ

If one then defines the Wannier centers as rjR = hwjR |r|wjR i, it is rather
straightforward to prove that Eq. (5.4) is equivalent to
n
2e X
P(el) () = rj0 . (5.7)
Vcell j=1

This means that the electronic term in the macroscopic polarization P is (twice)
the dipole of the Wannier charge distributions in the central cell, divided by the
cell volume. The nuclear term is obviously similar in form to Eq. (5.7); the sum
of both terms is charge neutral.
WFs are severely gauge-dependent, since the phases of the |jk i appearing in
Eq. (5.6) can be chosen arbitrarily. However, their centers are gauge-invariant
modulo a lattice vector. Therefore P(el) in Eq. (5.7) is affected by the same
quantum indeterminacy discussed above. We also stress, once more, that
Eq. (5.7)as well as Eq. (5.3)is to be used in polarization differences, and
does not define polarization itself.
The modern theory, when formulated in terms of WFs, becomes much
more intuitive, and in a sense vindicates the venerable Clausius-Mossotti view-
point [114]: in fact, the charge distribution is partitioned into localized contri-
butions, each providing an electric dipole, and these dipoles yield the electronic
term in P. However, it is clear from Eq. (5.6) that the phase of the Bloch or-
bitals is essential to arrive at the right partitioning. Any decomposition based
on charge only is severely nonunique and does not provide in general the right
P, with the notable exception of the extreme case of molecular crystals.

56
In the latter case, in fact, we may consider the set of WFs centered on a
given molecule; their total charge distribution coincidesin the weakly inter-
acting limitwith the electron density of the isolated molecule (possibly in a
local field). This justifies the elementary Clausius-Mossotti viewpoint. It is
worth mentioning that the dipole of a polar molecule is routinely computed in a
supercell geometry via the single-point Berry phase discussed below [115]. The
dipole value coincides with the one computed in the trivial way in the large
supercell limit. Finite-size corrections, due to the local field (different in the
two cases), can also be applied [116].
The case of alkali halideswhere the model is often phenomenologically
useddeserves a different comment [8]. The electron densities of isolated
ions (with or without fields) are quite different from the corresponding WFs
charge distributions, for instance because of orthogonality constraints: hence
the Clausius-Mossotti model is not justified in its elementary form, despite con-
trary statements in the literature. For a detailed analysis, see Ref. [117].

5.3.4 The surface charge theorem


The early occurrences of the theorem of quantization of the surface charge [35,
36, 37, 38] are discussed above, Sec. 2.3. The topological nature of this theorem
was first realized by Niu [118] in 1986; here we follow the treatment of Vanderbilt
and King-Smith [119] (see also Ref. [5]).
According to elementary electrostatics the macroscopic bound surface charge
density surface residing on the surface of a sample is related to the polarization
in the interior by surface = n P, where n is the surface normal. One defines
the bound charge surface by saying that no free charge is present, but what,
precisely, does this mean? The surface must be insulating, with the electron
chemical potential lying in a gap that is common to both bulk and surface.
But this is not a unique prescription, since there can be a surface band which
is entirely occupied or entirely empty. The two cases differ by a polarization
quantum in the corresponding P value. In fact, given that the bulk polarization
P is arbitrary modulo eR/Vcell, it follows that the charge per surface area is
defined modulo a quantum
e
surface = n P modulo . (5.8)
Asurface
An equivalent formulation of the surface charge theorem can be arrived at by
means of WFs. The WF approach is most perspicuous for quasi-1d systems (e.g.
insulating polymers); a pedagogical presentation is in Ref. [120]. Notice that all
terminations are insulating in a polymer, there are no surface conducting
states.
The bulk-surface correspondence encoded in Eq. (5.8) is an outstanding man-
ifestation of topology in condensed matter physics: the surface charge of an in-
sulating surface is topologically protected. The actual value of surface among
the discrete allowed values is then determined by energy considerations.
For a centrosymmetric crystal it is tempting to guess that P() in the Berry-
phase formula, Eq. (5.4), vanishes for any . Instead, this is not the case:
centrosymmetry only dictates that P = P modulo a quantum, and therefore
Eq. (5.4) yields P equal to an integer or half integer multiple of the quantum
eR/Vcell (for single band occupancy). Then from Eq. (5.8) it follows that the

57
Figure 5.4: A cen-
trosymmetric insulating
quasi-1d crystal with
two different termina-
tion: alternant trans-
polyacetylene. Here the
bulk is five-monomer
long.

charge per surface cell may only be an integer or half integer, as first discov-
ered many years ago, and previously discussed in Sec. 2.3. Therein, it was
observed that this important theorem is often ignored even by specialists in sur-
face physics. A thorough analysis of polar surfaces, in the light of the present
theorem, has appeared in a couple of 2011 papers: by Stengel [121] and by
Bristowe, Littlewood and Artacho [122].
Among the values dictated by topology, Nature chooses the minimum-energy
one. If the electric field vanishes outside the solid, a charged surface implies a
nonvanishing field inside. This has an extensive energy cost (proportional to the
square of the field times the volume). If the bulk of the solid is centrosymmetric,
the surface charge is zero; or otherwise the surface is metallic.
Quasi-1d crystalline systems (i.e. stereoregular polymers) are more in-
teresting. Therein: (1) the energy cost of the field is nonextensive and (2) the
surface (i.e. the termination) cannot be metallic. Therefore different values
of the surface quantum can be actually realized for the same bulk: we illustrate
this with Figs. 5.4 and 5.5, taken from Ref. [120].
Alternant trans-polyacetylene is a centrosymmetric quasi-1d insulating crys-
tal: its end charges are quantized in units of half an election charge. We con-
sider two different terminations, shown in Fig. 5.4: notice that in both cases
the molecule as a whole is not centrosymmetric, although the bulk is. The end
charges are trivially related to the dipole per unit length (or per monomer).
Fig. 5.5 shows that, in the long system limit, the end charges are either zero
or one, depending on the termination. The structure is either neutral or
charge-transfer; the end groups are donor (NH2 ) and acceptor (COOH).
The figure also shows that at finite sizes both molecules are polar; the quanti-
zation is exact in principle in the infinite system limit; in the present case it is
attained at about 10-20 monomers. This length is essentially the exponential
decay length of the one-body density matrix of bulk polyacetylene.
In the polyacetylene case considered so far topology mandates the end
charges to be zero modulo 1 (in units of e). Other cases where instead the
end charge is 1/2 modulo 1 are also considered in Ref. [120]. The actual value
of the end charges depends on the relative ionicity of the end groups vs. the
bulk. Counterintuitively, while the ionicity varies on a continuous scale, the end
charges may only vary by an integer. More about this is said below, Sec. 5.5.

58
7
[+-]
[NN]
6

4
Figure 5.5: Quantization of the
3
end charges in polyacetylene, af-
2
ter Ref. [120]. Dipole per
1 monomer as a function of the
0 number of monomers in the
-1
chain, for the two different ter-
1 10 100
minations.

5.3.5 Noncrystalline systems: The single-point Berry


phase
The key ingredient for computing the infrared spectrum of amorphous or liquid
systems is the power spectrum of the autocorrelation function of the macro-
scopic polarization hP(t) P(0)i. Since Car-Parrinello simulations are custom-
arily performed using only k = 0 in the (supercell) BZ, we need to analyze the
single-point version of the Berry-phase formula for polarization, Eq. (5.5).
We consider a simple cubic supercell of side L; in the large L limit the BZ
integral of any function f (k) is approximated as

(2)3
Z
dk f (k) f (0). (5.9)
BZ L3
We start with the Berry phase, i.e. with the 1d integral in square parenthesis
in Eq. (5.5):
Z Z 2
L L
z = i dkz hujk |kz ujk i = i dkz hujk |kz ujk i Im log det S(k1 , k2 ),

L 0
(5.10)
where k1 = (0, 0, 0) and k2 = (0, 0, 2/L). In Eq. (5.10) we have used the
discretized Berry phase, Eq. (3.46), with only one factor in the matrix product.
2z
Then, as in Sec. 3.9, we notice that |ujk2 i = ei L |ujk1 i: therefore the overlap
matrix in Eq. (5.10) becomes
2z
Sjj (k1 , k2 ) = huj |ei L |uj i, (5.11)

where all the orbitals |uj i = |j i are evaluated at k = 0. We then approximate


even the remaining integrals in Eq. (5.5) with a single point. At any time during
the simulation the electronic term in the polarization is thus
e e
Pz(el) (t) = z = Im log det S. (5.12)
L2 L2
The nuclear (or core) contribution has a very simple form. If zm is the instan-
taneous z coordinate of the m-th nucleus, and eZm the corresponding charge,
the total polarization is
e e X
Pz (t) = 2
Im log det S + 3 Zm z m . (5.13)
L L m

59
This the expression currently used in computing power spectra and infrared
spectra [123], and, more generally, whenever a single k point is used in the first-
principle simulations [124, 115]. As already observed, even the nuclear term is
affected by the quantum indeterminacy: we can therefore express it equivalently
into an Im log form. It is easy to show that Eq. (5.13) can be equivalently
written as
e 2
P
Pz (t) = 2
Im log [ (det S)2 ei L m Zm zm ], (5.14)
2L
where the square of the determinant owes to the double spin occupancy.
The polarization quantum in Eq. (5.13) is e/L2 , which vanishes in the L
limit. This does not make any problem, and in fact Eq. (5.12) is routinely used
for evaluating polarization differences in noncrystalline materials. The key point
is that the L limit is not actually needed; for an accurate description of
a given material, it is enough to assume a finite L, actually larger than the
relevant correlation lengths in the material. For any given length, the quantum
e/L2 sets an upper limit to the magnitude of a polarization difference accessible
via the Berry phase. The larger are the correlation lengths, the smaller is the
accessible P. This is no problem at all in practice, either when evaluating
static derivatives by numerical differentiation, such as e.g. in Ref. [124, 115],
or when performing Car-Parrinello simulations [123]. In the latter case t is
a Car-Parrinello time step (a few a.u.), during which the polarization varies
by a tiny amount, much smaller than the quantum (the typical size of a large
simulation cell nowadays is L 50 bohr). Whenever needed, the drawback
may be overcome by splitting t into several smaller intervals. More about
computing infrared spectra via molecular dynamics simulations will be said in
Sec. D.10.

5.4 Correlated wavefunctions


The treatment given so far assumes an independent-particle scheme, where po-
larization is evaluated as a Berry phase of one-electron orbitals, typically the
KS ones. Shortly after the appearance of the King-Smith and Vanderbilt paper
[95], Ortiz and Martin [96] provided the many-body generalization of the theory,
where polarization is expressed as a Berry phase of the many-body wavefunc-
tion.
A subsequent development [97] provides a unified treatment of macroscopic
polarization, dealing on the same footing with either independent-electron or
correlated systems, and with either crystalline or disordered systems.

5.4.1 Single-point Berry phase again


Suppose we have N electrons in a cubic box of volume L3 , with a many-body
Hamiltonian
N
1 X
H = |pi |2 + V , (5.15)
2me i=1

and eigenfunctions |n i, normalized in the hypercube of volume L3N . In


Eq. (5.15) the potential V includes one-body and two-body (electron-electron)

60
contributions. As usual in condensed-matter theory, we adopt periodic Born-
von-Karman boundary conditions over each electron coordinate ri indepen-
dently, whose Cartesian components ri, are then equivalent to the angles
2ri, /L. The potential V enjoys the same periodicity, which implies that the
electric field averages to zero over the sample.
The formula for correlated wavefunctions is quite similar to Eq. (5.12), which
in fact is a special case of the latter. The starting ingredient is the single-point
Berry phase
2 2
z = Im log h0 |ei L
P
zi
|0 i = Im log h0 |ei L
P
zi
i i |0 i, (5.16)

and the polarization formula obtains by replacing z in Eq. (5.13), or equiva-


lently
2
P
(det S)2 Im log h0 |ei L i zi |0 i. (5.17)
e 2
P
Pz(el) = Im log h0 |ei L i zi |0 i. (5.18)
2L2
The proof of Eq. (5.18) is provided in the original paper [97], as well as in some
review papers [3, 6, 7, 8].
Here we content ourselves to prove that Eq. (5.18) coincides with Eq. (5.12)
in the special case where |0 i is the Slater determinant of the (doubly occu-
pied) k = 0 orbitals |uj i = |j i. The key observation is that the many-body
wavefunction
2
P
|0 i = ei L i zi |0 i (5.19)
2z
is the Slater determinant built from the orbitals |uj i = ei L |uj i, hence
2
h0 |ei L
P
zi
i |0 i = h0 |0 i = ( dethuj |uj i )2 = (det S)2 , (5.20)

where the second power owes to double orbital occupancy.


Finally we observe that, analogously to Eq. (5.14), the nuclear contribution
can be included into Eq. (5.18) as
e 2 (n) P
Im log h0 |ei L ( m Zm zm i zi ) |0 i,
P
Pz = 2
(5.21)
2L
(n)
whereby the sake of claritywe denote as zm the (classical) nuclear coordi-
nates, and with zi the (quantum) electronic ones.
Incidentally, the form of Eq. (5.21) naturally provides a generalization for
the case where even the nuclei are quantum particles: it is enough to interpret
|0 i as the full (electron and nuclear) ground-state wavefunction. I am not
aware of any similar result in the published literature.

5.4.2 Kohn-Sham polarization vs. real polarization


All of the independent-electron formulas discussed in Sec. 5.3 are exact for
noninteracting electrons, but the obvious aim is to implement them with KS or-
bitals, in a given density-functional theory (DFT) framework. Since macroscopic
polarization applies to insulators only, we stress that we mean KS insulator
throughout: that is, we assume that the KS spectrum is gapped. In the class of
simple (i.e. computationally friendly) materials a genuine insulator is also a
KS insulator, although pathological cases (computationally unfriendly) do exist.

61
Having specified this, the key issue is then: Does the KS polarization coincide
with the physical many-body one? The answer is subtle, and is different whether
one chooses either open boundary conditions, as appropriate for molecules
and clusters, or periodic boundary conditions (Born-von Karman), as invariably
done in the present Notes.
Within open boundary conditions the KS orbitals vanish at infinity, as well
as the charge density of the sample. P is then the first moment of the charge
density, divided by the sample volume. The basic tenet of DFT is that the
microscopic density of the fictitious noninteracting KS system coincides with the
density of the interacting system. Therefore the exact P coincides by definition
with the one obtained from the KS orbitals.
Matters are quite different within periodic boundary conditions: we have
seen above that P is not a function of the charge density, hence the value of
P obtained from the KS orbitals, in general, is not the correct many-body P.
This was first shown in 1995 by Gonze, Ghosez, and Godby [125], and later
discussed by several authors. A complete account of the issue can be found in
Ref. [6]. Here we just mention that the exact P is provided by Eq. (5.18), while
the KS P is provided by Eq. (5.12), where the KS orbitals enter Eq. (5.11); both
expressions are to be evaluated in the large-L limit. The two expressions are
clearly different whenever the ground wave function is not a Slater determinant.
Therefore P cannot be exactly expressed within DFT, but the exact func-
tional is obviously inaccessible, and even sometimes pathological. The practical
issue is whether the current popular functionals provide an accurate approxi-
mation to the experimental values of P in a large class of materials.
A vast first-principle literature has accumulated over the years by either
linear-response theory [126]not reviewed hereor by the modern theory. The
errors are typically of the order of 10-20% on permittivity, and much less on most
other properties (infrared spectra, piezoelectricity, ferroelectricity) for many
different materials. It is unclear which part of the error is to be attributed to
DFT per se, and which part is to be attributed to the approximations to DFT.

5.5 Polarization as a Z2 topological invariant


We consider in this Section only crystalline systems in 1d or quasi-1d (i.e. stere-
oregular polymers): polarization P has the dimensions of a charge, and therefore
P/e is dimensionless. We further limit ourselves to the centrosymmetric case.
We start rewriting the most compact single-point formula, in the 1d case,
Eq. (5.21):
1 2
P (n) P
P/e = Im log h0 |ei L ( m Zm xm i xi ) |0 i. (5.22)
2
(n)
For the sake of clarity, we have denoted as xm the nuclear coordinates and as xi
the electronic ones. We remind that P is well defined only for a charge-neutral
system, and indeed Eq. (5.22) includes both electronic and nuclear contributions
to P .
For a centrosymmetric system the matrix element in Eq. (5.22) is real, hence
its phase is either 0 or (mod 2). We also remind that, in general, Eq. (5.22)
is meaningful in the large-system limit: N electrons in a periodic box of length
L, with N/L constant. In a centrosymmetric system the matrix element is real
at any N : the limit is no longer necessary.

62
Figure 5.6: Two paradigmatic
centrosymmetric lattices in 1d

It is expedient to consider 2P/e, which therefore assumes the values of either


0 mod 2, or 1 mod 2. This establishes a Z2 classification of 1d centrosymmetric
materials: they are either Z2 -even (P = 0 mod e) or Z2 -odd (P = e/2 mod e).
We have not yet proved that such classification is indeed topological. Let
us consider the two simplest nonprimitive (binary) centrosymmetric lattices in
1d. The top sketch refers to an idealized molecular crystal, or to an alternant
polymer, like polyacetilene (compare to Fig. 5.4); the bottom sketch refers to
an idealized ionic crystal. The former is Z2 -even (see Fig. 5.5), while the latter
is Z2 -odd. Both system can be described at the simplest level by an extreme
tight-binding model Hamiltonian (Huckel-like), whose only parameters are the
first-neighbor hoppings t and the onsite energies . The Z2 -even model has
alternating ts and constant ; the Z2 -odd has constant t and alternating s. It
is a straightforward exercise to show that one cannot continuously deformate
the Hamiltonian (and its ground state) from one case into the other without
closing the gap, while conserving centrosymmetry. In modern jargon, the Z2
invariant is protected by centrosymmetry. Provided we dont close the gap,
we may also deformate the tight-binding ground state into a first-principle one:
the classification is independent of the theory level.
Since PBCs are at the root of Eq. (5.22), we have addressed so far an un-
bounded sample: the Z2 invariant is a bulk property, but is determined only
modulo a quantum. Real systems are instead bounded: polarization remains
ambiguous until the termination of the sample has been specified. In that case
Nature chooses the minimum-energy eigenstate, while respecting the topology
constraint (see Sec. 5.3.4). We also observe that for a bounded sample topol-
ogy only appears in the large-N limit (Fig. 5.5), while the PBC expression of
Eq. (5.22) is quantized even at finite N .
One further comment about the ionic case: bottom sketch in Fig. 5.6. If we
identify the structure with a lattice of classical point charges e, it is obvious
that the dipole per unit length of a bounded sample is e/2, depending on the
termination. When replacing the classical point charges with real anions and
cations, one expects the dipole per unit length to be about e/2, but only in the
limit of a strongly ionic system. Conterintuitively, topology mandates thatin
the large system limitthe dipole per unit length is e/2 even in the case of
very weak ionicity, i.e. a binary chain made of Ga and As atoms.
The proof that the end charge in centrosymmetric linear polymers is topo-
logical can be equivalently arrived at by Wannier-function counting: see Ref.
[120]. Similar arguments also prove the topological nature of the soliton charge
in polyacetilene, first discovered by Su, Schrieffer, and Heeger in 1979 [127]. If we
insist that we want a singlet wavefunctionas everywhere in this Sectionthen
the soliton charge can only be e. But we may relax this condition, adopting
single occupancy (instead of double) for one of the Wannier functions in the
soliton region: the soliton is then neutral, but carries spin 1/2.

63
Chapter 6

Quantum metric and the theory of the


insulating state

6.1 Nongeometrical theories of the insulating


state
The standard textbook approach to the insulating state of matter is based on
band theory. Following Blochs theorem [128] in 1928, the main result is due
to Wilson in 1931 [129]. The single-particle spectrum of a lattice-periodical
Hamiltonian is in general gapped, and the electron count determines where the
Fermi level lies. If it crosses a band one has a conductor: an applied electric field
induces free acceleration of the electrons (at T = 0 in absence of dissipation).
If the Fermi level lies instead in a gap, one has an insulator: in presence of
a field the electronic system polarizes, but no steady-state current flows for
T 0. This very successful theory explains the insulating/conducting behavior
of most common materials across the periodic table, for which band structure
calculations became soon available.
At the root of band theory are two basic assumptions: the electrons are
noninteracting (in a mean-field sense), and the solid is crystalline. By the early
1960s, however, it became clear that there are solids where these two assump-
tions are very far from the truth, and where the insulating behavior is due to
completely different mechanisms. The works of Mott in 1949 [130] and of An-
derson in 1958 [131] opened new avenues in condensed matter physics. In the
materials which we now call Mott insulators the insulating behavior is due to
electron correlation [132], while in those called Anderson insulators it is due to
lattice disorder [133].
In a milestone paper appeared in 1964 Kohn [134] provided a more compre-
hensive characterization of the insulating state of matter, which encompasses
band insulators, Mott insulators, Anderson insulators, and eventually any kind
of insulating material. According to Kohn, the electrons in the insulating state
satisfy a many-electron localization condition [135]. This kind of localization
must be defined in a subtle way given that, for instance, the Hamiltonian eigen-
states in a band insulator are obviously not localized. According to the original
Kohns formulation, the insulating behavior arises whenever the ground-state
wavefunction of an extended system breaks up into a sum of contributions which
are localized in essentially disconnected regions of the many-electron configura-

64
tion space.
Kohns theory remained little visited for many years [136] until the 1990s,
when a breakthrough occurred in electronic structure theory: the modern theory
of polarization. Inspired by the fact that electrical polarization discriminates
qualitatively between insulators and metals, Resta and Sorella [137] in 1999
provided a definition of many-electron localization rather different from Kohns,
deeply rooted in the theory of polarization, and therefore based on geometrical
concepts. Their program was completed soon after by Souza, Wilkens and
Martin [138] (hereafter quoted as SWM), thus providing the foundations of the
modern theory of the insulating state. An early review paper appeared in 2002
[4], and a more recent one in 2011 [9]; the latter is at the root of the present
Chapter.

6.2 Metric-curvature tensor


As in Sec. 5.4 we address an interacting N -electron system, whose most general
Hamiltonian we write, in the Schrodinger representation and in Gaussian units,
as
N
1 X e
H() = |pi + A(ri ) + ~|2 + V . (6.1)
2me i=1 c

Equation (6.1) is exact in the nonrelativistic, infinite-nuclear-mass limit. With


respect to Eq. (5.15), the velocity is augmented with two terms: A(r) is a vector
potential of magnetic origin, and , having the dimensions of an inverse length,
is a flux or twist of the same kind as the single-particle flux thoroughly
discussed in Sec. 1.6.
In Sec. 5.4 we adopted periodic Born-von-Karman boundary conditions
(PBCs) over a cubic supercell of side L, and the eigenfunctions |n i were nor-
malized in the hypercube of volume L3N . In this Chapter, instead, it is expedi-
ent to consider in parallel both open boundary conditions (OBCs) and PBCs.
The former are appropriate to molecular physics, and require that the many-
electron wavefunction of a bound state is square-integrable over the whole co-
ordinate space R3N . PBCs are instead appropriate for extended systems, either
crystalline or disordered, either independent-electron or correlated. It is worth
emphasizing that the choice of boundary conditions (either OBCs or PBCs)
corresponds to choosing the Hilbert space, which in turn affects profoundly the
geometrical properties discussed in Ch. 3.

6.2.1 Open boundary conditions


The case of OBCs is by far the simplest. We write for the sake of simplicity
the ground state of Eq. (6.1) at = 0 as |0 i |0 (0)i, and we define the
many-body position operator as
N
X
r = ri . (6.2)
i=1

As in Sec. 1.6.2 the flux is easily gauged away: the state eir |0 i coincides
with the ground eigenstate |0 ()i of the twisted Hamiltonian, Eq. (6.1). It

65
is legitimate to multiply this eigenvector by any -dependent (and position-
independent) phase factor; our choice is then

|0 ()i = ei(rd) |0 i, (6.3)

where d = h0 |r|0 i is the electronic dipole of the molecular system. It follows


that the -derivative needed in our metric-curvature tensor, Eq. (3.24) is

| 0 i = i(r d)|0 i = iQ(0) r|0 i, (6.4)

Exploiting the idempotency of the projector, the tensor F , evaluated at = 0,


is

F (0) = h0 |r Q(0)r |0 i
= h0 |r r |0 i h0 |r |0 ih0 |r |0 i, (6.5)

where Q is the projector over the unoccupied many-body eigenstates. F (0)


clearly a real symmetric tensor, hence the Berry curvature vanishes within
OBCs, and the metric g (0) coincides with F (0). It will be shown that
within PBCs, instead, the tensor F (0) may acquire a nonvanishing imaginary
part.
Clearly, the metric tensor in Eq. (6.5) is the cumulant second moment of the
position operator, or equivalently the ground state fluctuation of the dipole of
the molecular system; this quantity is extensive (scales like N in macroscopically
homogenous systems). We anticipate that F (0)/N discriminates, in the large
N -limit, between insulators and metals.

6.2.2 Periodic boundary conditions


First of all, a key observation about the position operator is in order. The simple
multiplicative operator r, as defined in Eq. (6.2), is forbidden within PBCs:
in fact it maps any periodic wavefunction |i into the nonperiodic function
r|i. In other words, it maps a function which belongs to the Hilbert space
to a function outside of it [97]. Incidentally, this is the main reason why the
polarization problem has remained unsolved until the early 1990s. In the present
context, formulae like Eq. (6.5) are ill defined and absurd within PBCs.
Within PBC the state eir |0 i is not an eigenstate of Eq. (6.1), except
for the discrete values
2
m1 m2 m3 = (m1 e1 + m2 e2 + m3 e3 ), (6.6)
L
where m Z, and e are the Cartesian versors. For fractional values of i.e.
for values different from those in Eq. (6.6)the -dependence of the eigenvalues
and eigenvectors of Eq. (6.1) is nontrivial. We have thoroughly discussed the
single-particle version of this feature in Sec. 1.6.3.
If |0 ()i is the genuine ground eigenstate of Eq. (6.1) within PBCs, then
the auxiliary function |0 ()i = eir |0 ()i is a solution of H(0), but ful-
fills quasi-periodic boundary conditions: at any two opposite faces of the cube
the wavefunction differs by a -dependent phase factor. In other words the
problem can be formulated in two equivalent ways: either the Hamiltonian is
-dependent, as in Eq. (6.1), and the boundary conditions are -independent;

66
or the Hamiltonian is -independent but the boundary conditions are twisted
in a -dependent way.
Within PBCs the tensor F (0) cannot be simplifiedas e.g. in Eq. (6.5)
and must be therefore addressed in its original form by actually evaluating the
derivatives:
F () = h 0 ()|Q()| 0 ()i. (6.7)
As within OBCs, even within PBCs this tensor is extensive. In the thermody-
namic limit (i.e N , keeping N/L3 constant), the well defined quantity is
F (0)/N . We also observe that for time-reversal symmetric systems F (0) is
real symmetric, and coincides therefore with the metric tensor g (0).

6.2.3 Sum over states again


We express hr r ic using the sum-over-states formula, Eq. (3.28) :

1 X h0 | H(0)|n ihn | H(0)|0 i


hr r ic = . (6.8)
N (E0 En )2
n6=0

The -derivative of the Hamiltonian of Eq. (6.1) is


N
~ Xh e i
H(0) = pi + A(ri ) , (6.9)
me i=1 c

where the rhs is nothing else than ~ times the velocity operator v; Eq. (6.8)
becomes then
1 X h0 |v |n ihn |v |0 i
hr r ic =
~2 N (E0 En )2
n6=0
1 X h0 |v |n ihn |v |0 i
= 2 , (6.10)
N 0n
n6=0

where 0n = (En E0 )/~.


The velocity operator is also commonly expressed as v = i[H(0), r]/~, but
it is worth emphasizing that while the position r is well defined within OBCs
and ill defined within PBCs, the velocity v is well defined in both cases.
The basic sum-over states formula, Eq. (6.10), applies therefore to both
OBCs and PBCs. In general, it is not very useful on practical grounds, since it
would require the evaluation of slowly convergent sums. Nonetheless, Eq. (6.10)
is instead essential to gather understanding into the physical meaning of F ,
as will be shown below.

6.2.4 Localization tensor (a.k.a. cumulant moment)


The modern formulation of the theory of the insulating state is based on a
localization tensor (squared localization length in 1d), first introduced by Resta
and Sorella in 1999 [137]. This work was followed soon afterwards by SWM
[138], where additional results are found, most notably relating localization to
conductivity. Since then, most authors (including the present one) have adopted
the notation hr r ic for the localization tensor, where c stays for cumulant.

67
The formulation within OBCs, using the language and the notations familiar in
quantum chemistry, dates since 2006 [112].
The tensor hr r ic has the dimensions of a squared length; it is an inten-
sive quantity that characterizes the ground-state many-body wavefunction as a
whole. Its key virtue is that it discriminates between insulators and metals: it is
finite in the former case and divergent (in the large-system limit) in the latter.
This is the main message of the present Chapter (and of the modern theory of
the insulating state): we are going to prove it below, Sect. 6.4
Several expressions for the localization tensor have been given in the lit-
erature, all of them equivalent; here we define the localization tensor as the
intensive quantity
hr r ic = F (0)/N, (6.11)
where the thermodynamic limit is understood. A glance at the OBCs expression,
Eq. (6.5), shows the cumulant nature of F and explains the reason for the
notation, which we adopt within PBCs as well.
Until 2005 the theory of the insulating state implicitly addressed time-
reversal invariant systems only, where the F tensor is real symmetric, and
coincides with the metric: in such systems therefore

hr r ic = g (0)/N. (6.12)

It was found in 2005 [139] thatin absence of time-reversal symmetry and


within PBCsthe tensor hr r ic is naturally endowed with an antisymmetric
imaginary part, whose physical meaning is also outstanding. This is discussed
in Sects. 7.1 and 7.2.

6.2.5 Discrete formula (single point)


The formalism adopted so far is based on a ground state 0 () which is a differ-
entiable function of . As usual in numerical work there are no derivatives, and
one needs a discrete formulation, convergent by construction to the continuous
theory. In fact the original formulation of the theory of the insulating state, as
appeared in Ref. [137], was discrete. We are going to prove the main formula
therein in reverse, addressing the 1d case only.
Suppose we have system of N electrons over a segment of length L, within
PBCs. The localization tensor, Eqs. (6.7) and (6.11), becomes then a localiza-
tion length
1
2 = hx2 ic = h 0 ()| Q() | 0 ()i . (6.13)

N =0
By a small- expansion we have
1
| 0 ()i (|0 ()i |0 (0)i),

1
hx2 ic (1 |h0 (0)|0 ()i|2 ). (6.14)
N 2
Next we notice that whenever equals 2/L the wavefunction eix |0 (0)i
obeys PBCs and is an eigenstate of H(), with eigenvalue E0 . For insulating
systems the ground state is nondegenerate even in the thermodynamic limit,

68
hence |0 ()i = eix |0 (0)i. Replacing into Eq. (6.14) we get
2
2 1 L 2
(1 |h0 (0)|ei L i xi |0 (0)i|2 )
P
hx ic
N 2
 2
2 L 2
P
Re log h0 |ei L i xi |0 i. (6.15)
N 2

We may compare at this point with the polarization theory for a correlated
many-electron system. The main formula, Eq. (5.18), becomes in the 1d case
e 2
P (el) = Im log h0 |ei L i xi |0 i.
P
(6.16)
2
It is clear therefore that the same matrix element yields both polarization and
localization. In the metallic case the matrix element goes to zero in the ther-
modinamic limit: diverges, and P is ill defined. In fact, polarization is not a
bulk material property for a metal.

6.3 Localization vs. conductivity


6.3.1 Linear response
A summary of linear-response theory is provided in Appendix C. If we perturb
a system in equilibrium with the input signal finput (t) the corresponding output
foutput (t) is given (to linear order) by the convolution
Z
foutput (t) = dt (t t )finput(t ); (6.17)

its Fourier-transformed form is

foutput () = () finput (). (6.18)

We now address our N -electron system in its unperturbed ground state |0 i


and we perturb the many-body Hamiltonian with a term

H(t) = finput(t)A, (6.19)

where A determines the shape of the perturbation and finput(t) its ampli-
tude. We wish to measure the response to such perturbation by means of the
expectation value of some observable B, i.e.:
= h|B|i h0 |B|0 i,
foutput(t) = hB(t)i (6.20)

where = 0 + (t) is the perturbed time-evolved ground state. If we limit


ourselves to study terms which are linear in the response, it is expedient to
switch to the domain. As shown in Sec. C.7, the response function can be
written in the form
() = hhB|Aii (6.21)

69
The quantity hhB|Aii is by definition the linear response induced by the per-
turbation A at frequency on the expectation value hBi. Straightforward
first-order perturbation theory provides its explicit expression as

1 X h0 |B|n ihn |A|0 i


hhB|Aii = lim+
~ 0 0n + i
n6=0
!
h0 |A|n ihn |B|0 i
, (6.22)
+ 0n + i

where 0n are the excitation frequencies of the unperturbed system: expressions


of the kind of Eq. (6.22) go under the name of Kubo formulae. The positive
infinitesimal ensures causality in the response. More about causality is said
in the Appendix, Secs. C.3 and C.7.

6.3.2 Conductivity in the scalar potential gauge


If we adopt the scalar potential gauge, the conductivity tensor () measures
the current linearly induced by an electric field: j = E . We therefore
identify A with the potential of an electric field along , i.e. A = eE r , and
B with the current operator ev /L3 . We anticipate that the scalar potential
gauge is legitimately adopted only in insulators, while for metals the vector po-
tential gauge is mandatory in order to actually address conductivity. However,
the main entry in the SWM sum rule (upon which the theory of the insulating
state is based) is the scalar-potential-gauge conductivity. It turns out that this
conductivity, used in Sec. 6.4, is a fake conductivity in the metallic case. It
would be better to write the SWM sum rule in terms of the imaginary part of
the dielectric function: in fact it is a polarizability sum rule more than a con-
ductivity sum rule: this is pretty clear from the alternative derivation given in
Ref. [112]. We follow here the (unfortunate) notations adopted by the literature
so far, and we will adopt the same symbol both for the real conductivity and
for the fake (in metals) conductivity. More about this will be said below, Sec.
6.6.2
We pause to stress an important detail. The macroscopic field inside the
sample includes by definition screening effects due to the electronic system, while
the perturbation H entering Eq. (6.1)via the A operatoris the bare, or
unscreened one. This point will be discussed below (Sect. 6.3.4); for the time
being we simply identify screened and unscreened fields.
The Kubo formula for conductivity is therefore
e2
() = hhv |r ii ; (6.23)
L3
this is correct within OBCs, but meaningless within PBCs, owing to the explicit
presence of the position operator. This, however, makes no harm, since only
its off-diagonal matrix elements are required: see Eq. (6.22). As usual, we may
exploit the identity h0 |r|n i = ih0 |v|n i/0n . The Kubo formula becomes
then
ie2 X 1  h0 |v |n ihn |v |0 i
() = lim
~L3 0+ 0n 0n + i
n6=0

70

h0 |v |n ihn |v |0 i
+ . (6.24)
+ 0n + i

We introduce a compact notation for the real and imaginary parts of the
numerators in Eq. (6.24), i.e.

Rn, = Re h0 |v |n ihn |v |0 i, (6.25)


In, = Im h0 |v |n ihn |v |0 i, (6.26)

which are symmetric and antisymmetric, respectively. Using then


1 1
lim = P i(x), (6.27)
0+ x + i x
and omitting the principal part, we separate for > 0 the symmetric and
antisymmetric parts in the conductivity tensor as

(+) e2 X Rn,
Re () = ( 0n )
~L3 0n
n6=0

() 2e2 X In,
Re () = = 2 2 . (6.28)
~L3 0n
n6=0

6.3.3 Sum rules


At this point, we are ready to compare with the sum-over-states formulae for
the tensor hr r ic . In the present notations, we rewrite Eq. (6.10) as

1 X Rn,
Re hr r ic = 2
N 0n
n6=0
1 X In,
Im hr r ic = 2 . (6.29)
N 0n
n6=0

A glance at Eq. (6.28) shows that


Z
~L3 d (+)
Re hr r ic = 2
Re () (6.30)
e N 0
~L3 ()
Im hr r ic = Re (0). (6.31)
2e2 N
Eq. (6.30) has been arrived at by SWM in 2000 [138], and Eq. (6.31) by Resta
in 2005 [139]. First of all, these identities show that hr r ic , defined here as a
basic geometric feature, is indeed a measurable quantity (whenever it does not
diverge).
As emphasized throughout this work hr r ic is a ground-state property,
while the rhs of Eqs. (6.30) and (6.31) are properties of the system excitations,
owing to the Kubo formula. Indeed, both Eqs. (6.30) and (6.31) look like the
zero-temperature limit of a fluctuation-dissipation theorem, several forms of
which are known in statistical physics [140, 141]: in the lhs we have a ground-
state fluctuationsee in particular Eq. (6.5)while the ingredient of the rhs is
conductivity (dissipation).

71
6.3.4 Screened vs. unscreened field
The Kubo formula for conductivity has been obtained identifying the A operator
with E r , where E is the macroscopic field inside the sample. In general this
is not quite correct, since instead A = eE0 r , where E0 is the bare field, i.e.
the field that would be present inside the sample in absence of screening. The
latter originates from the two-body (electron-electron) terms in the potential V
entering Schrodinger equation. The relationship between E and E0 is not a bulk
property, and depends on the shape of the sample. Alternatively, it depends on
the boundary conditions assumed for integrating Poisson equation: we refer to
Ref. [8] for a thorough discussion. Whenever E = 6 E0 , the sum rule in Eq. (6.30)
must be modified.
The ground-state fluctuations, as e.g. Eq. (6.5), are fluctuations of the
macroscopic polarization, which in a finite sample induce a surface charge at
the boundary. This in turn generates a depolarizing field, which counteracts
polarization. Therefore the localization tensor hr r ic depends on the sample
shape, or equivalently on the boundary conditions assumed when taking the
thermodynamic limit. The choice of PBCs in Eq. (6.1), however, implies E = E0
[142]; hence Eqs. (6.30) and (6.31) are correct as they stand within PBCs. This
no longer holds within OBCs: in this case Eq. (6.30) needs to be modified, while
Im hr r ic = 0.
Ideally the equality E = E0 corresponds to choosing a sample in the form
of a slab, and to addressing the component of the fluctuation tensor hr r ic
parallel to the slab [8]; the thermodynamic limit amounts then to the infinite
slab thickness. Owing to the long range of Coulomb interaction the order of
the limits (first a slab, then its infinite thickness) is crucial. For instance, if
the limit is taken instead by considering spherical clusters of increasing radius,
the SWM fluctuation-dissipation sum rule, Eq. (6.30), assumes a different form:
this is discussed in Refs. [112, 142]. The explicit form of the generalized sum
rule is given therein.
Last but not least, the effect leading to E = 6 E0 within OBCs is a pure
correlation effect. It originates from explicitly correlated wavefunctions, and
does not occur within mean-field theories (Hartree-Fock and Kohn-Sham) [112,
142]. Within such theories, therefore, the sum rules hold in the simple form
of Eqs. (6.30) and (6.31); the conductivity therein is the independent-particle
conductivity (uncoupled response in quantum-chemistry jargon).

6.4 Localization in the insulating state


The basic tenet of the modern theory of the insulating state is that the localiza-
tion tensor hr r ic is the ground-state property which sharply discriminates
in the spirit of Kohns seminal work [134, 135]between insulators and metals.
The real part of hr r ic remains finite in the thermodynamic limit in any insu-
lator, while it diverges in any metal.
The theory is very general, and has found applications to various different
kinds of insulators: band insulators [143, 144, 145, 146]; correlated (i.e. Mott)
insulators, either by means of Hubbard-like model Hamiltonians [137, 147] or
realistic ones [148, 149, 150]; and Anderson insulators [151]. Some of these
applications are reviewed in Sect. 6.7. The cases of Chern insulators, quantum

72
Hall insulators, and topological insulators are discussed in Chap. 7.
The ultimate proof of the key property of Re hr r ic is based on the SWM
sum rule, Eq. (6.30). Since the tensor is real symmetric, it is enough to consider
the diagonal elements (over its principal axes)
Z
~L3 d
Re hr r ic = 2 Re (). (6.32)
e N 0

The f -sum rule yields


p2 e2 N
Z
d Re () = = , (6.33)
0 8 2me L3

where p is the plasma frequency. Therefore the integral in Eq. (6.32) always
converges at ; its convergence/divergence is dominated by the small- behav-
ior of Re ().
Suppose first that the spectrum is gapped, i.e. the spacing between the
ground state and the first excited state stays finite in the thermodynamic limit.
If the gap is Eg the conductivity vanishes for < Eg /~, and Eq. (6.33) yields

~L3 d
Z
Re hr r ic = Re ()
e2 N Eg /~
Z
~2 L 3
< d Re ()
e2 N Eg 0
~2
= . (6.34)
2me Eg

This inequality is due to SWM and clearly proves that Re hr r ic is finite in


any gapped insulator, as e.g. band insulators (considered in more detail in Sect.
6.4.2).
The main message of Kohns 1964 paper, however, is that insulating char-
acteristics are a strict consequence of electronic localization (in an appropriate
sense) and do not require an energy gap. For any gapless material, the small-
behavior of Re () is the result of a competition between numerators and
denominators in the Kubo formula, Eq. (6.24). Since we aim at a continuous
function of , the singularities in Eq. (6.28) must be smoothed: this can be done
by keeping the dissipation finite while performing the thermodynamic limit
first [152]. For a band metal the localization tensor diverges (see below). Ac-
cording to SWM, a gapless material is insulating whenever Re () 0 like
a positive power of , and metallic otherwise. The only example of gapless in-
sulator considered so far is a model Anderson insulator in 1d [151]. Simulations
prove indeed that hx2 ic is finite therein (Sect. 6.7.4).

6.4.1 Independent electrons


For noninteracting electrons the potential V in Eq. (6.1) is the sum of identical
PN
one-body terms: V = i=1 V (ri ). The many-electron Hamiltonian is separa-
ble and the exact ground state |0 ()i is a Slater determinant of one-particle
orbitals (doubly occupied in the singlet case). At a mean-field level, the one-
body potential V (r) includes electron-electron interaction in a selfconsistent

73
way. In the Hartree-Fock (HF) framework the Slater determinant is regarded
as an approximate many-electron wavefunction. Instead, in the DFT frame-
work the orbitalscalled Kohn-Sham (KS) orbitalsare auxiliary quantities,
individually devoid of physical meaning. In particular, their Slater determinant
does not coincide with the many-electron wavefunction |0 ()i, as a matter of
principle. Therefore the exact localization tensor does not coincide, at least in
principle, with the one obtained from the Slater determinant of KS orbitals; the
issue is similar to the one discussed above for macroscopic polarization P (Sec.
5.4.2). We stress, however, a difference: the exact P coincides with the KS P
within OBCs, and the problem occurs only within PBCs. Instead, the exact
localization tensor differs in principle from the KS one within both OBCs and
PBCs.
So much for the matters of principle. On practical grounds such difference
is routinely disregarded (e.g. when dealing with polarization, magnetization
[5, 6, 8], and more), given that it is not at all clear what is the relative im-
portance of this intrinsic error, compared with the errors due to the choice
of the functional itself. We therefore address here either the HF or the KS
wavefunction |0 ()i, having the form of a Slater determinant.
Whenever the wavefunction is a Slater determinant, all ground-state prop-
erties can be explicitly cast in terms of the one-body density matrix
N/2
X
(r, r ) = 2P (r, r ) = 2 j (r)j (r ), (6.35)
j=1

where a singlet ground state is assumed, and j (r) are the occupied one-particle
orbitals (either HF or KS); P (r, r ) is the projector over the occupied manifold.
The expression for the localization tensor is easily found within OBCs start-
ing from Eq. (6.5) [112]:

1
Z
hr r ic = drdr (r r ) (r r ) |P (r, r )|2 . (6.36)
N
If we define the complementary projector

Q(r, r ) = (r r ) P (r, r ), (6.37)

an equivalent expression is [143]


2
hr r ic = Tr {r P r Q}, (6.38)
N
where Tr is the trace over the single-particle Hilbert space (not on Cartesian
indices).
If we consider a cluster, cut out of a crystalline solid, Eq. (6.36) becomes in
the large-N limit
1
Z Z
hr r ic = dr dr (r r ) (r r ) |P (r, r )|2 , (6.39)
Nc cell all space

where Nc is the number of electrons per crystal cell. According to the discussion
in Sec. 6.3.4, we need not to worry about shape issues in taking the limit; we

74
also stress that the density matrix, Eq. (6.35), is independent of the boundary
conditions (either OBCs or PBCs) in the large-N limit.
Finally, we notice that the nasty position operator r is ill defined within
PBCs. But only the relative coordinates appear in Eq. (6.39), while instead in
Eq. (6.38) the operator r is sandwiched between a P and a Q. The main
feature making this expression viable in the thermodynamic limit is that P rQ
is well defined even within PBCs. More details about this are given below, Sec.
8.2.1.

6.4.2 Band insulators and band metals


As observed, Eq. (6.39) holds for a crystalline solids. Therefore the inner integral
on the rhs must converge for a band insulator, and must diverge for a band metal.
This is confirmed by the well known fact that the asymptotic behavior of P is
qualitatively different in insulators and in metals. In the former materials,
in fact, P (r, r ) decays exponentially [153, 154, 155, 156] for large values of
r r : therefore the integral converges and the localization tensor is finite.
In conducting materials, instead, P (r, r ) decays only polynominially, and the
inner integral diverges. This divergence can be explicitly verified for the simplest
conductor of all, namely, the noninteracting electron gas, whose density matrix
is exactly known in analytic form [143, 157] for 1d, 2d, and 3d. In all these
cases, the inner integral in Eq. (6.39) diverges like the linear dimension of the
system.
It is therefore clear that the localization tensor, when expressed in the form
of Eq. (6.39), measures in a perspicuous way the nearsightedness[158] of the
electron distribution. Such measure is qualitatively different in insulators and
in metals.
The one-particle orbitals (either HF or KS) in a crystalline solid have the
Bloch form. We therefore may wish to replace the orbitals in the expression for
P (r, r ), Eq. (6.35),
j (r) jk (r) = eikr ujk (r), (6.40)
where j is the band index and k is the Bloch vector. We stress that PBCs are
at the very root of Bloch theorem. If the orbitals are normalized to one over
the crystal cell of volume Vcell , the ground-state projector in insulating crystals
is
n Z
Vcell X
P (r, r ) = dk jk (r)jk
(r ) (6.41)
(2)3 j=1 BZ
n Z
Vcell X
= dk eik(rr ) ujk (r)ujk (r ),
(2)3 j=1 BZ

where n = Nc /2 is the number of occupied bands, and the integral is taken over
the Brillouin zone (equivalently, over the reciprocal cell).
Using Eq. (6.41), the localization tensor in a band insulator becomes (for
double occupancy) proportional to BZ integral of the Bloch metric-curvature
tensor, Eq. (3.47), i.e.

Vcell
Z
hr r ic = dk F (k)
(2)3 n BZ

75
n
X n
X
F (k) = h ujk | ujk i h ujk |uj k ihuj k | ujk i. (6.42)
j=1 jj =1

The proof is given in Refs. [4, 143], and will not be repeated here.
Within OBCs the localization tensor is always real: this is perspicuous
in Eqs. (6.36) and (6.39). Instead the imaginary part of the PBC hr r ic ,
Eq. (6.42), is the BZ integral of the Bloch Berry curvature, discussed below,
Sec. 7.1

6.5 Wannier and hermaphrodite orbitals


The Wannier and hermaphrodite orbitals are only defined for an independent-
electron insulator; as shown below, their theory makes use of expressions such
as Eq. (6.42) and the like. But these orbitals deserve a discussion of their own;
we limit ourselves to the crystalline case.

6.5.1 Wannier orbitals


The unitary transformation which defines the WFs, Eq. (5.6), labeled by band
j and unit cell R, is
Vcell
Z X
ikR
|wjR i = dk e |jk i ; |jk i = eikR |wjR i. (6.43)
(2)3 BZ
R

Bloch orbitals are normalized over the unit cell, while WFs are square-integrable
and normalized over R3 (in 3d). Any WF centered in R is just the translate of
a central-cell WF, i.e.:
Vcell
Z
hr|wjR i = hr R|wj0 i, |wj0 i = dk |jk i. (6.44)
(2)3 BZ
The density matrix is clearly invariant by any unitary transformation of the
occupied orbitals among themselves. The transformation of Eq. (6.35) to WFs
yields
n Z
Vcell X X
P = 3
dk |jk ihjk | = |wjR ihwjR |. (6.45)
(2) j=1 BZ
jR

In several circumstances, Eq. (6.45) is the most straightforward bridge be-


tween PBCs and OBCs; localized orbitals have a long history within quantum
chemistry.
The WF definition, Eq. (6.43), leaves a large arbitrariness (a.k.a. gauge
freedom), given that: (1) the Bloch orbitals |jk i are arbitrary by a k-dependent
phase factor, and (2) the band labeling j is ambiguous in case of band crossings.
It is expedient to fix a criterion for defining the best localized orbitals. Al-
though the criterion is somewhat arbitrary, the orbitals obtained in this way
often carry some intuitive physical meaning, and are in some sense additive: a
large system can be regarded as the sum of its parts. In quantum chemistry, a
very popular criterion is due to Boys [159]: it amounts to minimize the quadratic
spread of the orbitals, averaged over the molecule. The same criterion has been
adopted and implemented for crystalline systems by Marzari and Vanderbilt in
1997 [109], and has subsequently enjoyed a very large popularity [19].

76
6.5.2 Hermaphrodite orbitals
Hermaphrodite orbitals were first introduced in Ref. [143]. It is expedient to
start with a finite crystallite within OBCs, with N electrons in a singlet ground
state:
N/2
X
=2P =2 |i ihi |. (6.46)
i=1
The average quadratic spread in the x direction is by definition
N/2
2 X
2xx = (hi |x2 |i i hi |x|i i2 ). (6.47)
N i=1

We recast this identically as:


2 X X 2 X
2xx = hi | x ( 1 |j ihj | ) x |i i + |hi |x|j i|2 .
N i j
N
i6=j
(6.48)
The first term in Eq. (6.48) is gauge invariant, since we can identically write:
2 2 X
2xx = Tr xP x(1 P ) + |hi |x|j i|2 . (6.49)
N N
i6=j

The gaugeinvariant term in Eq. (6.49) coincides with the xx element hx2 ic of
the localization tensor hr r ic , Eq. (6.38).
If we look for the orbitals which minimize the average spread in the x di-
rection, the solution, after Eq. (6.49), is provided by those orbitals which diag-
onalize the position operator x, projected over the occupied manifold, i.e. the
operator P xP . Obviously, a set of orthonormal orbitals which diagonalize it
can always be found, since P xP is a Hermitian operator. The quadratic spread
of these orbitals is the minimum and equals then hx2 ic .
If we now switch to the thermodynamic limit for a crystalline system, trans-
lational symmetry implies P (r, r ) = P (r+ R, r + R). For the sake of simplicity
we limit ourselves to a rectangular lattice, where R = (X, Y, Z), and each of
the X, Y, Z are one-dimensional lattices. It is then obvious that the eigenstates
of P xP can be labelled with a Bloch vector in the yz directions. If we use the
notation s ky kz (r) for a generic eigenstate of P xP , then it obeys the Bloch
theorem in the form
s ky kz (x, y + Y, z) = eiky Y s ky kz (x, y, z), (6.50)

and analogous in the z direction.


We address the simple case of a single occupied band. If 0 ky kz (x, y, z)
is eigenstate of P xP with eigenvalue x0 , then its lattice-translate 0 ky kz (x
X, y, z) is also eigenstate of P xP , with eigenvalue x0 + X: this is reminiscent
of WFs, see Eq. (6.44). The proof is very simple: P (x X)P = P xP XP ,
and XP on the occupied manifold is just a constant multiplication by X.
In the many-band case we needas for WFsa band index j and a 1d cell
index X. All hermaphrodite orbitals are obtained by lattice translation from
the central cell ones:

hx, y, z|j X ky kz i = hx X, y, z|j 0 ky kz i; (6.51)

77
Figure 6.1: Two-masted vessels. Brig: both masts are square rigged. Schooner:
both masts are fore-and-aft rigged (gaff sails and topsails here). Hermaphrodite
brig: the foremast is square rigged, the mainmast is fore-and-aft rigged.

a glance at Eq. (6.44) explains the notations and the meaning. Notice that the
hermaphrodite-orbital center x0 does depend on the (ky , kz ) Bloch vector, as
well as on the band index j.
The name hermaphrodite orbitals is therefore pretty clear (see Fig 6.1):
they are Bloch-like in the yz direction, and Wannier-like in the x direction.
The present construction guarantees that they are maximally localized in the
x direction and their quadratic spread is equal to the xx component of the
(gauge-invariant) localization tensor. Similarly to standard Bloch and Wannier
orbitals, our hermaphrodite orbitals are an orthonormal basis for the occupied
manifold.

6.5.3 Maximally localized Wannier orbitals


For a one-dimensional system the hermaphrodite orbitals, as defined above, are
those localized orbitals which minimize the quadratic spread, a.k.a. MLWFs
(maximally localized WFs). The minimum spread equals the gauge-invariant
cumulant moment, a.k.a. squared localization length hx2 ic .
For a 3d system we are interested in minimizing the spherical quadratic
spread. In general we cannot diagonalize simultaneously P xP , P yP , and P zP ,
and the spherical spread will be strictly larger than the Cartesian trace of the
localization tensor. This is a key feature in the work of Boys [159], and of
Marzari and Vanderbilt well. The relationship between our localization tensor
hr r ic with the gauge-invariant quadratic spread I common in the MLWF
literature [109, 19] is
d
X 1
hr r ic = I , (6.52)
=1
n

where n is the number of occupied bands. As said above, the trace in Eq. (6.52)
is a lower bound (and not a minimum in dimension d > 1) for the spherical
second (cumulant) momenta.k.a. quadratic spreadof the WFs, averaged
over the sample.
We stress that Eqs. (6.41) and (6.42) make sense only insofar the Fermi level
falls in a gap, in which case hr r ic is always finite. If we vary the Hamiltonian
continuously, allowing the gap to close, then hr r ic diverges; the quadratic
spread of the Wannier functions diverges as well [112].

78
Our previous findings bear an important additional message concerning Boys
localization. For a cluster of finite size (or for a crystallite), no matter how large,
one can doubtless build localized Boys orbitals within OBCs. But our results
prove that, in the large N limit, the average quadratic spread of these Boys
orbitals diverges whenever the cluster is metallic.

6.6 More about conductivity in metals


6.6.1 Drude phenomenological conductivity
The conductivity is scalar in isotropic metals:
j() = ()E(). (6.53)
We use the standard conventions about Fourier transforms in the time-frequency
domains, as here in the Kubo formula and in Eq. (C.1); other conventions may
change the sign of the imaginary parts in the response functions (e.g. Kohn,
[134]). The Drude phenomenological formula is [160]
ie2 (n/m)eff
Drude () = , (6.54)
+ i/
where (n/m)eff is the quantum analogue of the original n/m in the classical
theory. The dc limit is purely dissipative:
Drude (0) = e2 (n/m)eff . (6.55)
We rewrite Eq. (6.54) as
i D
Drude () = , (6.56)
+ i
where D = e2 (n/m)eff is the Drude weight and = 1/ ; > 0 ensures a causal
response.
The real and imaginary parts of denote in-phase (dissipative) and out-of-
phase (reactive) response to the E field. Within the Drude model
1 D 1 D
Re Drude () = ; Im Drude () = . (6.57)
2 + 2 2 + 2
In the nondissipative ( 0+) limit we get
D 1
Re Drude () = D () ; Im Drude () = P , (6.58)

where P denotes the principal part. The dc ( = 0) in-phase conductivity has
a -like divergence: this accounts for the obvious fact that free electrons in a
constant field undergo free acceleration, and the current does not reach a steady
state limit. Equivalently, we may say that the system has an undamped normal
mode at = 0.
Allens review [160] does not explicitly define the Drude weight. Here we
follow Scalapino et al. [161, 162], who define D as the coefficient of the () in
the non dissipative limit or, equivalently, as
D = lim Im Drude () at = 0. (6.59)
0

79
Notice that the D as defined by SWM [138] is different (by a factor of 2 in
a.u.).
It is worth emphasizing that the Drude weight also has a more general mean-
ing. Eq. (6.56) has a pole at = i and D measures its -independent residue.
Using some well known algebra the conserved spectral weight is [160]
Z
d Drude () = D, 6= 0. (6.60)

We expect the leading term of () at small to behave exactly as Eq. (6.58)


in crystalline metals: (n/m)eff measures how many electrons in the metal
undergo free acceleration under the application of a constant field.
Clearly the Kubo formula within the scalar potential gauge, Sec. 6.3.2 lacks
the Drude peak (i.e. the leading term) in the dissipationless conductivity;
equivalently, the scalar potential gauge misses the free acceleration in a metallic
system. In order to recover free accelerations, one must adopt both PBCs and
the vector potential gauge. Other choices can only yield the fake conductivity
of Sec. 6.3.2

6.6.2 Conductivity in the vector potential gauge


Here the perturbation is an dependent vector potential along . One key
point is that the vector potential modifies the velocity operator. We stick to
the symbol v for the velocity in absence of the perturbation: this may include
a ground-state vector potential, but not the perturbing one. We thus have
eN
h0 | v |0 i h0 | v |0 i + A
mc
e e2 N
j 3 h0 | v |0 i A (6.61)
L mcL3
If A is the vector potential (along ) linearly induced by a an electric field E
in the direction, the linear response to the leading order in E is, in Zubarev
notations,

e2 N
 
e e
j = ()E = A hh (v | A v ii E
mcL3 L3 c
e2
 
N
= 3 + hhv |v ii A() E , (6.62)
cL m

where we are restoring the dependence. The term in A2 , being constant


in space, has zero matrix elements; it is also second order in E. We are using
here the same symbol in order to agree with the published literature, but we
stress that Eqs. (6.23) and (6.62) define two different linear responses of the
many-electron system.
The vector potential A and the field E are both along and are related
by E = 1c A/t, but only for t > 0; A = 0 for t < 0. Switching to the
domain  
1
A() = cE () , (6.63)
i

80
where the term is required by causality (see Sec. C.3). Therefore the current,
as expressed directly in terms of the field intensity, is
e2 N
  
1
j = ()E = 3 + hhv |v ii () E . (6.64)
L m i
Let us focus for the time being on longitudinal conductivity only:
e2
  
N 1
Re () = + Re hhv |v ii () Im hhv |v ii (6.65)
L3 m
2
  
e N 1
Im () = + Re hhv |v ii + Im hhv |v ii () (6.66)
.
L3 m
We then write explicitly Zubarevs double commutator, which by definition is
1 X  h0 |v |n ihn |v |0 i h0 |v |n ihn |v |0 i 
hhv |v ii = lim
~ 0+ 0n + i + 0n + i
n6=0
 
1 X
2 1 1
= lim |h0 |v |n i| .(6.67)
~ 0+ 0n + i + 0n + i
n6=0

We may now perform the weak limit of the distributions, using Eq. (6.27), from
which we get
2 X 0n
Re hhv |v ii = |h0 |v |n i|2 2 2
~ 0n
n6=0
X
Im hhv |v ii = |h0 |v |n i|2 [ ( 0n ) ( + 0n ) ](6.68)
.
~
n6=0

We also notice that the real part of hhv |v ii is even and the imaginary part
is odd: therefore the last term in parenthesis in Eq. (6.66) vanishes.
If we now evaluate the Drude weight from the imaginary part of the conduc-
tivity we get, using Eq. (6.59),

2 2
e N 2 X |h0 |v |n i|
D = lim Im () = 3 . (6.69)
0 L m ~ 0n
n6=0

If instead we look at the real part of the conductivity, the coefficient of ()


in Eq. (4.53) gives the same Drude weight as above, as it must be because
our response respects causality. The remaining term in Eq. (4.53) is even and
yieldsin the thermodynamic limitthe real part of the regular term in the
conductivity.
We may summarise the the above results in the following way. For either
metals or conductors, within either PBCs or OBCs, the longitudinal conductiv-
ity is always written as
(Drude) (regular)
() = () + (),
(Drude) D
() = D () + i ,

(regular) e2 1 X
Re () = |h0 |v |n i|2 [ ( 0n ) ( + 0n ) ] ,
~L3
n6=0

81
(regular) 2e2 X |h0 |v |n i|2 1
Im () = , (6.70)
~L3 0n 0n 2 2
n6=0

where the Drude weight is



e2 N 2 X |h0 |v |n i|2
D= 3 . (6.71)
L m ~ 0n
n6=0

Both (Drude) and (regular) obey the Kramers-Kronig relationships separately.


For insulators the Drude weight vanishes. This can be verified in various
ways; the simplest one is to relate (at > 0) the imaginary part of the conduc-
tivity to the real part of the dielectric function: () = () (0), and
then use Eq. (6.69).
Therefore in insulators the whole conductivity is given by the regular part
only. We notice that the regular part is identical in form to what one obtains
within the scalar potential gauge. Even if the matrix elements are different, it is
easy to guess that the two a conductivities are indeed identical. The reasoning
is based on general nearsightedness arguments in insulators. Using then the
SWM sum rule, we get the same localisation tensor (symmetric part only) within
either gauge, and even within either PBCs or OBCs.

We verify the f -sum rule, Eq. (6.33), by integrating () over all positive
frequencies. For insulators we get:

e2 X |h0 |v |n i|2 e2 p2
Z

d () = = = , (6.72)
0 ~L3 0n 2mL3 8
n6=0

where we have used Eq. (6.71) with D = 0. For metals the sum rule has a Drude
contribution and a regular contribution. We verify this:
Z
D e2 X |h0 |v |n i|2
d () = +
0 2 ~L3 0n
n6=0

e2 N 1 X |h0 |v |n i|2 e2 X |h0 |v |n i|2
= +
L3 2m ~ 0n ~L3 0n
n6=0 n6=0

p2
= , (6.73)
8
where the Drude peak at = 0 contributes by 1/2 to the integral.

6.7 Localization in different kinds of insulators


6.7.1 Small molecules
The modern theory of the insulating state clearly addresses extended sys-
tems, i.e. the N limit; indeed it makes little sense to ask whether a
small molecule is insulating or conducting. Nonetheless the concepts of lo-
calized/delocalized electronic states is of the utmost importance in quantum
chemistry as well, notably in relationship to aromaticity.

82
The tensor hr r ic within OBCs is always real symmetric. If the ground-
state wavefunction is a Slater determinant, then the trace of the tensor at finite
N has the meaning of a lower bound for the quadratic spread of the Boys
localized orbitals, averaged over all the occupied orbitals [112].
The small-N version of the main concepts of the present review (in their
OBCs flavour [112]) has been adopted in quantum chemistry by Angyan
[163, 164]. Besides providing HF calculations of hr r ic for a sample of small
molecules, Angyan even provides experimental values drawn from compilations
of the dipole oscillation-strength distributions: basically, from Eq. (6.30).

6.7.2 Band insulators


The theory warrants that the localization tensor is finite in any insulator. How-
ever, quantitative calculations for both model tight-binding Hamiltonians and
realistic solids within DFT have been used to illustrate the theory and to iden-
tify trends. For instance one expects much smaller diagonal elements hr r ic
for strong (i.e. large-gap) insulators than for weak (small-gap) ones. This is
also suggested by the SWM inequality, Eq. (6.34).
Let us start with a simple tight-binding (a.k.a. Huckel) Hamiltonian in 1d:

[ (1)j cj cj t(cj cj+1 + H.c.) ]


X
H = (6.74)
j

where t > 0 is the first neighbor hopping ( = t in most chemistry literature)


and H.c. stays for Hermitian conjugate. This toy model schematizes a binary
ionic crystal; the band structure is
p
(q) = 2 + 4t2 cos2 qa/2, (6.75)

where a is the lattice constant and q is the Bloch vector. The gap is equal to
2; at half filling the system is always insulating except for = 0. The squared
localization length (within OBCs) is the tight-binding version of Eq. (6.36), i.e.
N
a2 X 2
hx2 ic = Pjj (j j )2 . (6.76)
4N
j,j =1

This is a monothonical function of t/; it is easily verified that it vanishes in


the extreme ionic case (t = 0). In the metallic case ( = 0) the ground-state
projector has a simple analytical form:
1
Pjj = ; Pjj = 0 for even |j j| = 2s,
2
(1)s
Pjj = for odd |j j| = 2s + 1, (6.77)
(2s + 1)

which clearly implies divergence of Eq. (6.76). At any finite N within OBCs
Eq. (6.76) leads to a finite hx2 ic value; however Eq. (6.77) suggests that in the
metallic case hx2 ic diverges linearly with N , in qualitative agreement with the
case of the 1d electron gas. This has been verified by actual simulations, even
when 6= 0 but the Fermi level is not in the gap [151].

83
Figure 6.2: Diagonal element of
the KS localization tensor vs.
the inverse direct gap (theoret-
ical and experimental), for sev-
eral elemental and binary semi-
conductors (from Ref. [143])
The points corresponding to Si
and Ge with the theoretical
gaps are out of scale. From Ref.
[143].

Other simulations [146] have addressed dimerized chains, i.e. = 0 but


alternant hoppings in Eq. (6.74). While nothing relevant occurs within PBCs,
partly filled end states within OBCs at some fillings are at the root of some
noticeable features.
The first ab-initio study (in 2001) addressed several elemental and binary
cubic semiconductors at the KS level [143]. The tensor is real and isotropic.
The computed hx2 ic (Fig. 6.2) is smaller than 3 bohr2 in all the materials
studied: the ground many-body wavefunction is therefore very localized in this
class of materials. The SWM inequality was also checked, and found to be well
verified using both the theoretical KS gap and the experimental one (the latter
is typically larger).
Other studies have addressed the ferroelectric perovskites in their different
(cubic and noncubic) structures [144], and some model Hamiltonians in 1d and
2d [145].

6.7.3 Correlated (Mott) insulators


Several computational studies of the Mott transition have adopted the localiza-
tion tensor with a correlated wavefunction as a marker for the metal-insulator
transition. These studies are invariably for 1d systems, either with Hubbard-
like lattice Hamiltonians, or for quasi-1d atomic chains. In both cases the lo-
calization marker is hx2 ic , and the discrete (single-point) formula Eq. (6.16) is
adopted.
Arguably the simplest possible 1d highly correlated system obtains from the
noninteracting Hamiltonian of Eq. (6.74) and augmenting it with an on-site
repulsive term we get the two-band Hubbard model

[(1)j cj cj t(cj cj+1 + H.c.)] + U


X X
H = nj nj . (6.78)
j j

The explicitly correlated ground-state wavefunction has been found by exact


diagonalization [137], and the corresponding hx2 ic has been computed as a
function of U for fixed t/ = 1.75. The results are shown in Fig. 6.3 in
dimensionless units; it turns out that there is only one singular point U = 2.27t,

84
Figure 6.3: Squared localiza-
tion length for the Hamiltonian
in Eq. (6.78) at half filling for
t/ = 1.75. The system un-
dergoes a quantum phase tran-
sition from band-like insulator
to Mott-like insulator at U/t =
2.27. From Ref. [137].

where hx2 ic diverges. Indeed, it has been verified that at such value the ground-
state becomes degenerate with the first excited singlet state, i.e. the system is
metallic. The singular point is the fingerprint of a quantum phase transition:
on the left we have a band-like insulator, and on the right a Mott-like insulator.
The two insulating states are qualitatively different; by adopting the modern
jargon, nowadays we could say that they are topologically distinct. The static
ionic charges (on anion and cation) are continuous across the transition, while
the dynamical (Born) effective charge on a given site changes sign [165]. Other
studies of the localization tensor within the same Hubbard model can be found
in Ref. [166]. Other studies of the Mott transition with 1d model Hamiltonians,
and based on the localization length hx2 ic , have appeared [167, 168].
The transition from a band metal to a Mott insulator has been studied in
a model linear chain of Li atoms by Vetere et al. [148]. At a mean-field level
the infinite chain is obviously metallic at any lattice constant a, since there
is one valence electron per cell. However the mean-field description becomes
inadequate at large a, where the electrons localize and the system becomes a
Mott insulator. If electron correlation is properly accounted for at any a, the

Figure 6.4: Results for chains


of H atoms of different lengths
as a function of the interatomic
distance a, after Ref. [149].
Top
p panel: /a, where =
hx2 ic . Bottom panel: the
modulus of the matrix element
in Eq. (6.16).

85
system undergoes a sharp metal-insulator transition at a critical a.
The calculations addressed linear LiN systems (N up to 8) within OBCs,
where the finite size prevents a sharp transition; the tradeoff is that full config-
uration interaction was affordable with 6 atomic orbitals per site (yielding more
than 109 symmetry-adapted Slater determinants). The wavefunction of Vetere
et al. is therefore exempt from any bias insofar as the treatment of correlation
is concerned, although its quality is determined by the basis set. A study of the
longitudinal component hx2 ic of the localization tensor indicates rather clearly
the occurrence of the metal-insulator transition at a 7 bohr; other indicators
give concordant results [148]. For comparison, the nearest-neighbour distance
in 3d metallic lithium is 5.73 bohr.
The modern theory of the insulating state has also been applied to the
Mott transition on 1d hydrogen chains [149, 150]. We reproduce here Fig.
6.4 from Ref. [149] by Stella et al. who have performed variational quantum
Monte Carlo studies, up to 66 atoms, within PBCs. The crossover between
the weakly correlated (band) metallic regimeat small aand the strongly
correlated (Mott) insulating regimeat large ais clearly visible in both panels
of Fig. 6.4, which indicate the transition at a 3.5 bohr. The bottom panel
shows that the modulus of the matrix element in Eq. (6.16) switches from zero to
one in a narrow a region, the transition becoming sharper at increasing p size. The
top panel perspicuously shows that in the Mott-insulating regime = hx2 ic is
size-insensitive, while in the metallic regime it diverges with size. Unfortunately,
the authors have chosen to plot /a instead of itself. Therefore the value in
the large a limit cannot be verified: we expect that it goes to the isolated-atom
limit, i.e. = 1 bohr.

6.7.4 Disordered (Anderson) insulators


We start from the same Hamiltonian as in Eq. (6.74), and we replace the ordered
string (1)j by a random string of 1, chosen with equal (and uncorrelated)
probability. This system models a random binary alloy at 50% concentration.
It is well known both from analytical arguments and actual simulations that its
spectrum is gapless [133, 169]. The density of states for both the ordered and
disordered systems are shown in Fig. 6.5, and confirm the expected features.
The band structure of Eq. (6.75) yields obviously a gapped density of states; at
the band
edges it shows van Hove singularities, which in 1d have the character
of 1/ divergences. As discussed above, the system is insulating at half fill-
ing and conducting otherwise. The disordered system, instead, is gapless and
nonetheless insulating at any filling. In fact, this model Hamiltonian describes
a paradigmatic Anderson insulator in 1d.
The conventional theory of transport focusses on the nature of the one-
particle orbitals at the Fermi level; in Anderson insulators these are localized,
thus forbidding steady state currents [131]. More than fifty years of literature
have been devoted to investigate Anderson insulators under the most diverse
aspects [133, 169, 170, 171].
At variance with such wisdom, Ref. [151] addressed this paradigmatic An-
derson insulator from the nonconventional viewpoint of the modern theory of
the insulating state. In the spirit of Kohns theory the individual Hamilto-
nian eigenstates become apparently irrelevant, while the focus is on the many-
electron ground state as a whole. The squared localization length hx2 ic has been

86
Figure 6.5: Density of
states (arbitrary units) for
a model binary alloy in 1d.
The crystalline (band) case
corresponds to the Hamilto-
nian of Eq. (6.74) with =
0.25 and t = 1. The disor-
dered (Anderson) case cor-
responds to a random choice
of the anion/cation distri-
bution.

computed within OBCs from Eq. (6.76), and found to be finite, as expected.
Nonetheless its value is about 20 times larger than the one for the band insula-
tor, at the same value of the parameters (i.e. = 0.25, t = 1). This reflects the
fact that the scattering mechanisms are profoundly different: incoherent (An-
derson) versus coherent (band). In the latter case, the Hamiltonian eigenstates
are individually conducting butlocked by the Pauli principle if the Fermi level
lies in the gap.

87
Chapter 7

Topological invariants and topological


insulators

7.1 Chern invariant


Within OBCs the localization tensor is always real: this is perspicuous in
Eqs. (6.36) and (6.39). Instead the imaginary part of the PBC hr r ic ,
Eq. (6.42), is the BZ integral of the Bloch Berry curvature, Eq. (3.50):

Vcell 2Vcell
Z Z
Im hr r ic = dk Im F (k) = dk (k)
(2)3 n BZ (2)3 n BZ
n
Vcell
Z X
= dk Im h ujk | ujk i. (7.1)
(2)3 n BZ j=1

A necessary condition for this to be nonzero is the absence of time-reversal


symmetry.
The integral in Eq. (7.1) has (in 3d) the dimensions of an inverse length,
and is quantized in units of times a reciprocal vector. We call this reciprocal
vector Chern invariant; in fact it is a Chern number (Sect. 3.4) in k-space
and in appropriate units. In 2d the BZ integral is dimensionless and simply
related to the Chern number, defined as:
1 1
Z Z X
C1 = dk (k) = dk Im h1 ujk |2 ujk i. (7.2)
2 BZ BZ n

Band insulators where the Chern invariant is nonzero are called Chern insula-
tors (normal insulators otherwise). The possible existence of Chern insulators
has been pointed out by Haldane in 1988 [172], by means of a remarkable model
Hamiltonian in 2d, illustrated below (Sect. 7.3.1). We emphasize that the non-
vanishing of the Chern number prevents it prevents the existence of a smooth
periodic gauge, Eq. (3.42), across the whole reciprocal cell (or BZ): if, for in-
stance, Eq. (3.42) is enforced on the reciprocal cell boundary, an obstruction
will show up at some point inside. Because of this same reason, exponentially
localized Wannier functions do not exist [173, 174]; this is at variance with nor-
mal insulators where localized Wannier functions exist, and can even be chosen
as exponentially localized [19, 109, 175].

88
Figure 7.1: Discretization
of the reciprocal cell. The
periodic gauge is enforced
on the boundary; otherwise
the gauge is unspecified and
possibly erratic. The Chern
number is the sum of the
Berry phases computed on
the small squares.

7.1.1 Computing the Chern number


The computation of a Chern number proceeds similarly to what presented in
Sec. 4.1.4 for a toy-model Hamiltonian. Suppose we discretize the reciprocal
cell with a regular mesh, as in Fig. 7.1. We start enforcing the periodic gauge
|ujk+G i = eiGr |ujk i on the boundary, i.e. for all the couples of boundary
points which are related by a reciprocal lattice vector. At all the interior points
the gauge is chosen by the diagonalization routine and is therefore erratic. The
mesh actually shown in Fig. 7.1 would actually require 8 8 = 64 Hamiltonian
diagonalizations.
We recall that the curvature is the Berry phase per unit reciprocal area, and
we compute it on each of the small squares using the four-point discrete formula

= Im log huk1 |uk2 ihuk2 |uk3 ihuk3 |uk4 ihuk4 |uk1 i, (7.3)

which applies to the single-band case. In the many-band case this is replaced
by
= Im log det S(k1 , k2 )S(k2 , k3 )S(k3 , k4 )S(k4 , k1 ), (7.4)
where S is the overlap matrix

Snn (ks , ks ) = hunks |unks i; (7.5)

see Eq. (3.46). One has to choose the Im log branch with in [, ].
The Chern number C1 is the sum of all the s, covering the whole reciprocal
cell (64 in Fig. 7.1). In our discontinuous approach, the obstruction is not
located at any particular k point, although its effect becomes apparent after all
s are summed. we remind thatas discussed in Sec. 4.1.4the discretised
Chern number is always an exact integer, and not an approximate integer, even
for a coarse mesh. This feature has been apparently first pointed out in 2002
[176].

7.1.2 Chern number as the curvature for unit area


The Berry curvature (k) has the dimensions of a squared length; as said
above, it has the meaning of Berry phase per unit reciprocal area. Here instead
we consider a possibly disordered system, where the k vector is devoid of any
meaning. For a system which is macroscopically homogeneous, we are going to
define a global curvature , which is a numbernot a function of kand is
extensive, scaling like the system size (or like the electron number). The Chern

89
number is then proportional to our global curvature divided by the system area
in r space.
We start rewriting Eq. (7.2) for a square lattice as
/L /L
1
Z Z
C1 = dkx dky (k), (7.6)
2 /L /L

where L is the lattice constant. As usual, a noncrystalline system can be dealt


with in a supercell framework, by addressing an artificial crystal with a large
L, actually larger than the relevant correlation length. In this limit the BZ
becomes very small; owing to the mean value theorem
 2
1 2
C1 (0). (7.7)
2 L

We identify (0) with the (extensive) global curvature ; if A = L2 is the


system area, then

C1 = 2 (7.8)
A
in the thermodynamic limit.
At this point, it looks like we still need to deal with a k vector since, within
the the usual definition of (0), its essential ingredients are k derivatives of the
|unk i evaluated at k = 0. Instead, if we work within our discrete Berryology,
such derivatives are not needed. Incidentally, this shows once more that discrete
Berryology is more general and more powerful than differential Berryology. The
value of the global curvature can be obtained from a single Hamiltonian
diagonalization for a large supercelleither crystalline or disorderedand no
derivative is actually involved.
This bears some relationship to the so-called single-point Berry phase, dis-
cussed in Sec. 5.3.5. Here we follow the approach of Ref. [177]: to give a flavor
of how it works, we start from Fig. 7.1, where one needs 64 independent diag-
onalizations, as said above. One may look at the same system by doubling the
elementary cell in each direction, in which case the number of occupied bands
increases fourfold, and the new reciprocal cell has one quarter of the original
area. Now the calculation of requires 64/4 = 16 independent diagonaliza-
tions; the algorithm of Eq. (7.4) yields an value identical to previous one (up
to computer accuracy), because we are folding the same orbitals in a smaller
reciprocal cell (or BZ). We can go on with the folding, diagonalizing in 64, 16,
4, 1 points. We still get a numerically identical result insofar as no disorder
is introduced. If the supercell size L is large enough, the value computed
via a single Hamiltonian diagonalization and (super)lattice-periodical orbitals
is very close to the thermodynamic limit = C1 A/(2), Eq. (7.8), with C1
integer [177]. Once the large supercell is adopted, disorder can be introduced at
no extra cost; since C1 is a robust topological invariant, weak disorder cannot
change the value.
What we have presented here is an r space interpretation of the Chern num-
ber, Eq. (7.8), which gets rid of k space and deals with crystalline and disordered
systems on the same footing. Nonetheless, the approach of the present Section
requires a macroscopically homogeneous system and the adoption of PBCs. Is
it possible to overcome these limitations, and address topological order even

90
Figure 7.2: Upon dou-
bling the elementary cell,
the reciprocal-cell area is di-
vided by 4, and the number
of occupied bands is multi-
plied by 4.

for inhomogeneous systems (e.g. heterojuncions) and/or adopting OBCs? The


answer is yes; the presentation of this is deferred to Sec. 7.5, after we have in-
troduced the reference computational material (Haldanium) for dealing with
this kind of issues.

7.1.3 Chern number and hermaphrodite orbitals


The hermaphrodite orbitals have been defined and discussed in Sec. 6.5.2. Here
we address the simple case of a 2d square lattice of constant a and a single
occupied band. We simplify notations by indicating with ky (x, y) the central-
cell (ky -dependent) orbital; the complete set in the occupied manifold is, after
Eq. (6.51), m ky (x, y) = ky (x ma, y), where m Z and ky [0, 2/a). The
inverse transformation yields the Bloch orbitals as
X
k (r) = eimkx a ky (x ma, y) (7.9)
m
X
uk (r) = eikx (xma) ky (x ma, y), ky (x, y) = eiky y ky (x, y).
m

The uk orbitals are clearly lattice-periodical and normalized over the unit cell;
their k derivative yields
X
1 uk (r) = i eikx (xma) (x ma) ky (x ma, y),
m
X
2 uk (r) = eikx (xm a) k (x m a, y),
ky y
m
X
h1 uk |2 uk i = i eikx ma hm ky |(x ma)| 0 ky i, (7.10)
m
ky

where h i means r integral in (, ) (0, a). The Berry curvature and the
Chern number are therefore

(k) = 2 Im h1 uk |2 uk i
X
= 2 Re eikx ma hm ky |(x ma)| 0 ky i,
m
ky
Z 2/a Z 2/a X
1
C1 = Re dkx dky eikx ma hm ky |(x ma)| 0 ky i
0 0 m
ky
2 2/a
Z
= dky Re hky |x| k i, (7.11)
a 0 ky y

91
where |ky i |0 ky i is the central-cell orbital. If we define x0 as its (ky
dependent) center, then

x0 (ky ) = hky |x|ky i,


dx0
= 2 Re hky |x| k i (7.12)
dky ky y
2/a
1 dx0 1
Z
C1 = dky = [ x0 (2/a) x0 (0) ]. (7.13)
a 0 dky a
The hermaphrodite orbitals carry therefore a very clear topological signature.
In a topologically trivial 2d insulator their center is periodical in ky ; while
instead in the nontrivial case their center shifts by C1 a when ky ky + 2/a.
More generally, not only the center but even the function m ky (x, y) enjoys an
analogous property.
As for the k space periodicity of the Bloch functions, Eqs. (7.9) and (7.13)
yield

kx+2/a,ky (r) = kx ,ky (r)


kx ,ky+2/a (r) = eiC1 kx a kx ,ky (r). (7.14)

this confirms that in the Chern case the gauge cannot be periodical over the
reciprocal cell. The gauge implicit in Eq. (7.9) can be called cylindrical:
periodic in kx but not in ky .

7.1.4 Lowest Landau level


We consider a system of noninteracting electrons in a flat substrate potential and
in a normal magnetic field. Although the system is translationally invariant in
the (xy) plane, both the Hamiltonian and the density matrix P are not invariant.
Despite this, the orbitals of e.g. the lowest Landau level (LLL) share many of
the properties of the hermaphrodite orbitals.
If the LLL is fully occupied (i.e. at filling = 1) the electron density is uni-
form and equal to n0 = 1/(22 ), where = (~c/eB)1/2 is the magnetic length;
the modulus of the density matrix has the translationally invariant expression
(for single occupancy)
2
/(42 )
|P (r, r )| = n0 e|rr | , (7.15)

and the cumulant second moment is clearly finite. The trace hx2 ic + hy 2 ic of
the localization tensor is in fact equal to 2 , the squared magnetic length, and
the minimum quadratic spread in the x direction is 2 /2.
The LLL orbitals in the Landau gauge are labeled by a one-dimensional wave
vector q (, ) and have the form

q (r) = eiqy (x q2 ), (7.16)

where (x) is the normalized ground eigenfunction of a 1d harmonic oscillator


with frequency c = eB/me c
 1/4
1 2
/(22 )
(x) = ex . (7.17)
2

92
The quadratic spread of this function is indeed 2 /2; the q (r) are therefore
eigenstates of P xP , and are the magnetic analogue of the maximally localized
hermaphrodite orbitals. Notice that in Eq. (7.16) q (r) has a plane-wave-like
normalization over y.
To make contact
with our notations for hermaphrodite orbitals it is enough
to define a = 2 and

q = ky + 2m/a, ky [0, 2/a), m Z


1 i2my/a
mky (x, y) = e (x ky 2 2m2 /a)
a
1 i2my/a
= e (x ky a2 /(2) ma); (7.18)
a
as in the previous Section, the orbitals are normalized over (, ) (0, a).
Finally, we verify the behavior of vis-a-vis the Chern number. The center
of the central cell orbital is x0 = ky a2 /(2); if ky is increased by 2/a, then
x0 increases by a, as it must be: we remind that C1 = 1 in the LLL (sign
conventions are not uniform across the literature).

7.2 Quantum Hall insulators: Correlated wave-


function
The integer quantum Hall (QH) effect has been discussed at length above. The
most perspicuous featuresee Fig. 2.7is that whenever the 2d electron fluid
is in the QH regime: (i) the transverse resistivity is quantized; and (ii) the
longitudinal resistivity vanishes. The same two features are common to the
fractional QH effect as well. We remind that, while the integer QH effect can
be understood at the independent-particle level, a correlated wave function is
essential in the fractional case. We also remind that in the QH regime the static
longitudinal conductivity vanishes: this is why we speak of QH insulators,
both in the integer and fractional cases.
We are going to show that electron localizationas defined in the modern
theory of the insulating stateis the common cause for both the vanishing of
the dc longitudinal conductivity and the quantization of the transverse one.
Therefore to predict whether the dc transverse conductivity is quantized, it is
enough to inspect electron localization in the ground state. This outstanding
result was proved in 2005 [139].
Our system is a square of size L L, where the presence of a macroscopic B
field forbids simple PBCs; this, however, is no serious problem. We choose the
Landau gauge for the magnetic vector potential in Eq. (6.1), and we impose the
magnetic boundary conditions which are customary in the QH literature [178];
this requires the total flux = BL2 to be an integer times 0 , where 0 = hc/e
is the flux quantum.
When switching from 3d to 2d, many of the above formulae require obvious
modifications. For an isotropic system, the 2d analogue of Eqs. (6.30) and (6.31)
can be written as
Z
d e2 N
Re 11 () = Re hx2 ic , (7.19)
0 ~L2

93
2e2 N
Re 12 (0) = Im hxyic . (7.20)
~L2
A glance at the sum-over-states expression for hr r ic , Eq. (6.10), shows that
whenever Re hr r ic converges, then Im hr r ic converges as well. However,
we are going to show that the latter may only converge to quantized values in
the present 2d case.
By definition we have
1
Im hxyic = Im h1 0 (0)|2 0 (0)i. (7.21)
N
In the large-L limit, we may write

Im h1 0 (0)|2 0 (0)i (7.22)


 2 Z 2 Z 2
L L L
= Im d1 d2 h1 0 ()|2 0 ()i.
2 0 0

The dimensionless integral is over a closed surface (a torus), given the magnetic
boundary conditions. It is therefore equal to C1 , where C1 is the Chern
number of the first class (Sect. 3.4); hence

L2
Im hxyic = C1 , (7.23)
4N
and the transverse conductivity, Eq. (7.20) becomes

e2 e2
Re 12 (0) = C1 = C1 . (7.24)
2~ h
We have thus arrived at the famous Niu-Thouless-Wu formula [178] which holds
for both the integer and fractional QH effect. The original derivation was based
on an analysis of the Green function, under the hypothesis that the system
has a Fermi gap; in the present derivation the presence of a Fermi gap be-
comes apparently irrelevant, since the localization tensor is a pure ground-state
property. Underlying the modern theory of the insulating state, however, is a
fluctuation-dissipation theorem, relating the ground state to the excitations of
the system.
The main message of Ref. [139] (and of the present Section) can be summa-
rized by saying that any 2d insulator displays a quantized transverse conduc-
tance (nonvanishing only in absence of time-reversal symmetry).
An electron fluid is kept in the QH regime by disorder, and analytical imple-
mentations of the present formulae are obviously not possible. Nonetheless, in
order to illustrate how the theory works, it is expedient to consider the academic
case of noninteracting electrons in a flat substrate potential. As shown above, if
the first Landau level is fully occupied the trace hx2 ic + hy 2 ic of the localization
tensor is finite and equal to the square of the magnetic length = (~c/eB)1/2 .
If the filling is fractional the density cannot be uniform (for noninteracting elec-
trons), and, more important, the cumulant second moment diverges (see Ref.
[139] for details): the system is not insulating, and its transverse conductivity
is not quantized.

94
Figure 7.3: Four unit cells of crystalline Halda-
nium [172]. Filled (open) circles denote sites with
E0 = (+). Solid lines connecting near-
est neighbors indicate a real hopping amplitude t1 ;
dashed arrows pointing to a second-neighbor site in-
dicates a complex hopping amplitude t2 ei . Arrows
indicate sign of the phase for second-neighbor hop-
ping.

7.3 Topological insulators


The QH state of matter, discovered in 1980 and addressed above, provides the
first example of a quantum state which is topologically distinct from all pre-
viously known states of matter. This is the reason why its macroscopic quan-
tization properties are very robust (topologically protected), and insensitive
to small changes in material parameters. The debut of topological concepts in
electronic structure theory dates from two milestone papers: TKNN in 1982 [50]
and Niu-Thouless-Wu in 1985 [178]. These papers addressed a 2d electron fluid
(noninteracting and interacting, respectively) in the presence of a macroscopic
magnetic field: the Hamiltonian, therefore, cannot be lattice periodical. Topo-
logical insulators are defined as materials whose ground state is topologically
nontrivial in absence of an applied magnetic field.

7.3.1 Haldanium
The archetype of all topological insulators is the Haldane model Hamilto-
nian [172] (Haldanium): its hallmark is quantized Hall conductance in ab-
sence of a macroscopic magnetic field. Haldanium is the paradigm for Chern
insulators, and has been used as a workhorse in many simulations, provid-
ing invaluable insight into topological features of the electronic wavefunction
[172, 177, 173, 179, 180, 181], as well as into features of orbital magnetization
[182, 183, 177, 184].
The model is comprised of a 2d honeycomb lattice with two tight-binding
sites per primitive cell with site energies , real first-neighbor hopping t1 , and
complex second-neighbor hopping t2 ei , as shown in Fig. 7.3. Within this two-
band model, one deals with insulators by taking the lowest band as occupied.
The appeal of the model is that the vector potential and the Hamiltonian are
lattice periodical and the single-particle orbitals have the usual Bloch form.
Essentially, the microscopic magnetic field can be thought as staggered (i.e. up
and down in different regions of the cell), but its cell average vanishes.
For a 2d lattice-periodical Hamiltonian the Chern number C1 has been de-
fined before, Eq. (7.2). As a function of the flux parameter , this system
undergoes a transition from a normal insulator (C1 = 0) to Chern insulator
(|C1 | = 1). Its phase diagram is shown in Fig. 7.4. On the vertical = 0 axis
the Hamiltonian is time-reversal invariant and Haldanium becomes a model for
hexagonal boron nitride; the center of the plot ( = 0, = 0) corresponds
to graphene. Notice also that on the horizontal = 0 axis Haldanium is a
nonpolar and centrosymmetric insulator (except at = 0).
Like in the QH case discussed in Sect. 7.2: (i) the imaginary part of the

95
localization tensor is related to the Chern number by Eq. (7.23); and (ii) the
trace of the localization tensor is finite. This is confirmed by the simulations of
Ref. [173], where the actual value of I , Eq. (6.52), is computed. We remind
that, despite I being finite, localized Wannier functions (with finite quadratic
spread) do not exist in Chern insulators.
Haldanium and QH insulators display quantized transverse conductivity;
all are localized in the sense of the modern theory of the insulating state. It is
worth pointing out, though, that the decay of the density matrix is qualitatively
different: exponential in Haldanium [173] vs. Gaussian in the (noninteracting)
QH case, Eq. (7.15).
Chern insulators remained a curiosity of academic interest only for many
years; no actual material belonging to this class was actually synthesized until
2013. The hallmark of a 2d Chern insulator is quantum anomalous Hall effect
(QAHE). The effect was first observed in 2013 in Cr-doped (B,Sb)2 Te3 thin
films [185]; the quantization was later observed with much higher precision in
V-doped (B,Sb)2 Te3 -doped (B,Sb)2 Te3 [186].
In 2005 (well before the experimental discovery of a Chern insulator) it was
realized that the Haldane model is the forebear of a completely new class of
topological insulators, the time-reversal symmetric ones [51, 10, 11].

7.3.2 Time-reversal symmetric topological insulators


A novel kind of topological insulators in both 2d and 3d has been theoretically
predicted and experimentally discovered since 2005 [10, 11, 52, 53, 54, 55, 56]; in
these materials the macroscopic quantization properties are topologically pro-
tected, and insensitive to small changes in material parameters, analogously the
Chern (a.k.a. AQHE) insulators.
The time-reversal invariant topological insulators are characterized by a Z2
invariant (while the Chern number is a Z invariant in 2d). The paradigmatic
Z2 topological insulator is the Kane-Mele model Hamiltonian in 2d [51]. Most
important, at variance with Chern insulators, the modern topological insulators
are time-reversal symmetric. We have met before (Secs. 4.2.3 and 5.5) the group
Z2 : the additive group of the integers modulo 2. It has only two elements, which
can be labeled as 0 and 1, or as even and odd: topological insulators are
Z2 -odd, while a trivial insulator is Z2 -even.
Since spin-orbit interaction plays a major role, the definition of localization
tensor needs to be augmented to explicitly include the spin coordinates; this

Figure 7.4: Chern number C of the


bottom band of the Haldane model
as a function of the parameters
and /t2 ; the plot is drawn for
t1 = 1 and t2 = 1/3. At = 0
Haldanium is a model for graphene
and hexagonal boron nitride.

96
seems to be rather straightforward. After this is done, the scope of the modern
theory of the insulating state would include topological insulators. A closely
related investigation concerning the Wannier functions in a Z2 insulator has
appeared in 2011 [187] . The results confirm that the ground state of a topo-
logical insulator is localized in the sense of the modern theory of the insulating
state: I , Eq. (6.52), is in fact finite. Furthermore, at variance with the Chern-
insulator case: (1) localized WFs (with finite quadratic spread) do exist [187],
and (2) a smooth gauge over the whole BZ (with no obstruction) exists [188].
Both localized WFs and the smooth gauge requires breaking the Kramers-pair
symmetry in the Z2 -odd case.
We are not giving in these Notes a microscopic definition of the Z2 invariant,
and we refer to the (nowadays vaste) literature on the topic. Only one interesting
property of Z2 topological insulators in 3d is discussed below, Ch. 9.

7.4 Digression: Bulk-boundary correspondence


As discussed throughout these Notes, geometrical and topological properties of a
given system manifest themselves in reciprocal space. The concepts of reciprocal
space and of k vector are based on PBCs: the sample has no boundaries by
construction. However, for the same physical system, one could consider a
bounded sample within OBCs. In this case there is no k vector to speak of:
the geometrical properties manifest themselves as surface (in 3d) or edge (in
2d) properties. Bulk-boundary correspondence is the hallmark of geometry and
topology in electronic structure.
The archetype of bulk-boundary correspondence is the IQHE. When a
toroidal geometry is adopted there is no boundary, and the fingerprint of non-
trivial topology is the Chern number. If instead the 2d electron fluid in the
quantum Hall regime has the shape of a ribbon, then there are edge states which
carry a chiral current. However, these edge states are a bulk property and are
topologically protected. Similar considerations apply to Chern insulators in
2d.
For 2d time-reversal invariant topological insulators the edge states cannot
carry any net current; instead there are counterpropagating spin currents which,
once more, are topologically protected: one speaks then of quantum spin Hall
effect. In the 3d case, there are topologically protected surface states, whose
nature is nontrivial.
The bulk-boundary correspondence also occurs in the case of polarization,
where the topologically protected boundary quantity is not current but charge.
Here one has to pay attention to the fact that polarization is the sum of an
electronic and a ionic term: only the former term is expressed as a k inte-
gral. However, the bulk polarization determines the surface charge modulo the
quantum (Sec. 5.3.4). In the simple case where the bulk is centrosymmetric,
topology requires the charge per surface cell to be an integer or an half integer
(see also Sec. 5.5). We stress that, quite often, a polar surface is metallic and
topology does not apply.

97
7.5 Topological order in r space
The Chern number, as well as other topological invariants, labels the manner
in which the electronic ground state is topologically twisted or knotted in
k-space [10, 11]. But topological order must reflect a peculiar organization of
the electrons even when the concept of k-space does not apply, such as for
inhomogeneous systems (e.g. heterojunctions), as well as for finite systems
within OBCs. In Ref. [181] it has been shown that the Chern number in a
2d insulator can be mapped by means of a topological marker, defined in r
space, and which may vary in different regions of the same sample. Notably,
this applies equally well to periodic and open boundary conditions.
We start rewriting Eq. (7.1) in 2d for a single band:
Acell
Z
Im hr r ic = dk Im h uk | uk i, (7.25)
(2)2 BZ
whence the Chern number is related to the imaginary part of the localization
tensor as
4 4
C1 = Im hr r ic = Im Trcell{P xQy}, (7.26)
Acell Acell
where the second equality owes to Eq. (6.38).
At this point it looks like that we can use Eq. (7.26) even for a finite 2d
flake within OBCs, replacing the trace over the cell with the trace over the
whole system, and replacing Acell with the flake area. Instead, we have already
observed in Sec. 6.4.2 that the localization tensor is real within OBCs: this is
confirmed by noticing that
i
Im Tr{P xQy} = Im Tr{P xP y} = Tr{ [P xP, P yP ] }, (7.27)
2
and that the trace of the commutator is zero, whenever P projects on a finite-
dimensional manifold.
The key to the apparent paradox is that the thermodynamic limit and the
trace do not commute. In Eq. (7.26) the limit to the infinite system is performed
before taking the trace, while in Eq. (7.27) we are attempting at taking the trace
for a finite system, and eventually performing the limit afterwards.
The solution proposed in Ref. [181] is to address the commutator in
Eq. (7.27) before taking its trace. If we write
Z
Tr{ [P xP, P yP ] } = dr hr| [P xP, P yP ] |ri, (7.28)

then the integrand carries the information about the local topological order. A
topological marker is defined as the dimensionless function

C(r) = 2ihr| [P xP, P yP ] |ri; (7.29)

for a crystalline system, Eq. (7.26) guarantees that the cell average of C(r)
coincides with the Chern number. The function C(r) fluctuates over microscopic
dimensions; in the nonperiodc case, the cell average has to be replaced with the
macroscopic average hCimacro , defined as in electrostatics (see e.g. Jackson [90]).
We stress that the microscopic function C(r) itself, before macroscopic averaging,
bears no physical meaning.

98
The topological marker is validated over different Haldanium samples, by
means of simulations performed on finite flakes within OBCs. The samples are
crystalline, disordered, and inhomogeneous (heterojunctions), in both normal
and Chern regimes. The results are illustrated and commented in Ref [181].
The topological marker described here only addresses the Chern (i.e. Z)
topological order. A marker having similar properties and addressing instead
the Z2 topological order has not be found so far, and possibly cannot exist.

99
Chapter 8

Orbital magnetization

Like polarization P, even macroscopic magnetization M is a fundamental con-


cept that all undergraduates learn about in elementary courses [89, 90]. Macro-
scopic magnetization M is the sum of two terms, which are unambiguously
defined in nonrelativistic (and semirelativistic) quantum mechanics: spin mag-
netization M(spin) and orbital magnetization M(orb) . Experimentally, magneto-
mechanical measurements, based on the Einstein-de Haas effect, provide the
two terms separately. For instance, the values of M(spin) and M(orb) for the
three ferromagnetic metals (Fe, Co, and Ni) are accurately known since half
a century [189]. From the viewpoint of the present review, M(spin) is a dull
quantity. Electronic-structure codes routinely compute the spin density, which
is a simple lattice-periodical function. Its cell average (times a trivial factor)
coincides with M(spin) . In other words, a dipolar density is unambiguously
identified, while the same does not happen in the orbital case. Throughout
this Chapter we address orbital magnetization, using the symbol M to identify
M(orb) only.
Polarization P and orbital magnetization M share some analogy, but also
some significant differences. For the time being, we emphasize the main analogy:
textbooks define both P and M as the dipole moment of a sample, divided by
the volume V :
d 1 m 1
Z Z
(micro)
P= = dr r (r), M = = dr r j(micro) (r), (8.1)
V V V 2cV

where (micro) (r) and j(micro) (r) are the microscopic charge and (orbital) current
densities. Contrary to what most textbooks pretend, there is no such thing as a
dipolar density: the basic microscopic quantities are (micro)(r) and j(micro) (r)
[190]. If the sample is uniformly polarized/magnetized, then the microscopic
charge/current averages to zero in the bulk of the sample. Owing to the presence
of the unbound operator r, both P and M Eq. (8.1) are dominated by surface
contributions, while instead phenomenologically they are bulk properties.
The modern theory of magnetization dates from 2005-06 [183, 191], and is
still (2015) partly work on progress [191, 192, 193, 177, 194, 195, 184]. A couple
of reviews have appeared [8, 18], while only a few first-principle implementations
appeared so far (2015) [196, 197, 198].

100
8.1 Magnetization and magnetic field
The modern theory of magnetization only addresses the polarization M in a
null macroscopic B field; in this case M can be nonzero only if time-reversal
symmetry is broken in the spatial wavefunction. For instance, in a ferromag-
net the spin-orbit interaction transmits the symmetry breaking from the spin
degrees of freedom to the spatial (orbital) ones.
It must be realized that, insofar as we address an infinite system with no
boundaries, the B field is quite arbitrary, in full analogy with the case of E dis-
cussed in Sec. 5.1. The usual choice invariably performed within all electronic-
structure codes is to impose a lattice-periodical vector potential, i.e. B = 0.
When addressing a finite sample with boundaries, the B field is in principle
measurable inside the material, without reference to what happens at the sample
boundary: in fact, B obtains by averaging over a macroscopic length scale the
microscopic magnetic field B(micro) (r), which fluctuates at the atomic scale [90].
In a macroscopically homogeneous system the macroscopic field B is constant,
and in crystalline materials it coincides with the cell average of B(micro) (r).
As for the electrical case, discussed in Sec. 5.1, there is no need of addressing
finite samples and external vs. internal fields from a theoreticians viewpoint.
Nonetheless a brief digression is in order, given that experiments are performed
over finite samples, often in external fields. Suppose a finite macroscopic sample
is inserted in a constant external field B(ext) : the microscopic field B(micro)(r)
coincides with B(ext) far away from the sample, while it is different inside be-
cause of screening effects. If we choose an homogeneous sample of ellipsoidal
shape, then the macroscopic average of B(micro) (r), i.e. the macroscopic screened
field B, is constant in the bulk of the sample. The shape effects are embedded in
the demagnetization coefficients [89]: the simplest case is the extremely oblate
ellipsoid, i.e. a slab of a macroscopically homogeneous material; more details
are given in Ref. [8]. For the slab geometry in a vanishing external field B(ext)
the internal field B vanishes when M is normal to the slab (longitudinal polar-
ization), while B = 4M is the demagnetization field when M is parallel to
the slab (transverse polarization): see Fig. 8.1. Notice that this is the opposite
of what happens in the electrical case (Fig. 5.1).

Figure 8.1: Macroscopic magnetization M in a slab normal to z, for a vanishing


external field B(ext) . Left: When M is normal to the slab, no demagnetizing
field and no surface current is present. Right: When M is parallel to the slab,
a demagnetizing field B = 4M is present inside the slab, and dissipationless
currents Ksurface = c M n flow at the surfaces.

101
8.2 Magnetization itself
In Eq. (8.1) we have emphasized the main analogy between P and M; here,
instead, we anticipate the main difference. In the case of Pas stressed in Sec.
5.2the modern theory addresses the difference in polarization P between
two states of the material that can be connected by an adiabatic switching
process. As for P itself, the modern theory of polarization states that the
bulk electron distribution determines P only modulo a quantum, whose value
depends on the boundary of the sample [95, 119] (Sec. 5.3.2, and also Figs. 5.4
and 5.5). The modern theory of magnetization, instead, addresses M itself
directly, and is not affected by any quantum indeterminacy. We may express this
outstanding difference by saying that by tinkering with the boundary one can
change the polarization of a sample (by a quantum); while instead tinkering
with the boundary one cannot change the value of M.
Another key difference is that, while P makes sense only for insulators at zero
temperature, M is well defined even for metals and even for finite temperature.
In the present Notes we limit ourselves to discuss the zero-T insulating case,
both normal and Chern. We refer to the original papers [183, 191, 192], as well
as to the review of Ref. [8] for the metallic and finite-T cases.
The main concepts are more clearly formulated in the simple 2d case: elec-
trons in the xy plane and magnetization M along z; the 3d theory is not concep-
tually different, although it requires a more complex algebra. We only address
a system of spinless independent electrons. If the chemical potential is in the
bulk gap, the value of M is -independent in the normal case and -dependent
in the Chern case, owing to boundary currents. However, the boundary currents
are topologically protected (i.e. are a bulk effect): therefore in both cases the
M value is a unique function of the bulk electron distribution and of .
The macroscopic magnetization at fixed chemical potential is M =
(1/A) G/B, where A is the system area and G = U N is the grand
thermodynamic potential at T = 0:
1 U N
M = + = M1 + M2 . (8.2)
A B A B
We address the second term first: defining the areal density n = N/A, we get
M2 = n/B; in the thermodynamic limit we make contact with Stredas
formula [84], Eq. (4.60):
n eC1 C1
= = , (8.3)
B 2~c 0
where 0 = hc/e is the flux quantum.
In order to address the first term M1 in Eq. (8.2) we start considering a 2d
macroscopic flake of finite size: its energy is
X
U= j . (8.4)
j <

In the normal case (C1 = 0) the flake as a whole remains gapped in the large
system limit, while in the Chern case (C1 6= 0) the bulk is gapped but the
spectrum of the flake becomes gapless in the large system limit. Nonetheless,
for a given (finite) flake the number of terms in Eq. (8.4) is B-independent for

102
B 0. We may therefore invoke the Hellmann-Feynman theorem to get
U X H e X
= hj | |j i = hj | r v |j i; (8.5)
B <
B 2c <
j j

not surprisingly, M1 coincides then with the 2d analogue of Eq. (8.1). Using
v = i[H, r]/~ we have

ie X
M1 = hn | r [H, r] |n i (8.6)
2~cA <
j

ie X
= ( hn | xHy |n i hn | yHx |n i ).
2~cA <
j

Eq. (8.6) is a trace; since M1 is real,


e
M1 = Im iM1 = Im Tr {P xHyP }, (8.7)
~cA
where P is the ground-state projector, Eq. (6.35) (for single occupancy).
The trace in Eq. (8.7) has the same drawback as Eqs. (8.1) and (8.5): it is
dominated by boundary contributions. The key transformation comes next. We
write H = P HP + QHQ, where Q is the complementary projector, Eq. (6.37).
Then it is rather straightforward to transform Eq. (8.7) into
e
M1 = Im Tr {P xQHQyP QxP HP yQ}. (8.8)
~cA
A different derivation of the same expression is due to Souza and Vanderbilt
[193].

8.2.1 The operator P r Q


We have already observed in Sec. 6.4.1 that the nasty position operator r is
ill defined within PBCs, while instead it is well defined when sandwiched
between a P and a Q. Here we give more details.
First we show explicitly that P rQ commutes with a lattice translation R for
a crystalline system.
Z
hr + R| P rQ |r + Ri = dr P (r + R, r) r Q(r, r + R)
Z
= dr P (r + R, r + R) (r + R) Q(r + R, r + R)
Z
= hr| P rQ |r i + R dr P (r, r)Q(r, r ), (8.9)

and the last term obviously vanishes. Clearly, a similar property does not hold
for P rP ; instead

hr + R| P rP |r + Ri = hr| P rP |r i + R P (r, r ). (8.10)

103
Another useful property is that in a crystalline insulator the operator
hr| P rQ |r i decays quasi-exponentially for large values of |r r |:
Z

hr| P rQ |r i = dr P (r, r) r Q(r, r )
Z
= r P (r, r ) dr P (r, r) r P (r, r )
Z
= dr P (r, r) (r r) P (r, r ). (8.11)

The integral in Eq. (8.11) can be regarded as the overlap in the variable r of
two functions: the former is P (r, r), centered at r and exponentially vanishing
away from it; the latter is (r r) P (r, r ), centered at r and vanishing expo-
nentially (times a linear function) away from it. Such an overlap decays quasi-
exponentially (i.e. like an exponential times a polynomial) for |r r | .
With respect to the above two properties, P rQ behaves analogously to P .
Like for P (r, r ), in order to get the operator hr| P rQ |r i in the bulk of a sample
one can equivalently perform the thermodynamic limit within either OBCs or
PBCs. For a crystalline system P rQlike P itselfcan be equivalently ex-
pressed in terms of either localized orbitals or Bloch orbitals; we provide here
the explicit Bloch-orbital expression.
The P and Q projectors are, for a 2d crystal with Nc (singly) occupied bands
Nc Z
Acell X
P = dk |nk ihnk |,
(2)2 n=1 BZ

Acell X
Z
Q = dk |n k ihn k |. (8.12)
(2)2
n =Nc +1

Then, by some manipulations which are standard in linear-response theory [126],


we have

hn k | r |nk i = ihun k |k unk i n 6= n


2
(2)
hn k | r |nk i = i (k k ) hun k |k unk i. (8.13)
Acell
We stress once more that, while the position operator r is ill-defined within
PBCs [97], its off-diagonal elements over the Hamiltonian eigenstates are well
defined; more accurately, Eq. (8.13) should be interpreted as a definition of such
elements. Therefore
Nc
iAcell X X Z
QrP = dk |n k ihun k |k unk ihnk |
(2)2 n=1 BZ
n =Nc +1
Nc
iAcell X X Z
P rQ = dk |nk ihk unk |un k ihn k |. (8.14)
(2)2 n=1 n =Nc +1 BZ

8.2.2 Magnetization in k space


Let us start from a finite flake, as addressed above, whose magnetization is
given in Eq. (8.8). If the flake is crystalline, cut from the bulk, and if its

104
linear dimensions are much larger than the quasi-exponential decay length of
hr| P rQ |r i, then the genuine bulk value of M1 is retrieved from
e
M1 = Im Trcell {P xQHQyP QxP HP yQ}, (8.15)
~cAcell
where the trace is taken over the central cell in the crystalline flake. At this
point we only need the thermodynamic limit of P rQ, which is conveniently
expressed in terms of Bloch orbitals, as shown in Sec. 8.2.1:
Nc
Acell X X Z
P xQHQyP = dk
(2)2 n=1 BZ
n =Nc +1
|nk ih1 unk |un k i n k hun k |2 unk ihnk |
Nc
Acell X X Z
QxP HP yQ = dk
(2)2 n=1 BZ
n =Nc +1
|n k ihun k |1 unk i nk h2 unk |un k ihn k | (8.16)

Nc
Acell X X Z
Im Trcell {P xQHQyP QxP HP yQ} = dk
(2)2 n=1 BZ
n =Nc +1
[ n k h1 unk |un k ihun k |2 unk i nk h2 unk |un k ihun k |1 unk i ].

At this point we may replace the sum over the empty states with the sum over
all states, since the added contribution is real. Therefore
e
Z
M1 = Im dk h1 unk | (Hk + nk ) |2 unk i, (8.17)
~c(2)2 BZ

where Hk is the effective Hamiltonian acting on the us, i.e. Hk = eikr Heikr .
In order to get the total magnetization we still have to add the second term
M2 in Eq. (8.2). Using Stredas formula, Eq. (8.3), and the k-space expression
for the Chern number, Eq. (7.2), we get the full k-space expression for orbital
magnetization
e
Z
M = M1 + M2 = Im dk h1 unk | (Hk + nk 2) |2 unk i. (8.18)
~c(2)2 BZ

This is the ultimate formula for orbital magnetization at zero temperature. The
finite temperature formula is in Ref. [192], and is also reviewed in Refs. [8, 18].
The proof of Eq. (8.18) provided here is original and yet (2013) unpublished.
The formula applies to normal insulators, to Chern insulators, andwithout
proof hereto metals. We conjecture that the present proof can be extended
to the metallic case as well, although the ground state projector P (r, r ) is
only polynomial in |r r | and not exponential. Previous proofs were either
semiclassical [83] or limited to normal insulators [182, 183, 177]. In the Chern
insulating case Eq. (8.18) was heuristically guessed and numerically validated
in Ref. [183].

105
Figure 8.2: Convergence of M with
the flake size, normal insulating
phase. Filled circles: total mag-
netic moment divided by the flake
area, Eq. (8.7). Open circles: trace
over the central cell, Eq. (8.15). All
flakes have the same aspect ratio;
the abscissae indicate the length L
of one rectangle side in lattice pa-
rameter units. After Ref. [184].

8.2.3 Magnetization in r space


The basic expression for the magnetization of a finite flake, Eq. (8.8), has been
used so far to derive the previously known k space expression, Eq. (8.18); it is
remarkable that the same expression can be exploited directly in r space [184].
We start discussing a normal (zero Chern number) insulator first: in this
case the term M1 , Eqs. (8.8) and (8.15), is the only contribution to M . In order
to illustrate how the local approach works, we choose Haldanium samples with
/t2 = 3.67, = 0.1; with reference to the phase diagram of Fig. 7.4, the
material is clearly in the nontopological phase. Its macroscopic magnetization
M = M1 , Eq. (8.8), has been computed for rectangular Haldanium flakes of
various sizes and fixed aspect ratio within OBCs [184]; the trace over the cell in
Eq. (8.15) is just the average over the two central sites.
The result, shown in Fig. 8.2 is remarkable: when comparing two alterna-
tive definitions of macroscopic magnetization, namely the elementary one (total
dipole over sample area) vs. our Eq. (8.15), the latter converges much faster to
the bulk limit. In fact, the nearsightedness (i.e. exponential decay [158]) of
the density matrix implies an exponential convergence of Eq. (8.15), while the
elementary definition only yields a 1/L convergence.
Next we notice that the local approach in r space is much more general
than the k space one, since it can be applied to cases where the concept of k
vector does not make any sense, such as noncrystalline and/or inhomogeneous
materials (e.g. heterojunctions). Similarly to what we did in Sec. 7.5, we define
the microscopic function
e
M1 (r) = Im ( hr| P xQHQyP |ri hr| QxP HP yQ |ri ) . (8.19)
~c
In the crystalline case, our previous Eq. (8.15) is just the cell average of M(r);
this can be clearly generalized by replacing the cell average with the macroscopic
average hM1 imacro , defined as in electrostatics (see e.g. Jackson [90]).
Next, we address Chern insulators. Therein the spectrum of a finite sample
within OBCs becomes gapless in the large sample limit; when is in the bulk
gap, the bulk is insulating but M depends on , owing to boundary currents.
This is quite clear from Eqs. (8.2) and (8.3); since, as shown in Sec. 7.5, even
the Chern number admits a local description, we may write both contributions
M1 and M2 to magnetization as

M = hM1 imacro + h C imacro , (8.20)
0

106
where h C imacro is the macroscopic average of our topological marker, Eq. (7.29).
An equivalent definition is M = hMimacro , where
e
M(r) = Im ( hr| P xQHQyP |ri hr| QxP ( H 2 )P yQ |ri ) ; (8.21)
~c
this is clearly reminiscent of the k space expression, Eq. (8.18). Eqs. (8.20)
and (8.21) are due to Bianco and Resta [184]; an equivalent expression was
independently found by Schulz-Baldes and Teufel [195].
We stress that the macroscopic average may vary over a macroscopic scale;
at a given point r inside the sample it is uniquely defined by the behavior of
M(r) in a microscopic neighborhood of r, and is insensitive to the conditions of
the sample boundary.
At the beginning of this chapter we have investigated the concept of dipolar
density, showing that it is well defined for spin magnetization but ill defined
for orbital magnetization. Our local function M(r) plays indeed the role of a
magnetic dipolar density; we stress, however, that only its macroscopic average
bears a gauge-invariant physical meaning.

107
Chapter 9

Chern-Simons geometric phase

As discussed in Appendix A, and particularly in A.6, the orbital magnetoelectric


response of a crystal includes a pseudoscalar term of geometrical nature, also
dubbed axion term. This can be written as
1 e2 1 e2
CS = = , (9.1)
4 2 ~c 2 hc
where is an angle [199, 200, 201], and CS stays for Chern-Simons. The
literature often adopts SI units, where the magnetoelectric susceptibility is no
longer dimensionless; in this case
e2
CS = (SI units). (9.2)
2h
The angle is a bulk property defined modulo 2; therefore CS is a bulk prop-
erty with a quantum of arbitrariness, whose value (in Gaussian units, adopted
here) is 1/2 times the fine structure constant. When we address a bounded
crystallite, the value of CS is no longer ambiguous once the surface termina-
tion of the 3d sample has been specified. The analogy of CS with macroscopic
polarizationand of with the corresponding Berry phaseis evident.
The Chern-Simons geometric phasea.k.a. axion phase is defined only
in 3d. The analogy with polarization is most evident if we consider the latter
in 1d only; for the sake of completeness we will also illustrate analogies and
differences with the Chern number of a 2d crystal.
All three geometrical/topological properties that will be addressed in this
Chapter have a reciprocal space expression as a Brillouin-zone integral in 1d
(polarization), 2d (Chern number), 3d (Chern-Simons). All of them can be
also formulated for a bounded crystallite in r space, by means of expressions
involving the ground state projector P . In order to establish a common language
and common notations, the BZ integrands will be invariably expressed in terms
of the nonAbelian Berry connection of the band insulator
A,jj (k) = ihujk | uj k i, (9.3)
which for any Cartesian coordinate is an Hermitian matrix in the band indices
jj . In the following we will omit the band indices: the symbol A (k) will
indicate the Hermitian matrix. The symbol tr indicates the matrix trace over
the band indices, while instead the symbol Tr indicates the trace over the
Hilbert space.

108
9.1 Polarization revisited
We have shown in Ch. 5 that the electronic contribution to the polarization of
a 1d system can be written in the form

Px = e , (9.4)
2
where is the Berry phase. Notice that here we define for double occupancy,
i.e. inluding the factor of two. Polarization is a well defined observable only
for charge-neutral systems. The Berry phase can be generalized to include the
nuclear contribution as well: Eq. (9.4) defines the total polarization if we set
Z
2
P (n)
=2 dk tr [Ax (k)] + Im log ei a m Zm xm . (9.5)
BZ

(n)
In Eq. (9.5) Zm are the dimensionless nuclear charges, and xm the nuclear co-
ordinates in the 1d unit cell of size a. Thanks to charge neutrality, Eq. (9.5) in
invariant by translation of the origin (each of the two terms is not such). Inter-
estingly, mathematicians refer to the trace of Ax (k), the integrand in Eq. (9.5),
as to a Chern-Simons 1-form (we have not adopted such semantics so far).
Insofar as we adopt PBCs the system is unbounded and Px is well defined
only modulo e. If we consider instead a finite realization of the same 1d Hamil-
tonian with length L, its OBCs dipole dx divided by L assumesin the large-L
limitone of the quantized values dictated by the bulk. If we adopt Eq. (9.4)
even in the OBC case, the dimensionless polarization can be written as
" #
2dx 2 1X (n)
= = 2 Tr {xP } + Zm xm , (9.6)
eL L L m

where P is the ground-state projector and Tr is the trace over the Hilbert
space.
In the special case where the bulk of the 1d system is centrosymmetric, the
phase becomes topological, as thoroughly discussed in Sec. 5.5; the same
happens to in the large-L limit. Centrosymmetric systems in 1d admit a Z2
classification: Z2 -even means = 0 mod 2, Z2 -odd means = mod 2,

9.2 Chern number revisited


The Chern invariant has been thoroughly discussed from Sec 7.1 onwards. In
2d it is the Chern (integer) number of the first class, which we write here for a
many-band crystal as
Z Z
2C1 = i dk tr [x Ay (k) y Ax (k)] = i dk tr [ A (k)]. (9.7)
BZ BZ

This is written for spin channel, or for spinless electrons, as everywhere in


these Notes. In terms of the nonAbelian Berry curvature matrix (k) the
expression is Z
2C1 = i dk tr [xy (k)]. (9.8)
BZ

109
Notice that in the nonAbelian case

(k) = A (k) A (k) i[A (k), A (k)], (9.9)

but the commutator does not contribute to the trace in Eq. (9.7), which therefore
looks the same in either the Abelian or nonAbelian case. We anticipate that,
instead, the trace of the Chern-Simons 3-formEq. (9.14) belowis different
in the two cases.
The right-hand member integral in Eq. (9.7) can be interpreted as an angle,
equal to an integer number of 2. The form of Eq. (9.7) emphasizes the analo-
gies, but also the differences, with the polarization phase angle , Eq. (9.5). The
invariant has no modulo arbitrariness, hence it provides a Z (not Z2 ) topo-
logical classification; it can be nonzero only in crystals which lack time-reversal
symmetry.
The r-space form of C1 has been discussed in Sec. 7.5; we are going to
cast those results in a form which emphasizes, once more the analogies and
differences with polarization as expressed in Eq. (9.6). We notice that the
electronic contribution to is a trace over the whole system, divided by the
total length. If we tryby analogyto cast even C1 as trace over a bounded
crystallite, divided by its area A, we get invariably a null result. Whenever
OBCs are adopted, the trace per unit volume (or per unit cell) has to be taken
in the bulk region of the sample. We indicate this trace as TrV and we cast
the main result of Sec. 7.5 as

2C1 = 4 2 Im TrV {r P r P } (9.10)

9.3 Axion angle


We need to restore spinful electrons here, because spin-orbit coupling is essential.
The nonAbelian connection matrix, Eq. (9.3), is then defined in terms of two-
component spinors (a.k.a. spinorbitals). The full density matrix is therefore
a projector, at variance with the one entering Eq. (9.5), which projected on a
single spin channel (doubly occupied).
The axiona.k.a. Chern-Simonsangle has a rather simple expression in
r space for a bounded 3d crystallite within OBCs. According to the literature
[200] its expression in terms of the ground-state projector is

4 2
= Im Tr {r P r P r P }, (9.11)
3V
where V is the crystallite volume, in the large-V limit; a crucial requirement is
that the crystallite has an insulating boundary. A superficial look would hint
that the axion angle is just the trivial 3d generalization of the 2d Chern
angle, Eq. (9.10). This is definitely not the case, and the differences are more
outstanding than the analogies.
First of all, in Eq. (9.11) we must take the trace over the whole system, while
in Eq. (9.10) we must take the trace in the inner region of the system only. Sec-
ondly, Eq. (9.10) has no modulo indetermination and is therefore independent
of the actual crystallite termination; Eq. (9.11) instead assumes different values
according to how the crystallite is terminated at its surface. Therefore shares

110
a couple of important properties with the polarization angle in 1d, Eq. (9.6): it
is a trace over the whole system, and depends on the sample termination.
We may look closer at the three r space formulas: Eq. (9.6) in 1d, Eq. (9.10)
in 2d, and Eq. (9.11) in 3d, we notice the occurrence of the ground state projector
P once, twice, and thrice, respectively. This is the key to best understand
differences and analogies.
We start with Eq. (9.10); using P 2 = P and Q = I P we recast it as

2C1 = 4 2 Im TrV {P r P r P } = 4 2 Im TrV {P r Qr P }. (9.12)

According to the discussion in Sec. 8.2.1 the operator P r Qr P is bounded and


lattice periodical, and this is the major virtue making C1 boundary insensitive,
with no modulo arbitrariness.
A similar manipulation does not help when the number of P operators is
odd: in Eq. (9.6) the unbounded operator xP cannot be made bounded anyhow.
In Eq. (9.11), replacement of only one of the P operators gives some clue about
the 1d-3d analogy. In fact

4 2
= Im Tr {r (P r Qr P ) }, (9.13)
3V
where we have exploited antisymmetry. The inserted parenthesis isolateby
the associative propertythe operator P r Qr P , which is bounded and lattice
periodical, as said above. But this is multiplied by the unbounded operator r
before the trace is taken: the drawback is therefore not much different from the
case of 1d polarization.
The k space expression for in terms of the nonAbelian Berry connection
looks formidable, but is in fact a known formula in differential geometry. The
BZ integrand is the nonAbelian Chern-Simons 3-form [199, 200]:
 
1 2i
Z
= dk tr A (k) A (k) A (k)A (k)A (k) . (9.14)
4 BZ 3

Mathematicians have a very compact expression for such forms: on Wikipedia


you find the general expression for the Chern-Simons 1-form, 3-form, 5-form,
2n + 1-form.
In the Abelian case the connection matrices are diagonal and real: hence the
second term in the trace vanishes. In the simplest Abelian case (i.e. two-band)
the Chern-Simons geometric phase, Eq. (9.14), has a very simple expression in
terms of the connection A(k) (having two spin components) written in vector
form:
1 1
Z Z
= dk A(k) k A(k) = dk A(k) (k), (9.15)
4 BZ 4 BZ

where (k) is the 3d Berry curvature written in vector form as well.

9.4 Z2 topological insulators


The Chern-Simons term in the free energy is proportional to E B, ergo odd
under time-reversal (T ) and under inversion. This implies = if the crystal

111
has any of the above symmetries. Notice also that in order to have a genuine
magnetoelectric response both symmetries must be broken (Appendix A).
Since in an unbounded insulating sample is only defined modulo 2, this
means that both = 0 and = are possible in T -invariant crystals (re-
gardless of inversion symmetry) and in centrosymmetric crystals (regardless of
T symmetry). In both cases all linear-response (Kubo) terms accounting for
magnetoelectric coupling vanish, and the angle becomes topological: a Z2 in-
variant. This is in analogy with in 1d centrosymmetric insulators, mentioned
above and thoroughly discussed in Sec. 5.5.
The invariant equals for crystals in the class of strong Z2 topological
insulators, and therefore Eq. (9.14) provides an alternative way of detecting
topological order. Despite the difficulties in discretizing Eq. (9.14)illustrated
in Sec. 9.5 belowthe value of = has been numerically verified for Be2 Se3
[201], which is a paradigmatic material: its Z2 -odd character has been assessed
via other theoretical tools, and experimentally as well.
Notice that = , according to Eq. (9.1), would yield a magnetoelectric
susceptibility equal (in Gaussian units) to 1/4 times the fine structure constant,
i.e. 6 104 , a rather large value compared to common magnetoelectrics.
The value addressed so far is evaluated as a BZ integral, Eq. (9.14), for an
unbounded sample. Next we ideally address a bounded crystallite cut from a
Z2 -odd crystal, where the bulk T symmetry is preserved at the surface as well. If
the whole surface were insulating, the arguments given above would guarantee a
nonvanishing (and large) magnetoelectric susceptibility. This is clearly absurd:
any T -invariant finite system has a null magnetoelectric susceptibility (all terms
of it, not only the Chern-Simons term). The solution is obvious: the surface
must be metallic. It is known by independent arguments, in fact, that Z2 -
odd T -invariant insulators have topologically protected metallic states at their
surfaces.

9.5 Numerical considerations


The analytical BZ integrals for 1d polarization, 2d Chern number, and 3d axion
angle as given aboveEqs. (9.5), (9.8), and (9.14)all require twice differen-
tiable Bloch orbitals |ujk i throughout the Brillouin zone. Owing to differen-
tiability Eqs. (9.5) and (9.14) are gauge-invariant modulo 2, while Eq. (9.8)
is gauge-invariant with no ambiguity. We also observe that whenever C1 6= 0
the Berry connection has one or more obstructions, while nonetheless the
curvature entering Eq. (9.8) is everywhere continuous.
In numerical work the BZ integrals are discretized over a k-point mesh; the
orbitals |ujk i are then not a differentiable function of k. The phase factors are
erratic, and further unitary mixing of the occupied states at a given k is usual.
One therefore needs replacing Eqs. (9.5), (9.8), and (9.14) with their discretized
counterpart, which must converge to the respective analytical expression in the
dense-k limit.
In the present Notes we have described in detail the discretization of the
Berry phase (Sec. 3.9) and of the Chern number C1 (Sec. 7.1.1): in
both cases the expressions are numerically gauge invariant. By this we mean
that the expressions can be directly implemented with the raw Bloch states
(as provided by Hamiltonian diagonalizations) where the gauge is erratic. A

112
naive discretizationbased on replacing the k-derivatives in the integrands in
Eqs. (9.5) and (9.8) with finite differencesfails because the raw |ujk i are not
differentiable. The discretizations discussed in Secs. 3.9 and 7.1.1 follow a
different path and are unaffected by the erratic gauge.
In the case of the Chern-Simons angle a numerically gauge-invariant dis-
cretization has not be found so far (except in the Abelian case). Therefore
the gauge must be regularized before the BZ integral, Eq. (9.14), is discretized
[201, 202].

113
Appendix A

Magnetoelectrics (basic features)

A.1 Generalities
Let us consider first a normal, i.e. non magnetoelectric, material; we as-
sume it to be in general anisotropic, but macroscopically homogeneous. The
electromagnetic free-energy density is
1 1
F (E, H) = E E H H, (A.1)
8 8

where is the macroscopic dielectric tensor (same as in Appendix B), and
is the magnetic permeability tensor. We draw attention to the minus signs,
explained e.g. in Ref [89].
Eq. (A.1) is nicely symmetric andsince we are adopting Gaussian units
the fields have the same dimensions. This however hides an important feature:
electric and magnetic energies in condensed matter are not of the same order
of magnitude. The ratio of the magnetic energy scale to the electric one is of
the order of 104 . This ratio owes in fact to the actual value of the squared
fine-structure constant: (1/137)2.
As we will see in more detail below, magnetoelectric materials realize a cou-
pling between electric and magnetic phenomena. The conversion of magnetic
signals into electrical ones (and conversely) is obviously of the utmost techno-
logical interest. Nowadays, most devices exploit the phenomenon of the giant
magnetoresistance (Nobel prize in 2007), which has generated an economy of
billion of dollars. To date, the magnetoelectric effect is not commercially viable,
due to the fact that in all known materials the coupling is very weak, owing
again to the (1/137)2 bottleneck. An active area of research concerns therefore
mechanisms and materials where the effect would be enhanced.

A.2 B vs. H fields


In this Appendix we focus the free energy F (E, H), where E and H are cho-
sen as the independent variables. As for the electric field, this is clearly the
natural choice. We have stressed that E (not D) is the internal field, which is
measurable in principle inside the material. The macroscopic field E is also a
control parameter in first-principle calculations: standard crystalline Hamilto-
nians adopt a lattice-periodical selfconsistent potential, hence E = 0. Matters

114
are different in the magnetic case: the internal field, measurable in principle
inside the material is B, not H. Also, if a selfconsistent Hamiltonian is lattice-
periodical, the macroscopic B field, not H, is zero. And in fact the modern
theory of magnetization, discussed here in Ch. 8, addresses magnetization in
zero B field.
So, why instead we are adopting here H, and not B, as the independent
variable? There are several reasons. Phenomenologically, the experimenter
directly controls E (e.g. via capacitors) and H (via e.g. solenoids or generally
currents). In the laboratory, you will hear people speaking of E and H, more
often than of D and B, the reason being that E and H are directly read on
the instruments. Several textbooks even call H the magnetic field, which
I find strongly misleading. I adopt the nomenclature of the good textbooks,
such as Feynman [203] and Griffiths [204] popular textbooks, where B is called
magnetic field, and H is just H.
In the framework of magnetoelectric effects there are further reasons for
using H as the independent variable. All formulasincluding Eq. (A.1)have
a pretty symmetric expression in terms of E and H; Feynman warns however
that although the equations are analogous, the physics is not analogous [203].
The equations would look asymmetrical and ugly using the genuine magnetic
field B instead.
Custom dictates that spontaneous magnetization is defined as M0 =
4F /H: this is not what is addressed in Ch. 8. However B and H are

related by the magnetic permeability tensor as B = H, and in normal ma-
terials the permeability difference from one (in Gaussian units) is of the order
of (1/137)2 . There is then little difference between M0 as addressed in Ch.
8 and M0 as customarily addressed. Whenever needed, the conversion of the
response tensors from their (E, H) definition to the (E, B) counterpart requires
only straightforward algebra. The (E, B) must be adopted in first-principle
calculations, as indeed in Ch. 8; it is also adopted in Sec. A.6 below.

A.3 Multiferroics
The first order expansion of Eq. (A.1) reads

F (E, H) = F0 P0 E M0 H + . . . , (A.2)

where P0 is the spontaneous polarization and M0 is the spontaneous magneti-


zation. The materials where both are nonzero are called multiferroics: ferroelec-
tric and ferromagnetic at the same time. In such materials the electric-magnetic
coupling is expected to be strong.
There are very few multiferroic materials: ferroelectricity and ferromag-
netism seem to be almost mutually exclusive in nature. The reasons for this
have been investigated [205].

115
A.4 Linear magnetoelectrics
Here we address only crystalline materials where both P0 and M0 vanish. The
most general second order expansion of the free energy reads
1 1 1
F (E, H) = F0 E E H H E H, (A.3)
8 8 4
where textbooks ignore the last term here, and adopt Eq. (A.1) instead.
In high-energy physics a particle related to a B E coupling term added to
the electromagnetic Lagrangian has been postulated in the 1970s and dubbed
axion (see Sec. A.6). However, it is was known since much earlier that mag-
netoelectric coupling does occur in condensed matter. In 1960 Dzyaloshinskii

[206] realized that the magnetoelectric coupling tensor is in general nonzero in
crystals where both inversion symmetry and time-reversal symmetry are absent.
In fact E is odd under inversion and even under time-reversal, while the oppo-
site happens for B: therefore their combination in the last term of Eq. (A.3) is
odd under each of the two transformation.
From Eq. (A.3) the conjugate variables are

F
D = E + 4P = 4 = E+ H;
E
F
B = H + 4M = 4 = E+ H. (A.4)
H
Therefore in magnetoelectrics an H field induces an electrical polarization at
zero E field, and conversely an E field induces magnetization at zero H.
In its original paper, Dzyaloshinskii [206] proposed to look for the magneto-
electric effect in Cr2 O3 , which is non centrosymmetric and antiferromagnetic.
Shortly afterwards, Soviet physicists indeed experimentally demonstrated both
the direct and the converse magnetoelectric effect in Cr2 O3 [207]. In this mate-
rial the effect is by far too weak for being of technological use. There has been
a resurgence of interest in the magnetoelectric effect; the search for materials
where the effect could be stronger has been rather active in the last 15 years.

A.5 Parsing the magnetoelectric effect



In a non magnetoelectric cystalwhere = 0the only field coupled to lattice
coordinates is E: a magnetic field does not exert any force on the nuclei at

rest. The effects of lattice-field coupling on the dielectric tensor and on the
zone-center vibrational modes are analyzed in detail in Appendix D.

Whenever instead 6= 0 both fields E and H are coupled to lattice displace-
ments, formally on equal footing. As a consequence all three material constants

, , have an electronic contribution and a lattice contribution. The lattice-
field coupling in magnetoelectrics has been investigated in the literature: the
results presented in Appendix D for simple dielectrics have been generalised to
linear magnetoelectrics. We do not present such generalisations here, and we
refer to the original literature [208, 209, 210, 211].
Next, we consider the purely electronicalso called clamped nuclei re-
sponse functions. Since the macroscopic magnetisation M is the sum of

116
spin magnetization and orbital magnetization, Eq. (A.4) shows that both the

clamped-nuclei and have a spin and an orbital contributions. The full

susceptibility tensor of the paradigmatic magnetoelectric material Cr2 O3 has
been computed in 2012 [212]: the spin and lattice contributions are separately
addressed, and in both cases the response is decomposed into lattice and elec-
tronic parts.

A.6 Axion term


Here we adopt the (E, B) choice, according to the discussion at the end of Sec.
A.2. Since second order energies depend on first order wavefunctions, one needs
the linear response induced in the Bloch orbitals by either E or B, and then
uses them to evaluate the induced magnetization or polarization, respectively:
both approaches are feasible [213].
A very interesting geometrical feature contributes to the clamped-nuclei or-
bital term. Besides the genuine linear-response (a.k.a. Kubo) quantities, this
response includes a ground-state term, to be evaluated with the Bloch states of
the unperturbed crystal. This contribution to the magnetoelectic susceptibility

is scalar (more precisely a pseudoscalar), hence it provides a contribution


to the induced magnetization parallel to E, or equivalently a contribution to
polarization parallel to B.
Within PBCs this contribution is a novel geometrical quantity, discussed
above in Ch. 9, and given by the Brillouin-zone integral of a Chern-Simons 3-
form, Eq. (9.14) [199, 200, 201, 202]. The peculiar feature is that this geometrical
term is multivalued and thus reminiscent of the Berry-phase expression for the
spontaneous polarization: it remains ambiguous until the surface termination of
the 3d sample has been specified. The multivalued nature of the Chern-Simons
term is made explicit by writing it in terms of an angle as:

1 e2
CS = . (A.5)
4 2 ~c
The term axion comes from high-energy physics and was coined by F.
Wilczek in 1977-78. It is based on writing the Lagrangian for the electromagnetic
field as
1 1 2 1 2
L = + j A + E B + E B, (A.6)
c 8 8 4
where the postulated axion term is the last one. The equation of motion is
unaffected if is constant in space-time. Without the axion term the standard
L is a relativistic scalar in 4d space-time; the axion term is pseudoscalar, and
would account for the (very rare) processes which break charge-parity symmetry.

117
Appendix B

Fundamentals of piezoelectricity

B.1 Generalities
Piezoelectricity was first demonstrated in 1880 by Pierre and Jacques Curie.
Nowadays it is the most common effect industrially used to convert electrical
signals into mechanical ones, and conversely. Therefore the occurrence of the
piezoelectric effect in everydays life is ubiquitous. We make use of literally
thousands of piezoelectric sensors and actuators in our cars, computers....
Piezoelectricity is symmetry forbidden in centrosymmetric crystals. The
most symmetrical crystal structure where piezoelectricity is allowed is
zincblende: therein, uniaxial strain along the (1,1,1) direction is coupled to a
macroscopic E field along the same direction. Therefore not all piezoelectric ma-
terials are pyroelectric or ferroelectric, but the converse is true: all pyroelectric
and ferroelectric materials are also piezoelectric. A clarification about seman-
tics is in order: both pyroelectric and ferroelectric crystals have a preferred axis,
along which the spontaneous polarization P0 is oriented. In ferroelectrics P0 is
switchable (without crossing a metallic state), while in pyroelectrics switching
is impossible: breaking and reforming of covalent bonds would be needed. The
simplest pyroelectric symmetry classi.e. where P0 6= 0 is allowedis wurtzite
(see Fig. 5.2 and the discussion about it), while the prototypical ferroelectric is
BaTiO3 , whose crystal structure is a cubic perovskite undergoing spontaneous
inversion-symmetry breaking.
A wurtzite material much in fashion nowadays is GaN (see Nobel prize 2014).
Although the industrial interest in this material is about light emission and not
about the piezoelectric effect, its piezoelectricity has been thoroughly investi-
gated [214].
Most piezoelectrics in industrial and commercial use are ferroelectric ma-
terials, but they are generally ceramics, not crystalline. A new class of useful
materials has been discovered in 1997 [215]: they are single-crystal solid-solution
ferroelectrics. Some of these new materials have found their way into industrial
applications, e.g. in medical echographs.
For reasons explained in Sec. 5.1 (and in the previous Appendix) the most
convenient electrical variable in condensed matter is the field E (not D). We
remind that, when E is chosen as independent variable, the macroscopic free
energy density in isotropic dielectric media is F = E 2 /(8) in Gaussian units.
As for Eq. (A.1), we draw attention to the minus sign [89]. The conjugate field

118
obtains as D = 4F /E.

Here we are going to introduce macroscopic strain as an additional inde-
pendent variable besides E, and we address anisotropic media. Given that we
are going to address strain and volume changes, it is better to switch from the
free energy density F to the free energy per cell F = Vcell F , where Vcell is the
equilibrium volume.
It is expedient to use compact notations, where the Cartesian indices of the
relevant tensors (of various ranks) are left implicit. The macroscopic material
tensors discussed in this Appendix are genuinely static ones, i.e. include the
lattice contribution. We anticipate that, instead, in Appendix C the lattice
contribution to the dielectric constant (or tensor) will be separately investigated.

B.2 First order properties


The first order expansion of the free energy per cell is

F F
F ( , E) = Vcell F0 + + E + ...
E 0

0

= F0 Vcell 0 Vcell P0 E + ..., (B.1)

where P0 is the spontaneous polarization in zero field. Since we are by defini-


tion expanding around the equilibrium crystal structure, the macroscopic stress

tensor 0 in Eq. (B.1) vanishes.
However, the definition of equilibrium deserves a clarification. We remind
that E is the internal (screened) field inside the material. According to our
discussion in Sec. 5.1, whenever P0 6= 0 the field E is in general nonzero
inside a finite sample of arbitrary shape in zero external field. Therefore a slab,
cut parallel to P0 , and in zero external field, is in equilibrium and unstrained.
Insteadaccording to our expansion in Eq. (B.1)for a slab normal to P0 and
in zero external field we may consider two cases: (1) If ideally the structure
is kept unrelaxed the stress in nonzero (and proportional to P0 ); (2) In the
relaxed structure the stress is zero, but the equilibrium strain is nonzero (and
proportional to P0 ). This will appear more clearly when the second order
expansion is considered.

B.3 Second order properties



Setting 0 = 0 the expansion up to second order is

F ( , E) F0 Vcell P0 E (B.2)

1 2F 2F 1 2 F

+
+ E+ E E,

2 2 E E 0
0
E
0

where in the second derivatives we identify the elastic constants C (4th rank

tensor), piezoelectric tensor e (3rd rank tensor), and dielectric tensor (2nd

119
rank tensor), respectively. More precisely:

2 F 2 F 4 2 F

1 1
C = , e = , = .
Vcell
Vcell E Vcell E E 0
0 0
(B.3)
We therefore rewrite the second order expansion as

F ( , E) F0 Vcell P0 E
Vcell V
+ C Vcell E e cell E E, (B.4)
2 2
and we remind thatas stressed aboveall material constants entering

Eq. (B.4) are defined at = 0 and E = 0.
The conjugate variables are

= C +e E (B.5)

D = e + E + 4P0 ; (B.6)

here the dagger indicates the transpose, and e is called the converse piezo-
electric tensor. Eq. (B.6) can be recast as
1
P P0 = e + ( 1 ) E. (B.7)
4
Therefore the piezoelectric tensor can be defined as either the stress linearly
induced by a unit field at zero strain (converse), or the polarization linearly
induced by unit strain at zero field (direct).

B.4 Open circuit vs. closed circuit


In Fig. B.1 we show two different ideal measurements of a piezoelectric coeffi-
cient. The crystal is uniaxially strained along one of its piezoelectric axes.
In the left (short-circuit) figure one measures the time-integrated current
flowing through the sample, and we remind that the integrated current is the

+ + + + + +

Figure B.1: Ideal measurements of the piezoelectric effect. Left: closed circuit.
Right: open circuit.

120
polarization difference. The field E is zero at all times, hence according to
Eq. (B.7) the measurement directly provides the vertical component of e. We
pause at this point to remind that, as stressed in Sec. 5.2, the modern theory
of polarization directly evaluates the integrated current at zero field. Since
currents are easer to measure than charges, the ideal experiment in the left
figure is indeed close to the real experiments.
In the right (closed-circuit) figure one aims at measuring the dipole of the
slab, or equivalently the strain-induced surface charge. In this setting we have
D = 0, not E = 0, therefore for P normal to the slab E = 4P. Replacing
this into Eq. (B.7) one gets

P = ( )1 (e +P0 ). (B.8)

In the simple case where P0 = 0 (e.g. in zincblende crystals) Eq. (B.8) shows
that the open-circuit and closed-circuits measurements are simply related via
the dielectric tensor: this is no small difference, since dielectric constants in
common dielectrics are of the order 510.
The case with P0 6= 0, i.e. pyroelectric and ferroelectric materials, deserves
further discussion. Let us assume throughout that even P0 is normal to the

slab. At zero strain we set = 0 and we find that P is related to P0 by the
inverse dielectric tensor. This is in full agreement with our discussion in Sec.
5.1 about longitudinal vs. transverse polarization; see also Fig. 5.1. For a more
complete treatment see Ref. [8]. The open-circuit piezoelectric tensor is given
therefore by

P = ( )1 (e +P0 ) ( )1 P0 = ( )1 e . (B.9)

The formula looks therefore the same as for the P0 = 0 case, but a final subtlety
is worth pointing out.
The free energy expansion in Eq. (B.4) is around the equilibrium geometry
in zero field, and in fact according to Eq. (B.5) both strain and stress are zero
when E = 0. But in a slab where P0 is normal to the slab E is nonzero, as
thoroughly discussed in Sec. 5.1. Hence in a free standing slab in zero external
field, and zero stress, the equilibrium strain is nonzero. In fact, from Eqs. (B.5)
and (B.6), the equilibrium obtains by solving the linear system

0 = C +e E (B.10)

0 = e + E + 4P0 , (B.11)

which provides the values of E (depolarisation field) and (equilibrium strain


in a longitudinal geometry). Therefore the open-circuit response in Eq. (B.9) is
indeed the response to the additional strain besides the equilibrium strain.

B.5 Piezoelectricity from first principles


So far we have given a macroscopic phenomenological treatment of the linear
piezoelectric effect; the underlying microscopic concepts are nowadays rooted in
the modern theory of polarization. It is interesting, nonetheless, to address the
theory of piezoelectricity within an historical perspective.

121
Piezoelectricity has been for years an intriguing problem. The formal proof
that piezoelectricity is a well defined bulk propertyindependent of surface
terminationis due to R. M. Martin [216], in a paper appeared in 1972 under
the lapidary title of Piezoelectricity (no subtitle!). This proof has been chal-
lenged, and the debate lasted for about 20 years [217]. The very first quantum-
mechanical calculation ever of a piezoelectric tensor was published in 1989 [218].
Therein, the converse definition of Eq. (B.5) was adopted, i.e. the computed
quantity was the stress linearly induced by a unit field in III-V semiconductors.
Since the calculation was in terms of Bloch orbitals within PBCs, Ref. [218]
provided further evidence (if any was needed) that piezoelectricity is indeed a
bulk effect.
After the advent of the modern theory of polarization in the early 1990s it
became possible to implement even the direct definition of Eq. (B.7), i.e. to
compute the strain-induced linear polarization as a numerical derivative, via the
modern Berry-phase approach.

122
Appendix C

Linear-response theory

C.1 Definitions
To start with, we fix our conventions about Fourier transforms:
Z Z
it 1
f () = dt e f (t) f (t) = d eit f (). (C.1)
2

Suppose we have a general input signal finput (t) and the corresponding output
foutput (t), which is due to the response of a time-independent physical system.
The most general linear response is given by a convolution
Z
foutput (t) = dt (t t )finput(t ). (C.2)

The response , also known as generalised susceptibility, can be equivalently


written as the functional derivative
foutput (t)
(t t ) = , (C.3)
finput (t )

evaluated at equilibrium. The response is therefore an equilibrium property of


the system in absence of perturbation. The convolution theorem yields

foutput () = () finput (). (C.4)

C.2 Causality
If the input signal is chosen as finput (t) = (t), then we have foutput (t) = (t):
in a causal system therefore (t) = 0 for t < 0.
We pause to notice that not all linear systems must be causal: an obvious
example of a noncausal system is an amplifier. An amplifier can start self
oscillations (Larsen effect), which quickly become anharmonic; however it needs
to get energy from some power source.
We also stress that causality and dissipation are not synonymous. While a
dissipative system is causal, one can address causal systems where no mechanism
allows for dissipation: the paradigmatic example is the (undamped) harmonic

123
oscillator in classical mechanics. A formal trick to account for causality is to
consider a dissipative system in the nondissipative limit.
The Fourier transform of a causal (t) is
Z

() = () + i () = dt eit (t). (C.5)
0

We only address a response (t) which is real in the time domain: therefore
() is even and () is odd. Exploiting this fact, the antitransform is
Z
1
(t) = d [ () cos t + () sin t ]. (C.6)
2
The two terms in the integrand cancel for t < 0, and provide an identical
contribution for t > 0:
1
Z
(t) = d () eit , t > 0. (C.7)

In order to prove this, we exploit the identity (t) = [(t) + (t)] (t), hence
Z
1
(t) = d ()[ eit + eit ], t > 0. (C.8)
2
Since () is odd, we immediately arrive at Eq. (C.7).

C.3 Kramers-Kronig relationships


The Kramers-Kronig relationships are most frequently proved by addressing
the Fourier transform of the response function in the complex plane. Here
we provide a less commonand to my taste more elegantproof, found in the
literature.
It is expedient to write
(t) (t) (t) (t)
(t) = + sgn(t) , (C.9)
2 2
where sgn is the signum function, and to extend the integral to (, ):
1
Z

i () = dt eit [(t) (t)] (C.10)
4
Z
1
() = dt eit sgn(t) [(t) (t)]. (C.11)
4
We notice then that the Fourier transform of sgn(t) is the distribution 2i/, un-
derstood as a principal part. Using then the convolution theorem in Eq. (C.11),
and exploiting Eq. (C.10) we have
2i i ( ) 1 ( )
Z Z
() = d
= d . (C.12)
2
The real and imaginary parts of () are related via an Hilbert transform. The
inverse relationship reads
1 ( )
Z

() = d . (C.13)

124
Finally, we may take advantage once more of the fact that () is even and
() is odd, arriving at the Kramers-Kronig relationships in the equivalent
form
2 ( )
Z

() = d 2
0 2
Z
2 ( )
() = d 2 . (C.14)
0 2
Finally, we notice a subtle feature occurring when considering response func-
tions which are related by a time derivative, i.e.

X (t) = (t). (C.15)

If X(t) is causal, then (t) is obviously causal. The example most relevant to
the present notes concerns electrical polarizability and conductivity . In fact
P = E, j = E, and j = dP/dt.
If X() obeys Kramers-Kronig, then even () = iX() also obeys.
Instead, if () is causal, one would say that X() = i()/ is causal, but
this is in general incorrect: one has to fix the integration constant in the time
domain. The correct causal response function is
 
i
X() = + () (), (C.16)

where the two terms in parenthesis are easily recognised as the Fourier transform
of the causal function (t) = [sgn(t) + 1]/2; the second term can be neglected
only when () vanishes at = 0.

C.4 Fluctuation-dissipation theorem (classical)


We address the response of a classical system and the fluctuations of a classical
variable at finite T ; the quantum theorem is discussed below. The simplified
presentation given in the following is inspired by Ch. 8 in Chandlers text-
book [219]. We work in the canonical ensemble, hence we are addressing the
isothermal response throughout the present Notes.
Let us assume that the extensive quantity of interest obtains as the trace
(integral over the 3N -dimensional phase space) of a classical dynamical variable
B(rN , pN ). In order to simplify notations, and without loss of generality, we
assume that the equilibrium value of B in the unperturbed system vanishes:
R N N
dr dp B eH
hBi = R N N H = 0, (C.17)
dr dp e

where H is the classical many-body Hamiltonian.


Suppose next that the system is perturbed at t = with a time-
independent perturbation H. After relaxation, the new equilibrium value
at t = 0 is R N N
dr dp B e(H+H)
B(0) = R N N (H+H) 6= 0. (C.18)
dr dp e

125
Since we are addressing linear response, we neglect terms quadratic in H and
we get R N N
dr dp H B eH
B(0) = R . (C.19)
drN dpN eH
At time t = 0 we suddenly switch off the perturbation H. The system is now
out of equilibrium, and will evolve towards it, from B(0) 6= 0 to B() = 0.
We then define B(t) = B(t; rN , pN ), where each phase-space point in B(t)
evolves from the initial value (rN , pN ); the evolution is governed by the unper-
turbed Hamiltonian H. We assume that B(t) is a well behaved function of the
t = 0 distribution in phase space. We thus have
R N N
dr dp H B(t) eH
B(t) = R t 0. (C.20)
drN dpN eH

Now we identify H with minus the dynamical variable A(rN , pN ) and we


assume for the sake of simplicity that even hAi = 0. We then rewrite Eq. (C.20)
as
R N N
dr dp A(0) B(t) eH
B(t) = R = hB(t)A(0)i t 0, (C.21)
drN dpN eH

where we have introduced the simplified notation for the time-correlation func-
tion hB(t)A(0)i. Notice that it is a measure of the equilibrium fluctuations of
the dynamical variables A and B in the unperturbed system. Notice that B(t)
and A(0) commute in the classical case.
As a function of time, the perturbation we have considered so far can be writ-
ten as finput (t)A, where finput(t) = (t) is the externally applied disturbance
coupled to the dynamical variable A. If we further identify B(t) = foutput (t),
from the basic Eq. (C.2) we have
Z 0 Z
foutput (t) = hB(t)A(0)i = dt (t t ) = dt (t ). (C.22)
t

We thus arrive at the final statement of the (classical) fluctuation-dissipation


theorem

(t) = 0,t<0
d
(t) = hB(t)A(0)i, t > 0. (C.23)
dt
The response (t) provides the averaged perturbed value of the dynamical vari-
able B at time t when the system Hamiltonian H is perturbed at t = 0 by an
instantaneous kick (t)A.
In the simple case where A depends on the coordinates only (and not on the
momenta), the instantaneous kick corresponds to apply a -like force at t = 0
on the system at equilibrium. It is then clear that at t = 0 the coordinates
are continuous, while the momenta undergo a discontinuous jump. The system
relaxes toward equilibrium, and (t) is the ensemble average of the dynamical
variable B as a function of time.
In general, owing to dissipation, the correlation functionsand (t)are ex-
pected to vanish exponentially for t ; there are exceptions, though. Even

126
in a simple system, like the hard-sphere fluid, the velocity-velocity autocorrela-
tion function has a power-law tail, called hydrodynamic tail. This important
discovery came from computer simulations performed in in 1970 by Alder and
Wainwright [220]. The analytical explanation came only after the computer
experiment [221], as it happens in the most creative episodes of computational
physics.
Integrating by parts, the Fourier-transformed response is
Z
() = hA(0)B(0)i + i dt eit hB(t)A(0)i. (C.24)
0

The generalised susceptibility has a real and an imaginary part, which obey by
construction the Kramers-Kronig relationships:
Z

() = hA(0)B(0)i dt hB(t)A(0)i sin t
0
Z
() = dt hB(t)A(0)i cos t. (C.25)
0

The fluctuation-dissipation theorem applies unchanged to the case A = B,


and the response is then given by the time autocorrelation function of the vari-
able A. In the static limit the generalized susceptibility is

0 = (0) = hA2 i, (C.26)

a result which can be directly proved in a simpler way.


The imaginary part is related to absorption, as will become clear in the
following (see Secs. D.2 and D.11). For a system at equilibrium the zero of
time is arbitrary, therefore hA(t)A(0)i = hA(0)A(t)i. Since the (classical)
dynamical variables are real and commute, the imaginary part of the response
can be written as

Z
() = dt eit hA(t)A(0)i, (C.27)
2

where the integral is often called power spectrum.


In the present Notes we only consider in some detail the case where the
extensive variable A is identified with the macroscopic dipole d of the system:
see Sec. C.5; see also Eq. (D.96) below and the following text.

C.5 Finite-temperature polarizability


For pedagogical reasons we start addressing a classical model of a fluid (either
liquid or gas) whose components are identical polar molecules. At the simplest
level, all of the polarization is due to molecular orientation: the molecules are
rigid and nonpolarizable. The paradigmatic system of this kind, used as a
workhorse in many molecular dynamics simulation, is the Stockmayer fluid:
therein the two-body interactions are Lennard-Jones supplemented with point
dipoles.
There is no ambiguity about what the polarization P and the total dipole d
are in a dipolar fluid: within either OBCs or PBCs the dipole is the sum of the

127
point dipoles: X
d=VP= dj . (C.28)
j

Instead, we have often stressed in the present Notes that in a general system
(such e.g. a polar crystal) dealing with macroscopic polarization within PBCs
is not so straightforward. The dipole d = V P of the simulation cell cannot be
decomposed in individual dipoles dj .
In presence of long range interactions the interpretation of the fluctuation-
dissipation theorem requires some care. In order to see the problem, we start
with a bounded and macroscopically homogeneous system within OBCs, and we
apply an external field E0 , uniform in space. We address the static case first.
The field couples to the Hamiltonian via the term d E0 . The dipole-dipole
autocorrelation function provides, from Eq. (C.26), the response
P 1 d
= = hd di. (C.29)
E0 V E0 3V
We aim to relate this to the static dielectric constant of the fluid, whose defini-
tion is
P
0 = 1 + 4 , (C.30)
E
where now E is the screened macroscopic field inside the polarized fluid.
The static dielectric constant 0 is a well defined bulk material property,
while instead the relationship between E0 and E depend on the sample shape
(see Sec. 5.1). A glance at Eq. (C.29) shows that hd2 i must depend on shape as
well: the physical origin of this is the depolarization field, which indeed affects
the equilibrium quadratic fluctuation. Equivalently, the fluctuations depend on
the boundary condition chosen for integrating Poisson equation in the selfcon-
sistent Newton equations of motion.

C.5.1 Zero-field boundary conditions


The simplest choice is to evaluate the fluctuations at zero macroscopic field:
this corresponds ideally to a finite sample coated with a metallic layer. In prac-
tice, one adopts instead PBCs and integrates Poisson equation by choosing the
solution which is periodic over the simulation cell. In the molecular-dynamics
jargon this choice goes under the name of Ewald-Kornfeld [222]. Within this
choiceas thoroughly discussed in Sec. 5.1 there is no depolarization field,
hence E = E0 .
Provided the equilibrium fluctuations are evaluated at zero macroscopic field,
the dielectric constant has the simple expression
4 2
0 = 1 + hd i. (C.31)
3V
We also display the expression for the imaginary part of the -dependent
dielectric constant, which is the basic quantity to address infrared absorp-
tion. When the dipole-dipole correlation functions are evaluated using periodic
boundary conditions, the previous results and Eq. (C.27) yield
2
Z

() = dt eit hd(t) d(0)i. (C.32)
3V

128
The example of a fluid whose components are rigid molecules was used here
for pedagogical purposes, but the classical fluctuation-dissipation theorem has
a much wider scope. Since the pioneering work of Silvestrelli et al. in 1997
[123] first-principle infrared spectra are routinely evaluated from the dipole-
dipole correlation function, where the polarization P = d/V is provided by the
modern theory of polarization, and its fluctuations are computed via a Car-
Parrinello simulation. More about this will be said below, Sec. D.11. For the
time being, we notice that P is due to both nuclei and electrons; in the adiabatic
approximation the nuclear coordinates are purely classical, but their coupling
to the field includes the electronic contribution in the form P E. The case of
a polar crystal in the harmonic regime, dealt with in detail in Appendix D, will
make this point pretty clear.

C.5.2 Reaction field


In any quantum description of a condensed system within PBCs the adoption of
the zero-field condition is almost mandatory. Only in this way, in fact, the elec-
trostatic potentialand hence the Schrodinger Hamiltonianis periodic over
the simulation cell. In classical mechanics, instead, the potential plays a lesser
role: the Newton equation of motions are directly formulated in terms of the
electric field, not its potential. Other choices of boundary conditions may there-
fore be appealing.
Following Ch. 2 of the Landau-Lifshitz textbook [89], we consider an ho-
mogenous sphere of dielectric constant 0 embedded in a medium of dielectric
constant RF (reaction field). When an external constant field E0 is present
in the surrounding dielectric the screened field E inside the sphere is [89]
3RF
E= E0 ; (C.33)
2RF + 0
hence
P 3RF P
= . (C.34)
E0 2RF + 0 E
Next we need the bare field coupled to the molecular Hamiltonian, i.e. the
field present in the empty spherical cavity. This easily obtains from the same
Eq. (C.33) by replacing 0 with 1. The perturbation to the Hamiltonian is
therefore
3RF
H = d E0 . (C.35)
2RF + 1
The fluctuation-dissipation formula, Eq. (C.29), must be replaced with
2 2RF + 1 P 2RF + 1 P
hd i = = . (C.36)
3V 3RF E0 2RF + 0 E
Using then Eq. (C.30) we arrive at the final result
4 2 (2RF + 1)(0 1)
hd i = . (C.37)
3V 2RF + 0
Setting RF = we retrieve Eq. (C.31). Another special case is RF = 0 ,
leading to the (1939) Kirkwood-Frohlich formula [223, 224]:
4 2 (20 + 1)(0 1)
hd i = . (C.38)
3V 30

129
The other interesting special case is RF = 1, yielding the time-honored (1912)
Debye formula
4 2 3(0 1)
hd i = .. (C.39)
3V 2 + 0
This ideally corresponds to a spherical sample free-standing in vacuo: in prac-
tice, the formula apples to simulations performed within PBCs, but when the
Coulomb interaction is truncated with a spherical cutoff. By contrast the zero-
field formula, Eq. (C.31), requires the evaluation of conditionally convergent
Ewald-like sums.
The dielectric constant is a well defined property of the bulk material: one
gets the same 0 value by adopting different boundary conditions, and using the
appropriate fluctuation formula for each of them. This has been first demon-
strated by Neumann in 1983 by numerical simulations on the Stockmayer fluid
[225].
It is pedagogically useful to provide the explicit expression for the depo-
larization field acting on the fluctuating unperturbed system. Starting from
Eq. (C.33), we replace 0 with its definition, Eq. (C.30); the result can be recast
as
3RF 4
E= E0 P. (C.40)
2RF + 1 2RF + 1
The first term in the rhs is the bare field acting on the molecular system, as in
Eq. (C.35), and the second term is the depolarization field. For the unperturbed
system at thermal equilibrium E0 = 0 and the depolarization field is
4
E= P. (C.41)
2RF + 1
The meaning of this is pretty clear: the fluctuating dipole of the molecular
system induces a depolarization field E within the sphere. Two special cases
are worth mentioning: E = 0 for RF = (metallic coating, no depolarization
field), and E = 4P/3 for RF = 1 (our sphere is free-standing in vacuo).
Some final words about the literature. The correct fluctuation formulas
for dipolar systems as well as for ionic systemsfirst appeared in the 1980s
[226, 225]. The original literatureas well as the more recent oneis plagued
by clumsy and inelegant algebra. The simple proof provided here is, to the best
of my knowledge, unpublished.

C.6 Quantum vs. classical


Before switching to linear-response theory and to fluctuation-dissipation in the
quantum domain, it is expedient to emphasize analogies and differences.
The first difference concerns dissipation. The great majority of the systems
dealt with in classical molecular dynamics is ergodic: there are exceptions, the
most famous of them being the Fermi-Pasta-Ulam Hamiltonian [227, 228]. This
was a great discovery made possible by computational physics; before Fermi-
Pasta-Ulam (1953) it was postulated by everybody (Fermi included!) that all
non separable Hamiltonians were ergodic in the thermodynamic limit. The
discovery opened a new branch of physics: nonlinear systems and the physics
of chaos [228].

130
Apart from such peculiar exceptions, normal systems are ergodic, evolve
towards thermal equilibrium, fulfill the equipartition theorem, & the like. This
is demonstrated since the birth of computational physics in the 1950s, where the
first simulations on hard-sphere and Lennard-Jones fluids appeared. Dissipa-
tion is therefore present by itself in a canonical simulation, and the fluctuation-
dissipation theorem directly provides the causal response functions. The same
applies to Car-Parrinello simulations, where the nuclei are indeed classical par-
ticles. Matters are different within quantum mechanics, where the introduction
of temperature and dissipation at the first-principle level is much more prob-
lematic.
In the present Notes we keep, as far as possible, the same notations adopted
in the classical case. Now the dynamical variables A and B, and the Hamilto-
nian H, are Hermitian operators. They do not commute, while their classical
counterparts do commute. The fluctuations addressed by the theorem are there-
fore fluctuations of the quantum variables A and B in the many-electron ground
state of the unperturbed system.

C.7 Fluctuation-dissipation theorem (quantum)


In this section we give the zero-T formulation in detail. The finite-T quantum
formula is presented at the end, Eq. (C.55), without proof.
We have seen above that (t) can be defined as the response in the observable
B to an instantaneous -like kick at t = 0, where the perturbing probe is
H(t) = (t)A. For t < 0 the system is in its (non degenerate) ground state;
in order to simplify notations we assumeas we have done in the classical
casethat the ground-state expectation value of both A and B vanish, i.e.

hAi = h0 |A|0 i = 0, hBi = h0 |B|0 i = 0, t < 0. (C.42)

The time-dependent Schrodinger equation is


d
H|(t)i (t)A|(t)i = i~ |(t)i, (C.43)
dt
and the initial condition is |(0 )i = |0 i. The temporal evolution of this state
is
i t
R
|(t)i = T e ~ 0 dt [H(t )A] |0 i, (C.44)
where T is the time-ordering operator; |(t)i has a discontinuous jump at t = 0,
which to linear order in A is
 
i i X
|(0+ )i = 1 + A |0 i = |0 i + |n ihn |A |0 i. (C.45)
~ ~
n6=0

This state evolves for t > 0 with the unperturbed Hamiltonian H:


i X iEn t/~
|(t)i = eiE0 t/~ |0 i + e |n ihn |A |0 i, t > 0. (C.46)
~
n6=0

Since we are at T =0, the response is causal but non dissipative: our (t) will
have undamped oscillations. Dissipation and lifetime can be introduced phe-
nomenologically by hand.

131
Finally, the linear response is

(t) = hB(t)i = (t) h(t)|B |(t)i (C.47)


i X
= (t) (ei0n t h0 |B |n ihn |A |0 i ei0n t h0 |A |n ihn |B |0 i),
~
n6=0

where 0n = (En E0 )/~. Its Fourier transform obtains from


Z
1 1
i dt eit (t) = P i() = lim , (C.48)
0+ + i

and adopting the compact notation due to Zubarev [67, 229, 230]:

1 X h0 |B|n ihn |A|0 i


hhB|Aii = lim+
~ 0 0n + i
n6=0
!
h0 |A|n ihn |B|0 i
. (C.49)
+ 0n + i

The Kubo formula for the generalized susceptibility takes then the form

() = hhB|Aii . (C.50)

We draw attention to the fact that the sign conventions adopted in these Notes
agree with Zubarev [67, 229] and Chandler [219], but are opposite the the ones
of McWeeny [230] and other textbooks. We have given here the many-body
Kubo formula; for independent electrons Eq. (C.49) is easily transformed into
a double sum over occupied and unoccupied orbitals.
The previous formulation is useful in order to explicitly provide the spectral
representation of (), Eq. (C.49) . Nonetheless a more compact and elegant
formula obtains by adopting the Heisenberg representation. In fact Eq. (C.47)
can be equivalently rewritten as
i i
(t) = (t) h(0+ )| e ~ Ht Be ~ Ht |(0+ )i = h(0+ )| BH (t) |(0+ )i
i
= (t) h0 | [BH (t), AH (0)] |0 i. (C.51)
~
It is easy to prove that h0 | [BH (tt ), AH (0)] |0 i = h0 | [BH (t), AH (t )] |0 i,
hence we may also write
i
(t t ) = (t t ) h0 | [BH (t), AH (t )] |0 i
~
i
= (t t ) Tr {P [BH (t), AH (t )] }, (C.52)
~
where P = |0 ih0 | is the ground state projector. Zubarev defines his correla-
tion function for complex frequencies z, i.e.
i
Z
hhB|Aiiz = dt eizt (t) Tr {P [BH (t), AH (t )] }, (C.53)
~
where the integral converges for z in upper half of the complex z plane. In fact
the spectral representation in Eq. (C.49) immediately shows that the response

132
is analytical in the upper half, and has poles in the lower half, infinitesimally
close to the real axis. This enforces causality, and suggests a simple picture.
For z = + i, and finite positive , the oscillations are (phenomenologically)
damped with a relaxation time = 1/. The response is therefore dissipative,
and the 0+ limit can be interpreted as the limit were the response is
undamped (as it must be at T = 0), yet causal.
The finite-T Zubarev correlation formula appears as a simple modification
of Eq. (C.53): it is enough to replace the ground-state projector therein with
the finite-T equilibrium density matrix in the canonical ensemble, i.e.

eH
P eq = ; (C.54)
Tr eH

i
Z
hhB|Aiiz = dt eizt (t) Tr {eq [BH (t), AH (t )] }. (C.55)
~

Despite the formal elegance and simplicity of this formula, its implementation
at the first-principle level is a formidable task.

133
Appendix D

Advanced lattice dynamics

D.1 Huangs phenomenological theory


Huangs theory, published in 1950 [231, 232], is the simplest paradigm for lattice
dynamics in presence of long-range forces. It addresses the zone-center opti-
cal modes of a cubic binary crystal. Zone-center modes are lattice-periodical,
therefore there is only one microscopic dynamical variable: the relative sublat-
tice displacement u. The system is therefore a single oscillator, isotropic at the
harmonic level, and where the electric field contributes to the restoring force.
We start from a free energy expression where the independent variable is the
field E, as in Appendices A and B, augmented here with the dynamical variable
u. The most general expansion to second order around equilibrium of the free
energy per cell F = Vcell F is
1 2 Vcell
F (E, u) = F0 + M TO u2 E 2 Z eu E. (D.1)
2 8
The reasons for the symbols adopted will be clear in a moment. But we pause to
stress that Eq. (D.1) is the most general second order expansion: we are mak-
ing no hypotheses about the physical origin of the three expansion coefficients.
Therefore all subsequent results are exact within the harmonic approximation,
and must hold for lattice-dynamical models, for first-principle calculations, and
for the experimental measurements as well.
If we identify M with the reduced mass of the two atoms, the three phe-
nomenological parameters are TO , , and Z (the dimensionless Born charge).
Notice that for Z = 0 the field and the lattice displacements are uncoupled.
The conjugate variables are the force f and the field D, i.e.
F 2
f = = M TO u + Z eE (D.2)
u
4 F 4
D = = E + Z eu. (D.3)
Vcell E Vcell
From the second line it is clear that is the dielectric constant when the nuclei
are kept clamped at their equilibrium position. This is the dielectric constant
actually measured at frequencies much higher than the phonon frequencies, but
lower than the electronic transition; it is sometimes named with the oxymoron

134
static high frequency [91]. The physical origin of is clearly the dielec-
tric response of the many-electron system: it is therefore also called electronic
dielectric constant.
At equilibrium we set f = 0, and we have
Z e
u = 2 E
M TO
4e2 (Z )2
 
D = + 2 E. (D.4)
Vcell M TO

The first line yields the sublattice displacement, while the expression in paren-
theses in the second line is the genuinely static dielectric constant 0 , including
the lattice relaxation, with 0 .
We now study forced oscillations: from Eqs. (D.2) and (D.3)

M 2 u = M TO
2
u + Z eE (D.5)

Z e
u= 2 E (D.6)
2)
M (TO
4e2 (Z )2
 
D = + 2 2 ) E = ()E. (D.7)
Vcell M (TO
The quantity in parentheses in the last expression is therefore the infrared di-
electric constant (), with (0) = 0 and () = .
We have not identified the meaning of the other parameters yet. To this
aim we notice that for a long-wavelength phonon of wave vector q the solid
is macroscopically homogenous normal to q, while its physical properties are
modulated in the q direction. It follows that E vanishes normal to q, while
D vanishes parallel to q. Setting E = 0 in Eq. (D.6) confirms that the TO
parameter in the free-energy expansion is indeed the zone-center transverse op-
tical frequency, while instead D = 0 when () = 0. The (squared) longitudinal
optical frequency LO is therefore, from Eq. (D.7)

2 2 4e2 (Z )2
LO = TO + , (D.8)
Vcell M
with LO TO . We have already observed that the Born charge Z measures
the coupling of the macroscopic field to the ionic motion. This coupling vanishes
in nonpolar binary crystals, such as those having the diamond structure, where
in fact no longitudinal-transverse (LT) splitting of zone-center optical phonons
is observed.
The zone-center longitudinal oscillations in a polar material are driven by
an extra restoring force, of electrical origin. This is the same mechanism which
drives longitudinal plasma oscillations, and in fact if we ideally set TO = 0 in
Eq. (D.8) we retrieve a squared plasma frequency, i.e. 4 times a squared charge
density over a mass density; the interaction is screened with . The relevant
charge is the Born effective charge, and the relevant mass is the reduced mass.
From the previous equations it is straightforward to obtain the identity
2
LO 0
2 = , (D.9)
TO

135
i.e. the famous Lyddane-Sachs-Teller (1941) relationship [233]. It is remark-
able that all microscopic parameters (force constants, masses, Born effective
charges, cell volume) disappear from Eq. (D.9), which is exact in the harmonic
regime. The identity relates quantities measured in quite different ways: inelas-
tic neutron scatteringleft hand side of Eq. (D.9)vs. infrared spectroscopy
right hand side. A final comment about first-principle calculations: therein,
the quantities directly provided by the code are TO , , and Z (and their
generalization, see below) [59, 60, 62], while LO and 0 are obtained from the
phenomenological approach. Therefore Eq. (D.9) is fulfilled by construction.

D.2 Infrared absorption


The simple case of a cubic binary material is pedagogically useful to address
dispersion and absorption in infrared spectroscopy. The -dependent dielectric
constant found in Sec. D.1 is the real (dispersive) part of the complex response
function ()+i (), while absorption is provided by the imaginary part. The
absorption coefficient per unit path length is related to () as [234]:

() = (), (D.10)
c n()
where n() is the index of refraction.
We rewrite Eq. (D.7) as
4e2 (Z )2
() = + 2 2) . (D.11)
Vcell M (TO
Causality requires that () and () are related by the Kramers-Kronig re-
lations, Eq. (C.14), which in our case read
2 ()
Z

() = d 2
0 2
Z
2 ()
() = d 2 . (D.12)
0 2
It immediately follows that
4 2 e2 (Z )2 2 2 e2 (Z )2
() = 2
(TO 2) = (TO ). (D.13)
Vcell M Vcell M TO
The meaning of this is pretty clear: we have an undamped oscillator at the
frequency TO , which only absorbs when the exciting field E is at resonance, and
furthermore at resonance the response diverges. The coefficient of (TO )
in Eq. (D.13) is the oscillator strength of the spectral line; this also explains
why the Born effective charges are also called infrared charges.
The simple example of a cubic binary crystal, having only a single dynamical
variable u, is pedagogically useful to address the response in time domain:
Z
D(t) = dt (t t ) E(t ). (D.14)

The response (t) can be obviously found using Eqs. (D.11) and (D.13), and
antitransforming () = () + i (). But we also remind that (t) can be

136
seen as the response to a -like input signal at t = 0, i.e. E(t) = E0 (t). Using
Eq. (D.2) we have:

4Z e
(t) E0 = D(t) = E0 (t) + u(t), t 0, (D.15)
Vcell
where we consider the electronic response as instantaneous. The Newton equa-
tion for u(t), from Eq. (D.2), is

2 Z e
u(t) = TO u(t) + E0 (t). (D.16)
M
Since we are looking for a causal response, u(t) = 0 for t < 0, and the response
to the instantaneous kick at t = 0 is the t > 0 solution of the homogenous
equation with initial conditions:
Z e
u(0) = 0; u(0) = E0 . (D.17)
M
We thus get
Z e
u(t) = E0 sin TO t. (D.18)
M TO
Finally, the response function is, from Eq. (D.15):

(Z )2 e2
(t) = 0, t < 0; (t) = (t) + sin TO t, t 0, (D.19)
Vcell M TO
where we retrieve the undamped oscillations of the responding system. More
about infrared spectroscopy will be said below, Sec. D.10.
While () is the response to an E field, its inverse 1 () is the response
to a D field. Even 1 () must obey Kramers-Kronig relationships. We are not
giving the details here, but the derivation is straightforward by exploiting the
identity
() 2 2
= 2LO , (D.20)
TO 2
which clearly displays the symmetry. The real part of () has a pole at TO ,
and the corresponding imaginary part a delta peak at the same frequency; the
real part of 1 () has a pole at LO , and the corresponding imaginary part a
() peak at the same frequency.

D.3 Dissipation and lifetime


So far, our response function was causal but nondissipative. The single-mode
prototypical example dealt with so far gives also the opportunity for introducing
phenomenological dissipation (or equivalently lifetime) into our response func-
tion. We start addressing the issue in the time domain, and we include damping
into Eq. (D.16):

2 Z e
u(t) = 2 u(t) TO u(t) + E0 (t). (D.21)
M

137
The damping has the dimensions of an inverse time, and we have included a
factor of 2 for convenience. We are interested in the case TO ; the solution
with the initial conditions of Eq. (D.17) is
Z e
q
u(t) = E0 et sin 1 t, 1 = TO 2 2. (D.22)
M 1
(Z )2 e2 t
(t) = 0, t < 0; (t) = (t) + e sin 1 t, t 0, (D.23)
Vcell M 1
The dielectric function in the -domain obtains by antitransforming this; is
clearly the inverse lifetime.
It is simpler instead to address the forced oscillations at frequency . A
dissipative term into Eq. (D.5) yields
M 2 u = 2iM uM TO
2
u + Z eE (D.24)
Z e
u= 2 E (D.25)
M (TO 2i 2 )
4e2 (Z )2
() = + 2 2i 2 ) . (D.26)
Vcell M (TO
We notice that () is now endowed with a real and an imaginary part, which
obey Kramers-Kronig relationships by construction. In the complex plane
() is an analytic function, with poles at 1 i. The previous expressions
can be compared with the zero-T quantum response, and the location of its
poles: see Eqs. (C.49) and (C.50) and the following discussion. In both cases in
fact that the poles are in the lower half of the complex plane is a manifestation
of causality. Note that the location of the poles on the upper/lower half plane
depends on our convention on Fourier transforms, Eq. (C.1); the opposite choice
may be found in the literature.
Th most perspicuous difference between Eq. (D.26) and its quantum (T = 0)
counterpart, Eqs. (C.49) and (C.50), is in the denominator: second order in
vs. first order. This owes to the fact that the Newton equations of motion are
second order in t, while Schrodinger equation is first order. If we are interested
in the region around resonance we may replace
2
TO 2 2i 2 TO (TO i), (D.27)
2 2
2e (Z ) 1
() + . (D.28)
Vcell M TO TO i
The non dissipative (although causal) response obtains by taking the 0+
limit. Exploiting the well known identity
1 1
lim = P + i(x) (D.29)
0+ x i x
we get () identical to Eq. (D.13). At finite positive the distribution (TO
) is broadened into a normalized Lorentzian, and the oscillator strength of the
spectral line is conserved. In general we have
2e2 (Z )2 TO
() + , (D.30)
Vcell M TO (TO )2 + 2
2e2 (Z )2
() . (D.31)
Vcell M TO (TO )2 + 2

138
While () yields the absorptive part of the response, () yields the dispersive
one. While crossing the resonance, D() switches from in-phase to out-of-phase
with E().

D.4 Dynamical matrix


Let us consider a solid at zero temperature in the adiabatic approximation: the
ionic positions at equilibrium are Rls = Rl + Rs , where l is a cell index and s is
(in nonprimitive lattices) a basis index in the unit cell of n atoms. A distorted
configuration of the solid is described by the set of the ionic displacements
{uls }; as throughout the present Notes, Greek subscript are used for Cartesian
components.
The total energy is expanded in the displacements around equilibrium as
(0) (2)
Etot ({uls }) = Etot + Etot ({uls }) + .... (D.32)

and the harmonic force constant are second derivatives, defined through

(2) 1 X
Etot ({uls }) = css , (l, l )uls, ul s , . (D.33)
2
ll ss ,

Because of lattice periodicity, the force constants depend only on Rl Rl and


hence css , (0, l ) contain all of the information. Besides this, there are further
important constraints imposed by translational and rotational invariance [235]:
we consider explicitly only the former. If the solid is translated as a whole, the
energy is unchanged, and hence the energy expansion vanishes to all orders; this
is easily shown to imply
X
css, (0, 0) = css , (0, l ), (D.34)
l s

with the usual meaning of the primed sum.


Owing to the adiabatic approximation [236, 237] the ionic motion is classical
in the potential energy provided by Etot . The harmonic oscillations around
equilibrium are then governed by the equation of motion:
X
Ms uls, = css , (l, l )ul s , , (D.35)
l s ,

for which normal-mode solutions (within periodic boundary conditions) are


phonons of q wavevector:

uls, = us, (q)eiq(Rl +Rs ) ei(q)t . (D.36)

Owing to the transformation, the secular problem factorizes for different qs


(chosen by convenience within the first Brillouin zone). Using then the reciprocal
form of the force constants:
X
css , (q) = eiq(Rs Rs ) css , (0, l )eiqRl , (D.37)
l

139
and introducing the auxiliary quantities:
1
us, (q) = es, (q); (D.38)
Ms
1
Dss , (q) = (Ms Ms ) 2 css , (q), (D.39)
where Ms are the ionic masses, the secular problem becomes:
X
2 (q)es, (q) = Dss , (q)es , (q). (D.40)
s ,

The dynamical matrix is Hermitian, and has (in stable systems) nonnegative
eigenvalues: at every q there are 3n normal modes, whose frequencies are the
square roots of the eigenvalues; the eigendisplacements are related to the eigen-
vectors by a trivial mass-dependent factor.
At this point I stress the importanceboth in model and first-principle
lattice dynamicsof a standard technical trick. It is convenient not to bother
explicitly with the actual value of the on-site force constant css, (0, 0). To this
effect one uses an auxiliary set of force constants, indicated with an overbar,
which in general violate translational invariance and are therefore nonphysical.
The physical force constants are recovered from
X
css , (q) = css , (q) ss css , (0), (D.41)
s

which ensures translational invariance in form. It is easy to verify that css, (0)
cancels in this expression, and hence the value of the on-site force constant
css, (0, 0) is irrelevant, whenever Eq. (D.41) is explicitly used.

D.5 Coulomb forces


We discuss in this section the contribution to lattice dynamics Coulomb interac-
tions. This is an actual ingredient within the framework of the rigid-ion model,
but is also used here for pedagogical purposes, in order to introduce nonanalytic
features of the first-principle dynamical matrix in presence of Coulomb forces.
Suppose that the central two-body potential between ions of species s and
s is ss (r); then its contribution to the lattice-dynamical force constants is
2 ss (|r|)


css , (l, l ) = . (D.42)
r r r=R +R Rl Rs
l s

Explicit evaluation of the second derivatives of a radial function yields


2 ss (|r|) 1 r r r r
= ss (r)(, 2 ) + ss (r) 2 , (D.43)
r r r r r
and we only need to calculate (r)/r and (r) for each shell of neighbors.
These quantities are often called tangential and radial force constants.
The above expressions only apply to l 6= l and s 6= s ; using an arbi-
trary value for the on-site force constant css, (0, 0), we get the analogous of
Eq. (D.37) as
X
css , (q) = eiq(Rs Rs ) css , (0, l )eiqRl , (D.44)
l

140
and the physical force constants are obtained from Eq. (D.41). The expression
in Eq. (D.44) is usable as such only for short-range interactions, where the sum
over l is rapidly convergent; the resulting force constants are then analytic
functions of q.
It is now convenient to transform the force constants into an equivalent
expression involving the Fourier transform ss (k) of the pairwise interaction.
Standard manipulations, starting from Eq. (D.44), give
1 X
css , (q) = (q + G) (q + G) ss (|q + G|)eiG(Rs Rs ) . (D.45)
Vcell
G

Specializing to a pure Coulomb interaction between point charges, its re-


ciprocal space expression is ss (k) = 4Qs Qs /k 2 . The force constants are
nonanalytic, as a fingerprint of long-range Coulomb interactions: we cast them
as

(Coul) 4Qs Qs q q
X (q + G) (q + G) iG(Rs Rs )
css , (q) = + e , (D.46)
Vcell q2 |q + G|2
G6=0

where the nonanalytic term has been expressly separated. This expression is
valid at q 6= 0: a major problem is the fact that css , (q = 0) does not cancel
(for s 6= s ) in Eq. (D.41). What is the correct value to be used? There are
several ways to reach a rigorous answer: the final result can however be sum-
marized into a working recipe, first proposed by Pines in the 1950s. Under the
hypothesis that the extended system is overall charge-neutral, one gets always
correct results upon assumption that the Fourier transform of Coulomb interac-
tion is 4/k 2 at k 6= 0 and vanishes at k = 0. Therefore the first term in square
brackets must be taken as zero when q is exactly zero. When q 0, i.e. in the
long-wavelength limit, this term gives in general a finite contribution. In fact
q q /q 2 acts as a projector on the q-direction. Owing to the presence of terms
of this kind, the dynamical matrix of a polar dielectric is non analytic at the
zone center.
In the simple case of a cubic binary crystal q q /q 2 discriminates between
longitudinal and transverse modes, and is therefore responsible for the LT split-
ting, already introduced within Huangs phenomenological theory.
Still, a minor problem remains: the reciprocal space sum over G in
Eq. (D.46) does not converge. The solution is straightforward: one evaluates
the Coulomb term giving a small Gaussian spread to the ionic charges, in order
to ensure convergence. When the spread is much smaller than the typical inte-
rionic distances, the force constants assume their physical (spread-independent)
value. Alternatively, more efficient Ewald techniques can be used [235]. There is
one simple case where the sum can be evaluated in closed form: let us consider
(Coul)
c12, (0) in a cubic binary crystal. Because of symmetry, this must be diagonal
in and equal to one third of the trace:
(Coul) 4Q1 Q2 X iG(R2 R1 )
c12, (0) = e . (D.47)
3Vcell
G6=0

Since we know that such a sum over all Gs (including G = 0) vanishes, we get
the result
(Coul) 4Q1 Q2
c12, (0) = . (D.48)
3Vcell

141
D.6 A simple approach to alkali metals
Rather accurate phonon spectra in simple metals have been obtained during the
1960s from pseudopotential perturbation theory to second order. In a sense,
this was the archetypical total energy approach to the electronic properties
of solids. Although all this is a rather obsolete approachsuperseded by first-
principle calculations [126]we include this Section for pedagogical purposes.
Within pseudopotential perturbation theory the total energy is the sum of
two terms: the former depends only upon the average density, while the second
is in fact a pairwise central interaction.
Phonon modes do not affect the average density, and therefore the full dy-
namical matrix is straightforwardly obtained from the formalism of the previous
section, where the two body interaction is the screened interaction between the
pseudoions. For a primitive lattice we omit the s subscripts; using Eqs. (D.41)
and (D.45) we get the dynamical matrix in the form:
1 X
D (q) = [(q + G) (q + G) (|q + G|) G G (G)]. (D.49)
M Vcell
G

The crudest approximation for the two-body interaction obtains from the
electron-gas dielectric function as

4Z 2 e2
(k) , (D.50)
(k)k 2

with Z = 1 for monovalent methods. This form for the dynamical matrix of a
simple metal was first proposed in 1958 by Toya [238], and clearly approximates
the pseudopotential with the bare Coulomb potential of the ion core, screened by
the electron gas. It neglects the short-range repulsion which basically mimicks
orthonormalisation of the valence orbitals to the core ones.
Over the years, many forms have been proposed for the electron-gas dielectric
function (k). All of them are divergent for k 0, thus making the two-body
interaction and the dynamical matrix analytic. The simplest of them is the
Thomas-Fermi dielectric function yielding the screened interaction

4e2
(k) 2 , (D.51)
k2 + kTF

where kTF is given in many textbooks [78]. If we single out the G = 0 term in
Eq. (D.49) we have

4e2 q q
D (q) = 2 ) (D.52)
M Vcell (q 2 + kTF
1 X
+ [(q + G) (q + G) (|q + G|) G G (G)].
M Vcell
G6=0

If one further neglects the second line, i.e. neglects the discrete nature of the
lattice, one gets the speed of sound as p /kT F , where p is the electron-gas
plasma frequency. We thus recover an important result found in 1950 upon
macroscopic arguments by Bohm and Staver and derived in several textbooks
[78]. The result is in semiquantitative agreement with the experiment; this

142
proved for the first time that sound propagation in metals is dominated by
electronic screening therein.
A much more accurate expression for the two-body interaction is provided
by pseudopotential perturbation theory [239]. Without giving a derivation, we
only show here the main result. One starts defining defining the dimensionless
function: 2
4Ze2
  
2 1
G(k) = v (k) 1 , (D.53)
k2 (k)
known as the energy-wave-number characteristic, where v(k) is the (local)
pseudopotential. This accounts for the interaction between the ion cores, me-
diated by the electrons. When we add the bare repulsion between the ionic
charges we get the two-body interaction as

4Z 2 e2
(k) = [1 G(k)] , (D.54)
k2
to be used in Eq. (D.49). As said above, this approach has provided in the
1960s remarkably accurate phonon spectra for the simple metals. The crude
approximation of Eq. (D.50) is retrieved replacing v(k) in Eq. (D.53) with the
Coulomb interaction.
A more accurate expression for the electron-gas dielectric function (k) has
a logarithmic singularity for k=2kF . The simplest (k) displaying this feature is
the Lindhard dielectric function, discussed in many textbooks [78]. The singu-
larity is due to the sharpness of the Fermi surface in metals: notice that even at
room temperature the Fermi surface can be considered as sharp to the present
purposes.
Therefore the dynamical matrix of Eq. (D.49) is singular at the q vectors (in
the first Brillouin zone) which fullfill |q + G| = 2kF for some G. The surfaces
defined by this relationship can be found from a simple geometrical construction:
when q crosses one of these surfaces, one expects a kink in the experimental
phonon dispersion curves, as first proposed by W. Kohn [240]. Shortly after the
theory, the effect was observed in the spectra of lead [241], and goes under the
name of Kohn anomaly since then.

D.7 Ionic crystals: rigid-ion model


The lattice dynamics of ionic cristals has important qualitative features which
are better illustratedfor pedagogical purposesstarting from the simple rigid-
ion model, which was first applied to the lattice dynamics of alkali halides in 1940
[242]. The force model adopted here is exactly the same as used in the Kittel
textbook [91] to discuss the structural stability of alkali halides: short-range
Born-Mayer forces (which only act between first neighbours), and point-charge
Coulomb forces. The parameters in the model are the ionic charges Q (where
usually |Q| = e, i.e. full ionicity), plus other two parameters in the Born-Mayer
forces. The latter are fitted to a pair of empirical data (typically the equilibrium
lattice constant and the bulk modulus) [91]. One would naively guess that the
Born-Mayer forces are entirely responsible for TO , and the Coulomb forces are
responsible for the longitudinal-transverse splitting: this is incorrect, as we will
see below.

143
Figure D.1: The phonon spectrum
of NaCl for q along the (1,1,1) di-
rection, as computed in 1940 in Ref.
[242].

The dynamical matrix is therefore the sum of two contributions, which are
separately evaluated. The Coulomb term is given by Eq. (D.46), using for the
ionic charges the values Qs = (1)s Q, where s=1 (2) labels the anion (cation);
the short range term is evaluated directly in the form of Eq. (D.44), where
the sum includes only nearest neighbors (six terms for the rocksalt structure).
(s.r.) (s.r.)
Therefore the short-range css , (q) vanishes when s=s , while c12, (q) de-
pends on two parameters only, given in Eq. (D.43). A typical result, taken from
the original literature, is shown in Fig. D.1.
It proves useful to study in detail the diagonalization at the zone center.
Keeping only terms of order zero in q, and using Eq. (D.41), we get:
(s.r.) (s.r.) (s.r.) (s.r.) (s.r.)
c11, (q) c22, (q) c12, (0) ; c12, (q) c12, (0); (D.55)

furthermore bulk Cartesian tensors are diagonal in a cubic material, hence


(s.r.)
c12, (0) = R0 , where the constant R0 completely accounts for the short
range forces at the zone center. Using the same path for the Coulomb term,
Eq. (D.46), we find a nonanalytic term (homogeneous of degree zero in q), plus
an analytic term which behaves in all respects like an effectively short-range
additional interaction. From Eqs. (D.46) and (D.48) we cast the complete dy-
namical matrix at the zone center as
4Q2 q q 4Q2
 
Dss , (q) + (R0 ) Mss , (D.56)
Vcell q 2 3Vcell
where we define the 22 matrix M as :
!
1
M1 M1 M
M= 1 2
. (D.57)
M1 M 1
M2
1 2

One of the eigenvalues of M is zero, and the other is 1/M = 1/M1 + 1/M2 ,
i.e. the inverse reduced mass; the corresponding eigenvectors are easily rec-
ognized to be the acoustic and optic modes, respectively. Going back to the
zone-center dynamical matrix, the acoustic mode is threefold degenerate with
frequency zero, while the optic modes require further diagonalization over the

144
Cartesian coordinates. There are two transverse modeswhere the polarization
is perpendicular to qwith

1 4Q2
T2 O = (R0 ), (D.58)
M 3Vcell
and one longitudinal modewhere the polarization is parallel to qwhose fre-
quency is
2 4Q2
LO = T2 O + . (D.59)
M Vcell
The LT splitting is substantially overestimated with respect to the experi-
ment, and this drawback is simply due to the fact that no mechanism within
the rigid-ion model allows for electronic screening of the Coulomb interaction.
In fact comparing Eq. (D.59) to the exact expression in Eq. (D.8), we easily
realise that the rigid-ion model sets Z = 1 and = 1. While the former is
a reasonable approximation to the real Z in ionic crystals, the latter is not:
2 3 in alkali halides.
Models correcting this problem are well known in the literature: they agree
better with the experimental spectra, at the price of using more empirical pa-
rameters. A comprehensive atlas of phonon spectra, as computed by means
of empirical models for many materials, was published in 1979 [243]. Nowa-
days such models have historical interest only, since phonon spectra are rou-
tinely computed by first principles, even for very complex crystal structures
[59, 60, 62].
One final lesson can be drawn from the rigid-ion model, looking at
Eq. (D.58), where it is perspicuous that TO is not due to the Born-Mayer
forces (summarized in R0 ) only: it also has a sizeable contribution from the
Coulomb forces. This was to be expected, since the Coulomb forces are essen-
2
tial to crystal stability as well. We notice that TO is the eigenvalue of the
analytical part of the dynamical matrix: Born-Mayer forces provide obviously
an analytic term, but there is another analytic term of Coulomb origin, coming
from the G 6= 0 terms in Eq. (D.46).
In a first-principle approach there is no way of disentangling the short-
range from the Coulomb forces: the interatomic forces are due to quantum
mechanics and electrostatics intertwined. We are only allowed to partition into
analytic and nonanalytic contributions. The key ingredient to the latter is
q q /q 2 , i.e. an homogeneous function of degree zero, whose value for q 0
depends on the q direction.

D.8 Zone-center modes: general case


Exploiting the results in Secs. D.1 and D.7, the exact zone-center dynamical
matrix of a cubic binary crystal is

4e2 (Z )2 q q
 
2
Dss , (q) = M TO + Mss , (D.60)
Vcell q 2

where the 22 matrix M is given in Eq. (D.57), and M is the reduced mass.
The nonanalytic expression in Eq. (D.60) is correct to order zero in q. In order

145
Figure D.2: A typical phonon spec-
trum of a low-symmetry dielectric
(crystalline BiFeO3 ). Owing to
Coulomb interactions, the dynami-
cal matrix is nonanalytic at the zone
center. The q 0 limit along dif-
ferent q directions may take differ-
ent values.

to proceed further, it is expedient to define the Born charges of each sublattice


Zs , with Z1 = Z2 , and rewrite Eq. (D.60) as

4e2 Zs Zs q q
 
1 (analytic)
Dss , (q) = Css , + , (D.61)
Ms Ms Vcell q2
(analytic)
whereowing to cubic symmetryCss , is diagonal in its Cartesian indices.
Several first-principle electronic structure codes [59, 60, 62] routinely com-
(analytic)
pute Css , , Zs , and , as well as their generalisations, discussed below.
We notice that Zs and require ad-hoc linear-response algorithms, while
(analytic)
Css , could even be computed in principle via frozen phonons, i.e. eval-
uating the ground-state energy of the crystal for non equilibrium geometries,
and expanding it to second order. We remind that standard ground-state cal-
culations implicitly assume a lattice-periodical Hamiltonian, hence E = 0. Any
lattice-periodical calculation is unable to provide the longitudinal phonon fre-
quency.
In a generic crystal with a basis of n atoms Eq. (D.61) generalises to
1 
(analytic) (nonanalytic)

Dss , = Css , + Css , (q) , (D.62)
Ms Ms
(nonanalytic)
where Css , (q) is a nonanalytic function of degree zero in q.
In a low-symmetry crystal the phonon spectrum is more complex: a typical
example is shown in Fig. D.2, illustrating a novel qualitative feature. We have
seen that for a cubic binary crystal the nonanalytic term in Eqs. (D.60) and
(D.61) is responsible for the LT splitting, but the q 0 limit is independent of
the q directions, even for the longitudinal mode. In a low-symmetry crystal the
zone-center modes are neither longitudinal nor transverse, and the q 0 limit
along different q directions may take different values.

D.8.1 Free energy and equations of motion


(nonanalytic)
In order to obtain the explicit expression for Css , (q) in a generic in-
sulating crystal we need to generalize our starting Eq. (D.1) to an arbitrary
lattice. We remind that a zone-center mode is lattice-periodical, hence inde-
pendent variables in the free energy per cell F are now the field E and the
sublattice displacements us , with s = 1, n. In a ferroelectric crystal the free
energy includes a term linear in E, and us -independent.

146
We limit the present treatment to crystals which are neither pyroelectric
nor ferroelectric. The most general second order expansion of F (E, {us }) reads
then
1 X (analytic) Vcell
F (E, {us }) = F0 + C us, us , E E
2 ss , 8
ss
X

e Zs, us, E ., (D.63)
s

(analytic)
where the sum over Cartesian (Greek) indices is left implicit. Therein Css ,
are the standard force constants, computed at zero E field, is the electronic

dielectric tensor, and Zs, is the dimensionless Born (a.k.a. infrared, a.k.a.
dynamical) effective charge tensor of sublattice s. It is defined as the mixed
second derivative

2F 2F
eZs, = , eZs, = . (D.64)
us, E E us,
(analytic)
Notice that, while Css , and are symmetric tensors in their Cartesian

indices, the Born charge Zs, in general is not a symmetric Cartesian tensor.
The Born effective charge tensors must fulfill the acoustic sum rule [244]:
X

Zs, = 0. (D.65)
s

If the sum of the tensors does not vanish identically, some acoustic phonons do
not have a vanishing frequency at q = 0.
The equations of motion are:
F X (analytic)

fs, = = Css , us , + eZs, E
us,
s
4 F 4e X
D = = Z us, + E . (D.66)
Vcell E Vcell s s,

We remind once more that all the coefficients occurring in Eqs. (D.63) and
(D.66) are routinely computed in some electronic-structure codes [59, 60, 62].

D.8.2 Microscopic meaning of the Born charges


We have said above that the s-th Born effective charge tensor measures the
macroscopic polarization linearly induced by a unit displacement of the s-th
sublattice in zero E field, or equivalently the force on the s-th atom induced by
a unit macroscopic E field at zero displacement. We adopt the latter viewpoint
in the following, and we notice that the force on the s-th atom is also equal to
the microscopic field E(micro)(r) at the nuclear site. Therefore

fs, = eZs, E = eZs E(micro) (Rs ), (D.67)

where eZs is the bare nuclear charge. This applies in an all-electron picture;
the force has a different expression in a pseudopotential framework.

147
The dimensionless s-th Born charge tensor measures therefore the ratio of
the microscopic field at site s to the macroscopic field. More precisely
(micro)
E (Rs )
Zs, = Zs . (D.68)
E
P
Since the acoustic sum rule [244] requires s Zs, = 0, the microscopic field
must bein simple cases where the tensors are diagonalparallel and antipar-
allel to the macroscopic field on different sites.
All experiments (either anelastic neutron scattering or infrared spectroscopy)
measure quantities which are quadratic in the Born tensors: therefore the sign
of Zs cannot be experimentally determined. Quantum-mechanical calculations
of the response functions, instead, do determine the sign. The first pioneering
calculations appeared in the early 1980s [245, 246]; nowadays computation of the
Born charge tensors is a standard feature of many electronic structure codes [59,
60, 61, 62]. Data are available for many complex materials, even noncrystalline.
Based on a simple rigid-ion viewpoint, one would expect the diagonal ele-
ments of Zs to be negative on anions and positive on cations, or equivalently
that the force on the nuclei is parallel to the macroscopic field on cations, and
antiparallel on anions. This is indeed what actual computations confirm, for
normal materials at least; some materials with counterintuitive Zs tensors are
discussed in the following. The counterintuitive features, where present, are
typically due to covalency and/or correlation.
The reasons why the Born charges are trivial or non trivial may be examined
in terms of Wannier functions: in fact the modern theory of polarization, in one
of its formulations, states that the polarization difference induced by a zone-
center phonon can be simply expressed in terms of the displacements of the
bare nuclear charges and of the Wannier-function centers (see Sec. 5.3.3). Let
us start with the simple case of alkali halides, where the computed Born charges
are close to their intuitive value of 1. In the paradigmatic example of NaCl
the nuclear charges are 11 and 17, respectively, and in the unperturbed solid
at equilibrium there are 5 and 9 doubly occupied maximally localized Wannier
functions centered at the two sites, respectively. When the two sublattices are
displaced, the Wannier centers are displaced as well. The modern theory of
polarization tells that Zs would be exactly 1 if the Wannier centers follow
rigidly the nuclear motion. What actually happens is that, in alkali halides, the
Wannier-center displacements are almost rigid. Notice that we are not assuming
that the static charge, belonging to each atom, is 1: the static charge is an
approximate concept, ill defined as a matter of principle [87, 88]. The Born
(a.k.a. dynamical) charge, instead, is well defined both from first principles and
experimentally. So much for alkali halides; at variance with them, in materials
where covalency plays an important role the Wannier-center displacements are
remarkably nonrigid.
In elemental crystals with a binary lattice, such as those having the diamond
structure, the Born tensors vanish and there are no infrared active modes. This
is due to symmetry and to the acoustic sum rule. However, elemental crystals
of lower symmetry in general do have infrared active modes. One known case
is solid hydrogen (a molecular crystal with several molecules in the elementary
cell).
Another interesting case is Se: its crystal structure, with three equivalent

148
Figure D.3: The crystal structure
of selenium. There are three equiv-
alent atoms per cell, arranged on
a spiral. The bonding is cova-
lent within the spiral (coordination
two), but there are weaker, almost
covalent, interspiral bonds.

atoms Pin the elementary cell, is shown in Fig. D.3. Owing to the acoustic sum

rule s Zs, = 0, the Zs, are purely off-diagonal, and they are equivalent
by rotations of 2/3. This means that a macroscopic field exerts on an atom
a force which is orthogonal to the field direction. The infrared activity can be
qualitatively understood as an effect of the bond charges belonging to the weak
interchain bonds
The next case that we wish to outline is the paradigmatic example of the
perovskite BaTiO3 in its undistorted cubic structure (nonferroelectric, a.k.a.
aristotype), where the oxygen ions sit at a noncubic sites. The Cartesian effec-

tive mass tensor ZO is diagonal, but strongly anysotropic. Based on a rigid-ion

picture, one would expect the elements of ZO to be about 2; instead, one

of them is of the order of 6. The tensor ZBa is isotropic, but is about +6,
much larger than the nominal Ba ionicity +4. The effect is due to the mixed
ionic-covalent character of this material, and of many perovskites as well. As a
consequence, the LO-TO splitting at the zone center is giant [248, 247].
The phonon spectrum of BaTiO3 in its cubic (aristotype) structure, com-
puted from first principles by Ghosez et al. [247], is shown in Fig. D.5. Besides
the giant LO-TO splitting, another prominent feature is worth a comment.
Some phonon frequencies are imaginary, i.e. the dynamical matrix has negative
eigenvalues. We have already observed that this material undergoes a sponta-
neous symmetry breaking, and therefore the cubic structure is not the relaxed
equilibrium geometry. In other words the occurrence of imaginary modes is a
manifestation of the fact that our harmonic expansion, Eqs. (D.33) and (D.63),
is in this case performed at a saddle point and not at a minimum.
Finally, we wish to discuss the hypothetical case of an highly correlated alkali
halide. It might happen that the electron-electron repulsion dominates over the

Figure D.4: The undistorted cubic per-


ovskite structure (aristotype) of BaTiO3 .
Solid, shaded, and empty circles represent
Ba, Ti, and O atoms, respectively. The O
atoms sit at orthorhombic, and not cubic,
sites. The oxygen Born tensors are diagonal
but anisotropic.

149
Figure D.5: The phonon
spectrum of BaTiO3 in
its cubic structure (aristo-
type). Notice the giant
LO-TO splitting, and also
the occurrence of imaginary
modes. After Ref. [247].

standard behaviour discussed above. In this highly correlated regime the anion
would have a positive Zs , and the cation a negative one. This feature has been
discovered via simulations on a two-band Hubbard model in one dimension [165].

D.8.3 Cochran-Cowley formula


We find expedient from now on to switch to compact formulas, adopting
tensorial-vectorial notations, leaving the Cartesian indices implicit, and without
a change of symbols; the dagger indicates the transpose. We thus equivalently
cast Eq. (D.66) as
X (analytic)
fs = Css us + e Zs E (D.69)
s
4e X
D = Z us + E. (D.70)
Vcell s s

We have learned from the paradigmatic example of the cubic binary crystal
that the field E must depend on the relative orientation between the wavevector
q and the sublattice displacements us (q). We make this explicit by rewriting
Eqs. (D.69) and (D.70) as
X (analytic)
fs (q) = Css us (q) + e Z s E(q) (D.71)
s
4e X
D(q) = Z us (q) + E(q), (D.72)
Vcell s s

and we remind that we remain at the lowest order: the dependence is nonan-
alytic and of order zero in q. Eq. (D.72) tells us that the polarization in zero
field is
e X
P0 (q) = Z us (q). (D.73)
Vcell s s
Next, it is expedient to define the unit vectors in the q-direction as

qx
q = qx qy qz .

q = qy , (D.74)
qz

Therefore the norm is q q = 1, while the dyadic product q q is the projector


in the direction of q, whose components are q q /q 2

150
We have already observed that in presence of a phonon of wavevector q, the
solid is macroscopically homogeneous in the plane normal to q, while all macro-
scopic properties display a modulation in the direction of q. It is immediate to
realize that the component of D(q) parallel to q and the component of E(q)
normal to q both vanish [89]:

q D(q) = 0, (1 q q )E(q) = 0. (D.75)

Whenever nonvanishing, both D(q) and E(q) are nonanalytic functions of order
zero in q. In the following equations, it is tacitly assumed that only the leading
term in q is considered. Whenever convenient, we may therefore replace D(q)
and E(q) with D(q) and E(q), respectively. Eqs. (D.72) and (D.73) are clearly
equivalent to
D(q) = E(q) + 4P0 (q). (D.76)
We now exploit Eqs. (D.75) and (D.76) as follows:

0 = q D(q) = q E(q) + 4q P0 (q) (D.77)

E(q) = q q E(q). (D.78)


From these it easily follows that

0 = q q q E(q) + 4q P0 (q) (D.79)


4
E(q) = q q P0 (q),
q q
4e X
=
q q Zs us (q) (D.80)
Vcellq q s

which can be interpreted as the depolarization field for an arbitrary q-direction.


We can finally eliminate the field in the equations of motion, Eq. (D.71), which
reads
X  (analytic) 4e2 (Z q)(q Z ) 
s s
fs (q) = Css + q
us (q). (D.81)

V cell q
s

The quantity in parenthesis is indeed the usual expression for the force-constant
matrix at the zone center, including the nonanalytic term, first obtained in
1962 by Cochran and Cowley [249] and implemented much later in some first-
principle codes [250, 126, 62, 59]. This confirms that the matrix elements at the
zone center are indeed nonanalytic functions, homogeneous of degree zero in q;
we also remind that Eq. (D.81) applies to crystals of any symmetry and is exact
at the harmonic level. The simple expression previously found in Eq. (D.61) is
clearly a special case of Eq. (D.81).
When we restore the Cartesian subscripts the nonanalytic contribution to
the dynamical matrix in Eq. (D.62) takes the more familiar (and prolix) form
P P
(nonanalytic) 4e2 Zs, q Zs , q
Css , (q) = P , (D.82)
Vcell q q

where even the sums on dummy indices are explicitated.

151
To inspect where nonanalyticity enters the dynamical matrix, we exploit
associativity in the matrix products, and we write

(nonanalytic) 4e2
Css (q) = Z (qq )Zs . (D.83)
Vcell (q q) s
The scalar q q is clearly analytic, hence the only source of nonanalyticity is
the projector in the q-direction qq , discussed above, and whose elements are
q q /q 2 .
Finally we observe that in a crystal of arbitrarily low symmetry all zone-
center phonons are coupled to the field. Therefore a zero-field calculation of the
force constants, as routinely provided by the popular computer codes [62, 59]
does not provide by itself the frequency (and the eigenvectors) of any zone-center
(analytic)
mode. It only provides Css , which must be used into the full Cochran-
Cowley formula in order to get all zone-center modes.

D.9 High-symmetry cases


We consider here only crystals whose symmetry is orthorombic or higher; then
all crystalline tensors can be simultaneously diagonalized. We obviously orient
the Cartesian axes along the orthorhombic crystal axes.
Since all Zs are diagonal Cartesian tensors, we rewrite Eq. (D.83) as

(nonanalytic) 4e2
Css , (q) = (Z Z )(qq ). (D.84)
Vcell (q q) s s
Even the products Zs Zs are diagonal, and the modes are purely longitudinal
or purely transverse.
For phonons polarised along the principal axes, the dynamical matrix can be
dealt with as a scalar (in general different in different directions), and one may
proceed analogously to our treatment of a cubic binary crystal. In that case the
6 6 dynamical matrix factorized in three 2 2 blocks, each with one acoustic
mode and only one optical mode; of the three blocks one was longitudinal and
two transverse. In the generalised case the matrix is 3n 3n, with n 1 optical
modes per direction.
Adopting from now on a scalar notation, we write the free energy per cell
using as independent variables the field E and the transverse (zero field) normal
mode coordinates un . The second order expansion of the free energy per cell is
n1 n1
1 X 2 2 Vcell 2
X
F (E, {uj }) = F0 + j u j E eE Zj uj . (D.85)
2 j=1 8 j=1

We have used for the expansion only 1/3 of the optic modes: those polarised
along the given Cartesian axis. We are thus generalizing the Huang single-mode
theory to the (n 1)-mode case. Notice however that here the normal-mode
coordinates uj include a factor with the dimensions of (mass)1/2 , while the mode
effective charges Zj include a factor with the dimensions of (mass)1/2 .
In Eq. (D.85) the mode frequencies j are obviously the transverse ones
(E = 0). The equation of motion for the normal modes are
fj = j2 uj + eZj E (D.86)

152
n1
4e X
D = Z uj + E. (D.87)
Vcell j=1 j

Proceeding as above the dielectric constant in the given direction is


n1
4e2 X (Zj )2
() = + (D.88)
Vcell j=1 j2 2
n1
4 2 e2 X 2
() = (Zj ) (j2 2 ), (D.89)
Vcell
j=1

and obviously obeys Kramers-Kronig relationships. While the poles of () are


the frequencies of the transverse zone-center modes, its zeros are the frequen-
cies of the longitudinal ones: it is not straightforward to find such zeros from
Eq. (D.88).
We write therefore the equations of motion for the longitudinal modes by
setting D = 0 in Eqs. (D.86) and (D.87):

uj = j uj + eZj E
n1
X
0 = e Zj uj + E, (D.90)
j=1

which yields
n1
4e2 X
2 uj = j2 uj + Zj Zj u j . (D.91)
Vcell
j =1

This equation shows that, in general, the longitudinal normal mode coordinates
are different from the transverse ones. This may happen even in cubic crystals,
like e.g. a cubic perovskite [248], whose structure is shown in Fig. D.4.
Suppose then we have diagonalized the longitudinal dynamical matrix,
Eq. (D.91), and found its eigenvalues j2 : since these are the zeros of (),
we may rewrite identically Eq. (D.88) as
Q 2 2
() j (j )
= Q 2 2
, (D.92)
j (j )

thus generalising the analogous single-mode formula, Eq. (D.20). This result
was found in 1961 by Kurosawa [251]. We thus immediately get the generalised
Lyddane-Sachs-Teller relationship in the form
Q 2
0 j j
= Q 2, (D.93)
j j

and we remind that this only holds if all the Born charge tensors are diagonal
on the Cartesian axes.
For a cubic crystal all the zone-center modes are threefold degenerate (at
E = 0): we may thus rewrite Eqs. (D.88) and (D.89) as
3n3
4e2 X (Zj )2
() = + (D.94)
3Vcell j=1 j2 2

153
3n3
4 2 e2 X 2
() = (Z ) (j2 2 ), (D.95)
3Vcell j=1 j

Where now the sum is over all the 3n 3 zone-center optic modes.

D.10 Finite temperature


In the present Appendix, we have implicitly assumed zero temperature so far.
In this framework, the imaginary part of the isotropic dielectric response ()
is given by Eq. (D.95) for a cubic crystalline system. We wish to compare with
the fluctuation formula, Eq. (C.32) valid for a general classical system at finite
T . We reproduce here Eq. (C.32) for the sake of clarity:

2
Z

() = 4 () = dt eit hd(t) d(0)i. (D.96)
3V

As discussed in Sec. C.5 this formula applies when the equilibrium fluctuations
are evaluated at E = 0, and d = V P, where V is the volume of the periodic
simulation cell. Ideally, the thermodynamic limit obtains for V .
We pause to comment on a very important feature: our expression in
Eq. (D.89) was evaluated at zero T , while Eq. (D.96) provides by definition
the finite-T response of the system. The key point is thatfor a purely har-
monic systemthe correlation function in Eq. (D.96) is exactly proportional
to 1/, and () is T -independent. In fact in a harmonic system at ther-
mal equilibrium the energy is in average half kinetic and half potential. From
the equipartition theorem and from Eq. (D.85) at E = 0 the average potential
energy per degree of freedom is
1 2 2 1
j huj i = , (D.97)
2 2

where we remind that u2j includes a factor with the dimension of a mass. The
1/ factor cancel the factor in Eq. (D.96): a virtue of harmonic systems only.
We are now going to do the bookkeeping, in order to show in detail that for
our harmonic system Eq. (D.95) and Eq. (D.96) are indeed identical. The free
evolution of each normal modewith zero fieldis uj (t) = uj (0)eij t , hence
the dipole of each normal mode is

dj (t) = eZj uj (0)eij t . (D.98)

We remind that here the normal-mode coordinates uj include a factor with the
dimensions of (mass)1/2 , while the mode effective charges Zj include a factor
with the dimensions of (mass)1/2 . Due to cubic symmetry, the j-th normal-
mode coordinate is parallel to one of the Cartesian axes.
Since correlation functions are quadratic, it is safer to separate the real from
the imaginary part:

uj (t) = uj (t) + iuj (t) = [uj (0) + iuj (0)](cos j t i sin j t)


uj (t) = uj (0) cos j t + uj (0) sin j t
uj (0)uj (t) = uj (0)uj (0) cos j t + uj (0)uj (0) sin j t. (D.99)

154
When the system is at equilibrium with a thermostat at the inverse temperature
the j 6= j fluctuations are uncorrelated. Eq. (D.97) yields
jj
huj (0)uj (t)i = jj h[uj (0)]i2 cos j t = cos j t (D.100)
j2

for any given Cartesian direction; orthogonal fluctuations are uncorrelated.


Summing over the modes polarised in the three directions, the dipole-dipole
time correlation function is
e2 X (Zj )2
hd(t) d(0)i = cos j t. (D.101)
j j2

Taking then the Fourier transform (a.k.a. power spectrum) we get

e2 X (Zj )2
Z
dt eit hd(t) d(0)i = [ (j ) + (j + ) ]
j j2
2e2 X 2
= (Zj ) (j2 2 ), > 0. (D.102)
j

Inserting this into the fluctuation-dissipation expression, Eq. (D.96), we


cancelas anticipatedthe apparent -dependence, and we arrive at
4 2 e2 X 2
() = (Zj ) (j2 2 ), (D.103)
3V j

which coincides with Eq. (D.95), the only difference being V vs. Vcell in the
denominator. Clearly, the equipartition theorem, Eq. (D.97), holds even for a
single cell with a microscopic number of degrees of freedom, provided these 3n3
degrees of freedom fluctuate at equilibrium with a thermostat at inverse temper-
ature . In anharmonic systems with a non separable Hamiltonian, instead, a
large supercell of volume V is mandatory for the use of the fluctuation formula,
Eq. (D.96). In any case, the dipole of a cell is extensive and the response is
intensive.
The equipartition theorem no longer holds in quantum statistical mechanics;
nonetheless in the harmonic case it is easy to correct Eq. (D.96) for quantum
effects. It is enough to replace

1 2 2 1 1 ~ 2
hu i = . (D.104)
2 j j 2 2 tgh ~
2

One obtains in this way the T -dependent quantum response of the harmonic
system as
4 ~
Z

() = 4 () = tgh dt eit hd(t) d(0)i. (D.105)
3~V 2

D.11 Anharmonic systems


As anticipated, the dipole-dipole autocorrelation functions are a standard tool to
compute the infrared spectra of disordered, anharmonic systems, from molecular

155
dynamics simulations, both classical and Car-Parrinello. The previous discus-
sion of the harmonic case was provided for pedagogical purposes only.
There is some disagreement between different authors about the relative
merits of Eq. (D.105) vs. Eq. (D.96). In fact Eq. (D.96) is the exact formula
at the purely classical level, while the quantum-corrected formula applies in
principle to the harmonic case only. The issue mostly concerns liquid water and
in general hydrogen-bonded systems, since H atoms are not classical particles at
room temperature. The relative merits of the different quantum corrections for
hydrogen-bonded anharmonic systems are investigated in Ref. [252]. We point
out, nonetheless, that only Eq. (D.96) obeys Kramers-Kronig relationships, as
it is clear from the previous derivation.
The absorption coefficient per unit path length is related to () as [234]:

() = () (D.106)
c n()

where n() is the index of refraction. We therefore obtain the two main formulas
on the market, from either Eq. (D.96) or Eq. (D.105), as:

2 2
Z
() = dt hd(0) d(t)i eit ; (D.107)
3cV n()

4 ~
Z
() = tgh dt hd(0) d(t)i eit . (D.108)
3~cV n() 2
Within Car-Parrinello simulations the dipole of the simulation cell is evalu-
ated as d = V P, where the electronic term in P is given by the modern theory
of polarization, by computing the single-point Berry phase (see Sec. 5.3.5) at
each time step orequivalentlythe Wannier centers. Not surprising, the ma-
terial whose infrared spectrum has been most studied is liquid water. The very
first Car-Parrinello infrared spectrum for liquid water appeared in 1997 [123].
Many other followed over the years; the literature is also flooded by simulations
based on a vast zoology of classical force models: all of them inadequate (in my
view) to capture the key features of hydrogen bonding.
Notice that the simulation only needs polarization differences in small time
steps. In fact at any discretized time nt the polarisation is

P(nt) = P(0) + [P(t) P(0)] + [P(2t) P(t)] + . . .


+ [P(nt) P((n 1)t). (D.109)

For small enough t the polarization quantum is harmless, as discussed in Sec.


5.3.5.

156
Bibliography

[1] R. Resta, Rev. Mod. Phys. 66, 899 (1994).


[2] R. Resta, in: Quantum-Mechanical Ab-initio Calculation of the Properties
of Crystalline Materials, Lecture Notes in Chemistry, Vol. 67, edited by
C. Pisani (Springer, Berlin, 1996), p. 273.
[3] R. Resta, J. Phys.: Condens. Matter 12, R107 (2000).
[4] R. Resta, J. Phys.: Condens. Matter 14, R625 (2002).
[5] D. Vanderbilt and R. Resta, in: Conceptual foundations of materials: A
standard model for ground- and excited-state properties, S.G. Louie and
M.L. Cohen, eds. (Elsevier, 2006), p. 139.

[6] R. Resta and D. Vanderbilt, in: Physics of Ferroelectrics: a Modern Per-


spective, Topics in Applied Physics Vol. 105, Ch. H. Ahn, K. M. Rabe,
and J.-M. Triscone, eds. (Springer-Verlag, 2007), p. 31.
[7] R. Resta, J Phys.: Conference Series 117, 012024 (2008).
[8] R. Resta, J. Phys.: Condens. Matter 22 123201 (2010).
[9] R. Resta, Eur. Phys. J. B 79, 121 (2011).
[10] M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010).
[11] J. E. Moore, Nature 464, 194 (2010).
[12] N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald, and N. P. Ong, Rev.
Mod. Phys. 82, 1539 (2010).
[13] D. Xiao, M.-C. Chang, and Q. Niu, Rev. Mod. Phys. 82, 1959 (2010).
[14] E. Prodan, J. Phys. A 44, 113001 (2011).
[15] X.-L. Qi and S.-C. Zhang, Rev. Mod Phys. 83, 1057 (2011).
[16] M. Z. Hasan and J. E. Moore, Annu. Rev. Condens. Matter Phys. 2, 55
(2011).

[17] J. Maciejko, T. L. Hughes, and S.-C. Zhang, Annu. Rev. Condens. Matter
Phys. 2, 31 (2011).
[18] T. Thonhauser, Int. J. Mod. Phys. B 25, 1429 (2011).

157
[19] N. Marzari, A. A. Mostofi, J. R. Yates, I. Souza, and D. Vanderbilt, Rev.
Mod. Phys. 84, 1419 (2012).
[20] N. A. Spaldin, J. Solid State Chem. 195, 2 (2012).
[21] R. Resta, Berrys Phase and Geometric Quantum Distance: Macroscopic
Polarization and Electron Localization, Lecture Notes for the Troisieme
Cycle de la Physique en Suisse Romande (Lausanne, 2000).
http://www-dft.ts.infn.it/~resta/publ/notes2000.ps.
[22] M. V. Berry, Nature Phys. 6, 148 (2010).
[23] Y. Aharonov and D. Bohm, Phys. Rev. 115, 485 (1959); reprinted in
Geometric Phases in Physics, edited by A. Shapere and F. Wilczek (World
Scientific, Singapore, 1989), p.104.
[24] R. P. Feynman, R. B. Leighton, and M. Sands, The Feynman Lectures in
Physics, Vol. 2 (Addison Wesley, Reading, 1964), Sect. 15-4.
[25] R. G. Chambers, Phys. Rev. Lett. 5, 3 (1960).
[26] M. Peshkin and A. Tonomura, The AharonovBohm Effect (Springer,
Berlin, 1989).
[27] http://en.wikipedia.org/wiki/SQUID.
[28] P. Bocchieri and A. Loinger, Nuovo Cimento 47, 475 (1978).
[29] http://www.phy.bris.ac.uk/people/berry mv/index.html.
[30] M. V. Berry, Proc. Roy. Soc. Lond. A 392, 45 (1984).
[31] H. C. Longuet-Higgins, U. Opik, M. H. L. Pryce, and R. A. Sack, Proc.
Roy. Soc. A 244, 1 (1958).
[32] G. Herzberg and H. C. Longuet-Higgins, Discuss. Faraday Soc. 35, 77
(1963).
[33] C. A. Mead and D. G. Truhlar, J. Chem. Phys. 70, 2284 (1979).
[34] C. A. Mead, Chemical Physics 49, 23 (1980).
[35] V. Heine, Phys. Rev.145,593 (1966).
[36] J. A. Appelbaum and D. R. Hamann, Phys. Rev. B 10, 4973 (1974).
[37] L. Kleinman, Phys. Rev. B 11, 858 (1975).
[38] F. Claro, Phys. Rev. B 17, 699 (1977).
[39] J. A. Appelbaum, G. A. Baraff, and D. R. Hamann, Phys. Rev. B 14,
1623 (1976).
[40] K. von Klitzing, G. Dorda, and M. Pepper, Phys. Rev. Lett. 45, 494
(1980).
[41] T. Ando, J. Phys. Soc. Jpn. 37, 622 (1974).

158
[42] R. B. Laughlin, Phys. Rev. B 23, 5632 (1981).
[43] R. E. Prange, S. M. Girvin, M. E. Cage, and K. von Klitzing, The Quan-
tum Hall Effect, Second Edition (Springer, New York, 1990).
[44] D. Yoshioka, The Quantum Hall Effect (Springer, Berlin, 2002).
[45] D. Bures, Trans. Am. Math. Soc. 135, 199 (1969).
[46] Geometric Phases in Physics, edited by A. Shapere and F. Wilczek (World
Scientific, Singapore, 1989).
[47] D. J. Thouless, Topological Quantum Numbers in Nonrelativistic Physics
(World Scientific, Singapore, 1998).
[48] A. Bohm, A. Mostafazadeh, H. Koizumi, Q. Niu, and J. Zwanzinger, The
Geometric Phase in Quantum Systems (Springer, Berlin, 2003).
[49] J. J. Sakurai, Modern Quantum Mechanics (Addison-Wesley, Reading,
1994), p.140.
[50] D. J. Thouless, M. Kohmoto, M. P. Nightingale, and M. den Nijs, Phys.
Rev. Lett. 49, 405 (1982).
[51] C. L. Kane and E. J. Mele, Phys. Rev. Lett. 95, 226801 (2005).

[52] D. N. Sheng, Z. Y. Weng, L. Sheng, and F. D. M. Haldane, Phys. Rev.


Lett. 97, 036808 (2006).
[53] S.-C. Zhang, Physics 1, 6 (2008).
[54] Y. L. Chen et al., Science 325, 178 (2009).
[55] J. E. Moore, Physics 2, 82 (2009).
[56] X.-L. Qi and S.-C. Zhang, Phys. Today 63(1), 38 (2010).
[57] J. P. Provost and G. Vallee, Commun. Math Phys. 76, 289 (1980).
[58] F. Wilczek and A. Zee, Phys. Rev. Lett. 52, 2111 (1984); reprinted
in:Geometric Phases in Physics, edited by A. Shapere and F. Wilczek
(World Scientific, Singapore, 1989), p. 141.
[59] http://www.abinit.org/.
[60] http://www.crystal.unito.it/.
[61] http://cms.mpi.univie.ac.at/vasp/.
[62] http://www.quantum-espresso.org.
[63] M.-C. Chang and Q. Niu, Phys. Rev. B 53, 7010 (1996).
[64] G. Sundaram and Q. Niu, Phys. Rev. B 59, 14915 (1999).
[65] C. A. Mead, Rev. Mod. Phys. 64, 51 (1992).
[66] M. Kohmoto, Ann. Phys. 160, 343 (1985).

159
[67] D. N. Zubarev, Non-Equilibrium Statistical Mechanics (Consultants Bu-
reau, New York, 1974).
[68] P. Schmelcher, L. S. Cederbaum, and H.D. Meyer, Phys. Rev. A 38, 6066
(1988).
[69] L. Yin and C. A. Mead, J. Chem. Phys. 100, 8125 (1994).
[70] D. Ceresoli, R. Marchetti and E. Tosatti, Phys. Rev. B 75, 161101(R)
(2007).
[71] R. Karplus and J. M. Luttinger, Phys. Rev. 95, 1154 (1954).
[72] T. Jungwirth, Q. Niu, and A. H MacDonald, Phys. Rev. Lett. 88, 207208
(2002).
[73] M. Onoda and N. Nagaosa, J. Phys. Soc. Jpn. 71, 19 (2002).
[74] Y. Yao, L. Kleinman, A. H. MacDonald, J. Sinova, T. Jungwirth, D.-S.
Wang, E. Wang, and Q. Niu, Phys. Rev. Lett. 92, 037204 (2004).
[75] X. Wang, J. R. Yates, I. Souza, and D. Vanderbilt, Phys. Rev. B 74,
195118 (2006).
[76] F. D. M. Haldane, Phys. Rev. Lett. 93, 206602 (2004).
[77] X. Wang, D. Vanderbilt, J. R. Yates, and I. Souza, Phys. Rev. B 76,
195109 (2007).
[78] N. W. Ashcroft and N. D. Mermin, Solid State Physics (Saunders,
Philadelphia, 1976).
[79] J. C. Slater, Phys. Rev. 76, 1592 (1959).
[80] J. M. Luttinger, Phys. Rev. 84, 814 (1951).
[81] J. Zak, Phys. Rev. 168, 686 (1968).
[82] G. Panati, H. Spohn, and S. Teufel, Commun. Math Phys. 242, 547
(2003).
[83] D. Xiao, J. Shi, and Q. Niu, Phys. Rev. Lett. 95, 137204 (2005).
[84] P. Streda, J. Phys. C 15, L717 (1982).
[85] D. J. Thouless, Phys. Rev. B 27, 6083 (1983).
[86] J. B. Pendry and C. H. Hodges, J. Phys. C 17, 1269 (1984).
[87] J. Meister and W. H. E. Schwarz, J. Phys. Chem. 98, 8245 (1994).
[88] Ph. Ghosez, J.-P. Michenaud, and X. Gonze, Phys. Rev. B 58, 6224
(1998).
[89] L. D. Landau and E. M. Lifshitz, Electrodynamics of Continuous Media
(Pergamon Press, Oxford, 1984).
[90] J. D. Jackson, Classical Electrodynamics (Wiley, New York, 1975).

160
[91] C. Kittel, Introduction to Solid State Physics, 7th. edition (Wiley, New
York, 1996).
[92] R. M. Martin, Phys. Rev. B 9, 1998 (1974).
[93] M. Posternak, A. Baldereschi, A. Catellani and R. Resta, Phys. Rev. Lett.
64, 1777 (1990).
[94] R. Resta, Ferroelectrics 136, 51 (1992).
[95] R. D. King-Smith and D. Vanderbilt, Phys. Rev. B 47, 1651 (1993).
[96] G. Ortz and R. M. Martin, Phys. Rev. B 43, 14202 (1994).
[97] R. Resta, Phys. Rev. Lett. 80, 1800 (1998).
[98] http://www.uam.es/departamentos/ciencias/fismateriac/siesta/.
[99] http://www.cpmd.org/.
[100] R. Resta, Modelling Simul. Mater. Sci. Eng. 11, R69 (2003).
[101] W. H. Duan and Z. R. Liu, Curr. Opin. Solid State Mater. Sci. 10, 40
(2006).
[102] M. Rabe, and J.-M. Triscone, eds., Physics of Ferroelectrics: a Modern
Perspective, Topics in Applied Physics Vol. 105, Ch. H. Ahn, K. (Springer-
Verlag, 2007).
[103] M.P. Marder, Condensed Matter Physics (Wiley, New York, 2000).
[104] G. Grosso and G. Pastori-Parravicini, Solid State Physics (Academic, San
Diego, 2000).
[105] I. Souza, J. Inguez, and D. Vanderbilt, Phys. Rev. Lett. 89, 117602
(2002).
[106] P. Umari and A. Pasquarello, Phys. Rev. Lett. 89, 157602 (2002).
[107] R. Resta, M. Posternak, and A. Baldereschi, Phys. Rev. Lett. 70, 1010
(1993).
[108] G. H. Wannier, Phys. Rev. 52, 191 (1937).
[109] N. Marzari and D. Vanderbilt, Phys. Rev. B 56, 12847 (1997).
[110] A. A. Mostofi, Y.-S. Lee, I. Souza, D. Vanderbilt, and N. Marzari, Comput.
Phys. Commun. 178, 685 (2008).
[111] J. R. Yates, C. J. Pickard, and F. Mauri, Phys. Rev. B 76, 024401 (2007).
[112] R. Resta, J. Chem. Phys. 124, 104104 (2006).
[113] E. I. Blount, in Solid State Physics, edited by H. Ehrenreich, F. Seitz and
D. Turnbull, vol 13 (Academic, New York, 1962), p. 305.

161
[114] O. F. Mossotti, Memorie di Matematica e di Fisica della Societa Ital-
iana delle Scienze Residente in Modena, 24, 49 (1850); R. Clausius, Die
Mechanische Behandlung der Electrica (Vieweg, Berlin, 1879).
[115] A. Pasquarello and R. Resta, Phys. Rev. B 68, 174302 (2003).
[116] I. Dabo, B. Kozinsky, N. E. Singh-Miller, and N. Marzari, Phys. Rev. B
77 , 115139 (2008); ibid. 84, 159910(E) (2011).
[117] P. Umari, A. Dal Corso, and R. Resta, in: Fundamental Physics of Fer-
roelectrics: 2001 Williamsburg Workshop, H. Krakauer, ed. (AIP, Wood-
bury, New York, 2001), p. 107.
[118] Q. Niu, Phys. Rev. 33, 5368 (1986).
[119] D. Vanderbilt and R. D. King-Smith, Phys. Rev. B 48, 4442 (1993).
[120] K. N. Kudin, R. Car, and R. Resta, J. Chem. Phys. 127, 194902 (2007).
[121] M. Stengel, Phys. Rev. B 84, 205432 (2011).
[122] N. C. Bristowe, P. B. Littlewood, and E. Artacho, J. Phys.: Condens.
Matter 23, 081001 (2011).
[123] P. L. Silvestrelli, M. Bernasconi, and M. Parrinello, Chem. Phys. Lett.
277, 478 (1997).
[124] A. Pasquarello and R. Car, Phys. Rev. Lett. 79, 1766 (1997).
[125] X. Gonze, Ph. Ghosez, and R. W. Godby, Phys. Rev. Lett. 74, 4035
(1995).
[126] S. Baroni, S. de Gironcoli, A. Dal Corso, and P. Giannozzi, Rev. Mod.
Phys. 73, 515 (2001).
[127] W. P. Su, J. R. Schrieffer, and A. J. Heeger, Phys. Rev. Lett. 42, 1698
(1979).
[128] F. Bloch, Z. Phys. 52, 555 (1928).
[129] A. H. Wilson, Proc. Roy. Soc. A 133, 458, and 134, 277 (1931).
[130] N. F. Mott, Proc. Phys. Soc. (London) 62, 416 (1949).
[131] P. W. Anderson, Phys. Rev. 109, 1492 (1958).
[132] N. Mott, Metal-Insulator Transitions, 2nd ed. (Taylor & Francis, London,
1990).
[133] E. Abrahams (Ed.), 50 Years of Anderson Localization, (World Scientific,
Singapore, 2010).
[134] W. Kohn, Phys. Rev. 133, A171 (1964).
[135] W. Kohn, in ManyBody Physics, edited by C. DeWitt and R. Balian
(Gordon and Breach, New York, 1968), p. 351.

162
[136] Only about 80 citations (ISI) in the 25 years from 1964 to 1991.
[137] R. Resta and S. Sorella, Phys. Rev. Lett. 82, 370 (1999).
[138] I. Souza, T. Wilkens, and R. M. Martin, Phys. Rev. B 62, 1666 (2000).
[139] R. Resta, Phys. Rev. Lett. 95, 196805 (2005).
[140] R. Kubo, M. Toda, and N. Hashitsume, Statistical Physics II, Nonequi-
librium Statistical Mechanics, Springer Series in Solid-State Sciences, Vol.
31, (Springer, Berlin, 1985).
[141] D. Forster, Hydrodynamic Fluctuations, Broken Symmetry, and Correla-
tion Functions (Benjamin, Reading, 1975).
[142] R. Resta, Phys. Rev. Lett. 96, 137601 (2006).
[143] C. Sgiarovello, M. Peressi, and R. Resta, Phys. Rev. 64, 115202 (2001).
[144] M. Veithen, X. Gonze, and Ph. Ghosez, Phys. Rev. B 66, 235113 (2002).
[145] N. D. M. Hine and W. M. C. Foulkes, J. Phys: Condens. Matter 19,
506212 (2007).
[146] A. Monari, G. L. Bendazzoli, and S. Evangelisti, J. Chem. Phys. 129,
134104 (2008).
[147] C. Aebischer, D. Baeriswyl, and R. M. Noack , Phys. Rev. Lett. 86, 468
(2001).
[148] V. Vetere, A. Monari, G.L. Bendazzoli, S. Evangelisti, and B. Paulus, J.
Chem. Phys. 128, 214701 (2008).
[149] L. Stella, C. Attaccalite, S. Sorella, and A. Rubio, Phys. Rev. B 84, 245117
(2011).
[150] M. El Khatib et al., J. Chem. Phys. 142, 094113 (2015).
[151] G. L. Bendazzoli, S. Evangelisti, A. Monari, and R. Resta, J. Chem. Phys.
133, 064703 (2010).
[152] E. Akkermans, J. Math. Phys. 38, 1781 (1997).
[153] W. Kohn, Phys. Rev. 115, 809 (1959).
[154] J. des Cloizeaux, Phys. Rev. 135, A685 (1964); ibid. 135, A697 (1964).
[155] S. Ismail-Beigi and T.A. Arias, Phys. Rev. Lett. 82, 2127 (1999).
[156] L. He and D. Vanderbilt, Phys. Rev. Lett. 86, 5341 (2001).
[157] G. F. Giuliani and G. Vignale, Quantum Theory of the Electron Liquid
(Cambridge University Press, Cambridge, 2005).
[158] W. Kohn, Phys. Rev. Lett. 76, 3168 (1996).
[159] S.F. Boys, Rev. Mod. Phys. 32, 296 (1960); J.M. Foster and S.F. Boys,
ibid. 300.

163
[160] P. B. Allen, in: Conceptual foundations of materials: A standard model
for ground- and excited-state properties, S.G. Louie and M.L. Cohen, eds.
(Elsevier, 2006), p. 139.
[161] D. J. Scalapino, S. R. White, and S. C. Zhang, Phys. Rev. Lett. 18, 2830
(1992).
[162] D. J. Scalapino, S. R. White, and S. C. Zhang, Phys. Rev. 47, 7995 (1993).
[163] J. G. Angyan, Int. J. Quantum Chem. 109, 2340 (2009).
[164] J. G. Angyan, Curr. Org. Chem. 15, 3609 (2011).
[165] R. Resta and S. Sorella, Phys. Rev. Lett. 87, 4738 (1995).
[166] T. Wilkens and R. M. Martin, Phys. Rev. B 63, 235108 (2001).

[167] S. Tamura and H. Yokoyama, JPS Conf. Proc. 3, 013003 (2014).


[168] V. K. Varma and S. Pilati, Phys. Rev. B 92, 134207 (2015).
[169] B. Kramer and A. MacKinnon, Rep. Prog. Phys. 56, 1469 (1993).
[170] D. J. Thouless, Phys. Rep. 13, 93 (1974).
[171] A. Lagendijk, B. van Tiggelen, and D. S. Wiersma, Phys. Today 62(8),
24 (2009).
[172] F. D. M. Haldane, Phys. Rev. Lett. 61, 2015 (1988).
[173] T. Thonhauser and D. Vanderbilt, Phys. Rev. B 74, 235111 (2006).
[174] D. J. Thouless, J. Phys. C 17, L325 (1984).

[175] C. Brouder, G. Panati, M. Calandra, Ch. Mourougane, and N. Marzari,


Phys. Rev. Lett. 98, 046402 (2007).
[176] T. Fukui, Y. Hatsugai, and H. Suzuki, J. Phys. Soc. Japan 74, 1674 (2005).
[177] D. Ceresoli and R. Resta, Phys. Rev. B 76, 012405 (2007).
[178] Q. Niu, D. J. Thouless, and Y. S. Wu, Phys. Rev. B 31, 3372 (1985).
[179] N. Hao et al., Phys. Rev. B 78, 075438 (2008).
[180] S. Coh and D. Vanderbilt, Phys. Rev. Lett. 102, 107603 (2009).
[181] R. Bianco and R. Resta, Phys. Rev. B 84, 241106(R) (2011).
[182] T. Thonhauser, D. Ceresoli, D. Vanderbilt, and R. Resta, Phys. Rev. Lett.
95, 137205 (2005).
[183] D. Ceresoli, T. Thonhauser, D. Vanderbilt, and R. Resta, Phys. Rev. B
74, 024408 (2006).
[184] R. Bianco and R. Resta, Phys. Rev. Lett. 110, 087202 (2013).
[185] C.-Z. Chang et al. Science 340, 167 (2013).

164
[186] C.-Z. Chang et al. Nature Materials 14, 473 (2015).
[187] A. A. Soluyanov and D. Vanderbilt, Phys. Rev, B 83, 035108 (2011).
[188] A. A. Soluyanov and D. Vanderbilt, Phys. Rev, B 85, 115415 (2012).
[189] A. J. P. Meyer and G.Asch, J. Appl. Phys. 32, S330 (1961).
[190] L. L. Hirst, Rev. Mod. Phys. 69, 607 (1997).
[191] D. Xiao, Y. Yao, Z. Fang, and Q. Niu, Phys. Rev. Lett. 97, 026603 (2006).
[192] J. Shi, G. Vignale, D. Xiao, and Q. Niu, Phys. Rev. Lett. 99, 197202
(2008).
[193] I. Souza and D. Vanderbilt, Phys. Rev. B 77, 054438 (2008).
[194] K.-T. Chen and P. A. Lee, Phys. Rev. B 84, 205137 (2011).
[195] H. Schulz-Baldes and S. Teufel, Commun. Math. Phys. 319, 649 (2013).
[196] T. Thonhauser, D. Ceresoli, A.A. Mostofi, N. Marzari, R. Resta, and D.
Vanderbilt, J. Chem. Phys. 131, 101101 (2009).
[197] D. Ceresoli, U. Gerstmann, A. P. Seitsonen, and F. Mauri, Phys. Rev. B
81, 060409(R) (2010).
[198] M. G. Lopez, D. Vanderbilt, T. Thonhauser, and I. Souza, Phys. Rev. B
85, 014435 (2012).
[199] A. M. Essin, J. E. Moore, and D. Vanderbilt, Phys. Rev. Lett. 102, 146805
(2009).
[200] A. Malashevich, I. Souza, S. Coh, and D. Vanderbilt, New J. Phys. 12,
053032 (2010).
[201] S. Coh, D. Vanderbilt, A. Malashevich, I. Souza and D. Vanderbilt, Phys.
Rev. B 83, 085108 (2011).
[202] J. Liu and D. Vanderbilt, Phys. Rev. B 92, 245138 (2015).
[203] R. P. Feynman, R. B. Leighton, and M. Sands, The Feynman Lectures in
Physics, Vol. 2 (Addison Wesley, Reading, 1964), Sect. 36-6.
[204] D. J. Griffiths, Introduction to Electrodynamics (Prentice-Hall, 1999).
[205] N. A. Hill, J. Phys. Chem. B 104, 6694 (2000).
[206] I. E. Dzyaloshinskii, Sov. Phys. JETP 10, 628 (1960).
[207] D. N. Astrov, Sov. Phys. JETP 11, 708 (1960).
[208] J. Iniguez, Phys. Rev. Lett. 101, 117201 (2008).
[209] J. C. Wojdel and J. Iniguez, Phys. Rev. Lett. 103, 267205 (2009).
[210] R. Resta, Phys. Rev. Lett. 106, 047202 (2011).
[211] R. Resta, Phys. Rev. B 84, 214428 (2011).

165
[212] A. Malashevich, S. Coh, I. Souza, and D. Vanderbilt, Phys. Rev. B 86,
094430 (2012).
[213] A. M. Essin, A. M. Turner, J. E. Moore, and D. Vanderbilt, Phys. Rev.
B 81, 205104 (2010).
[214] F. Bernardini, V. Fiorentini, and D. Vanderbilt, Phys. Rev. B 56, 10024
(1997).
[215] S. E. Park and T. R. Shrout, J. Appl. Phys. 82, 1804 (1997).
[216] R. M. Martin, Phys. Rev. B 5, 1607 (1972).
[217] R. M. Martin, Phys. Rev. B 6, 4874 (1972); W. F. Woo and W. Landauer,
Phys. Rev. B 6, 4876 (1972); R. Landauer, Solid St. Commun. 40, 971
(1981); C. Kallin and B. J. Halperin, Phys. Rev. B 29, 2175 (1984); R.
Landauer, Ferroelectrics 73, 41 (1987); A. K. Tagantsev, Phase Transi-
tions 35, 119 (1991).
[218] S. de Gironcoli, S. Baroni, and R. Resta, Phys. Rev. Lett. 62, 2853 (1989).
[219] D. Chandler, Introduction to Modern Statistical Mechanics (Oxford Uni-
versity Press, Oxford, 1987).
[220] B. Alder and T. E. Wainwright, Phys. Rev. A 1, 18 (1970).
[221] A. Widom, Phys. Rev. A 3, 1394 (1971).
[222] H. Kornfeld, Z. Phys. 22, 27 (1924).
[223] J. G. Kirkwood, J. Chem. Phys. 7, 911 (1939).
[224] H. Frohlich, Theory of Dielectrics, 2nd edition (Clarendon, Oxford, 1958).
[225] M. Neumann, Mol. Phys. 50, 841 (1983).
[226] S. W. de Leeuw, J. W. Perram, and E. R. Smith, Proc. Roy. Soc. London
Ser. A 373, 27 (1980); ibid. 373, 57 (1980).
[227] J. Ford, Physics Reports 213, 271 (1992).
[228] See Fermi-Pasta-Ulam on Wikipedia.
[229] D. N. Zubarev, Soviet Phys. Ushpekhi 3, 320 (1960).
[230] R. McWeeny, Methods of Molecular Quantum Mechanics, Second Edition
(Academic, London, 1992).
[231] K. Huang, Proc. Roy. Soc. A203, 178 (1950).
[232] M. Born and K. Huang, Dynamical Theory of Crystal Lattices (Oxford
University Press, Oxford, 1954).
[233] R. H Lyddane, R. G. Sachs, and E. Teller, Phys. Rev. 59, 673 (1941).
[234] D. A. McQuarrie, Statistical Mechanics (University Science Books, Sausal-
ito, California, 2000).

166
[235] A. A. Maradudin, E. W. Montroll, G. H. Weiss, and I. P. Ipatova, Theory
of Lattice Dynamics in the Harmonic Approximation, Solid State Physics,
Suppl. 3 (Academic, New York, 1971).
[236] G. V. Chester, Adv. Phys. 10, 357 (1961).
[237] E. G. Brovman and Yu. M. Kagan, in: Dynamical Properties of Solids,
edited by G. K. Horton and A. A. Maradudin, vol. 1 (North-Holland,
Amsterdam, 1974), p. 191.
[238] T. Toya, J. Res. Inst. Catalysis, Hokkaido Univ. 6, 161 and 183 (1958).
[239] W. A. Harrison, Pseudopotentials in the Theory of Metals (Benjamin, New
York, 1966).
[240] W. Kohn, Phys. Rev. Lett. 2, 393 (1959).
[241] Brockhouse et al, Phys. Rev. Lett. 7, 93 (1961).
[242] E. W. Kellermann, Phil. Trans. Roy. Soc. , A238, 513 (1940).
[243] H. Bilz and W. Kress, Phonon Dispersion Relations in Insulators
(Springer, Berlin, 1979).
[244] R. Pick, M. H. Cohen, and R. M. Martin, Phys. Rev. B 1, 910 (1970).

[245] R. Resta and A. Baldereschi, Phys. Rev. B 24, 4839 (1981).


[246] R. M. Martin and K. Kunc, Phys. Rev. B 24, 2081 (1981).
[247] Ph. Ghosez, E. Cockayne, U. V. Waghmare, and K. M. Rabe Phys. Rev.
B 60, 836 (1999).
[248] W. Zhong, R. D. King-Smith and D. Vanderbilt, Phys. Rev. Lett 72, 3618
(1994).
[249] W. Cochran and R. A. Cowley, J. Phys. Chem. Solids 23, 4471 (1962).
[250] P. Giannozzi, S. de Gironcoli, P. Pavone, and S. Baroni, Phys. Rev. B 43,
7231 (1991).
[251] T. Kurosawa, J. Phys. Soc. Jpn. 16, 1298 (1961).
[252] R. Ramirez, T. Lopez-Ciudad, P. Kumar P, and D. Marx, J.Chem. Phys.
121, 3973 (2004).

167

Anda mungkin juga menyukai