Resta Lecture Notes
Resta Lecture Notes
Resta Lecture Notes
Raffaele Resta
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 . . . . . . . . . . . . . . . . . . . . 2
1.2.2 Euler characteristic . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Electronic wavefunctions . . . . . . . . . . . . . . . . . . . . . . . 4
1.4 Units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.5 Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.6 Gauge and flux . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.6.1 Classical mechanics . . . . . . . . . . . . . . . . . . . . . . 6
1.6.2 Quantum mechanics, open boundary conditions . . . . . . 6
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 . . . . . . . . . . . . . . . . . . . . 34
4.2.1 Aharonov-Bohm effect . . . . . . . . . . . . . . . . . . . . 34
4.2.2 Molecular Aharonov-Bohm effect . . . . . . . . . . . . . . 36
4.2.3 Digression: the Z2 invariant . . . . . . . . . . . . . . . . . 37
4.2.4 Integer quantum Hall effect . . . . . . . . . . . . . . . . . 38
4.3 Adiabatic approximation in a magnetic field . . . . . . . . . . . . 39
4.4 Anomalous Hall effect . . . . . . . . . . . . . . . . . . . . . . . . 41
4.5 Semiclassical transport . . . . . . . . . . . . . . . . . . . . . . . . 42
4.5.1 Textbook equations of motion . . . . . . . . . . . . . . . . 42
4.5.2 Modern equations of motion . . . . . . . . . . . . . . . . . 42
4.5.3 Equations of motion in symplectic form . . . . . . . . . . 43
4.5.4 Geometrical correction to the density of states . . . . . . 44
4.5.5 Outstanding consequences of the modified density of states 45
4.6 Quantum transport . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.6.1 Transport by a single state . . . . . . . . . . . . . . . . . 46
4.6.2 Current carried by filled bands . . . . . . . . . . . . . . . 47
4.6.3 Quantization of charge transport . . . . . . . . . . . . . . 47
5 Macroscopic polarization 49
5.1 Polarization and electric field . . . . . . . . . . . . . . . . . . . . 49
5.2 Polarization “itself” vs. polarization difference . . . . . . . . . . 50
5.3 Independent electrons . . . . . . . . . . . . . . . . . . . . . . . . 52
5.3.1 The King-Smith and Vanderbilt formula . . . . . . . . . . 52
5.3.2 The quantum of polarization . . . . . . . . . . . . . . . . 53
5.3.3 Wannier functions . . . . . . . . . . . . . . . . . . . . . . 54
5.3.4 The surface charge theorem . . . . . . . . . . . . . . . . . 55
5.3.5 Noncrystalline systems: The single-point Berry phase . . 57
5.4 Correlated wavefunctions . . . . . . . . . . . . . . . . . . . . . . 58
5.4.1 Single-point Berry phase again . . . . . . . . . . . . . . . 58
5.4.2 Kohn-Sham polarization vs. real polarization . . . . . . . 59
ii
6.4.1 Independent electrons . . . . . . . . . . . . . . . . . . . . 69
6.4.2 Band insulators and band metals . . . . . . . . . . . . . . 71
6.5 Wannier and hermaphrodite orbitals . . . . . . . . . . . . . . . . 72
6.5.1 Wannier orbitals . . . . . . . . . . . . . . . . . . . . . . . 72
6.5.2 Hermaphrodite orbitals . . . . . . . . . . . . . . . . . . . 72
6.5.3 Maximally localized Wannier orbitals . . . . . . . . . . . 74
6.6 Localization in different kinds of insulators . . . . . . . . . . . . 75
6.6.1 Small molecules . . . . . . . . . . . . . . . . . . . . . . . . 75
6.6.2 Band insulators . . . . . . . . . . . . . . . . . . . . . . . . 75
6.6.3 Correlated (Mott) insulators . . . . . . . . . . . . . . . . 76
6.6.4 Disordered (Anderson) insulators . . . . . . . . . . . . . . 77
8 Orbital magnetization 90
8.1 Magnetization and magnetic field . . . . . . . . . . . . . . . . . . 91
8.2 Magnetization “itself” . . . . . . . . . . . . . . . . . . . . . . . . 92
8.2.1 The operator P r Q . . . . . . . . . . . . . . . . . . . . . . 93
8.2.2 Magnetization in k space . . . . . . . . . . . . . . . . . . 94
8.2.3 Magnetization in r space . . . . . . . . . . . . . . . . . . 95
Bibliography 99
iii
iv
Chapter 1
Introduction
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.
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
Ω = 1/R2 .
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 ∂x∂y
H= ∂2 z ∂2z
= . (1.2)
∂y∂x ∂y 2 0 1/R
In general, Ω 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
dσ Ω = 2(1 − g), (1.4)
2π 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
(“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 χ = 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 χ = 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: χ = 2(1 − g). Polyhedra can also have χ 6= 2, like the
doughnut-shaped one shown in Fig. 1.4.
4
Figure 1.4: A doughnut shaped polyhe-
dron. This surface has Euler character-
istic χ = 0 or, equivalently, genus g = 1.
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 ε0 and µ0 disappear; SI formulas are converted
by setting 4πε0 = 1 and 4π/µ0 = 1.
For a single particle, the Newton equation of motion and the Hamiltonian
read, respectively
dv 1
M =f =Q E+ v×B , (1.5)
dt c
2
1 Q
H= p − A(r) + QΦ(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.
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
5
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, P̂ , P, P, P. Despite this, I
found unavoidable to use—in different Chapters—the 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.
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 potentials—scalar
and vector—are auxiliary quantities, devoid of physical meaning.
6
where κ, having the dimensions of an inverse length, will be referred to as
“twist” in the following. The Schrödinger equation is
7
2
~2
2πn
ǫn (κ) = +κ , (1.17)
2m L
where the nontrivial κ-dependence is perspicuous. The velocity operator can be
written as
1 ∂H
v= , (1.18)
~ ∂κ
and the Hellmann-Feynman theorem yields
1 dǫn (κ)
hψn |v|ψn i = . (1.19)
~ dκ
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~κ/e is present along the rail; eigenvectors and eigenfunctions depend on
its value.
The flux breaks time-reversal symmetry (κ → −κ), and the spectrum is non-
degenerate, except when φ = 0 or φ = φ0 /2, the latter also called “π flux”. In
8
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
dǫn
I = −ev = −c . (1.21)
dφ
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
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 Φ”.
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].
11
Figure 2.3: A schematic representation of a (counterclockwise) pseudorota-
tion, where subsequent snapshots differ by 2π/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 ⊗ ε 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 ξ = (ξ1 , ξ2 ), the ionic displacements are parametrized as:
xA = ξ1 √
yA = ξ2 √
xB = − 21 ξ1 + 3
√2 2
ξ yB = − 3
√ 2 1
ξ − 21 ξ2
xC = − 12 ξ1 − 3
2 ξ2 yC = 3 1
2 ξ1 − 2 ξ2
12
The meaning of this coordinate choice is transparent with reference to Fig 2.3:
when the atom A is displaced by ξ, the displacements of B and C are of equal
magnitude |ξ|, but pointing in directions rotated by −2π/3 and −4π/3, re-
spectively. If we neglect Jahn-Teller coupling beyond linear order, no potential
energy is associated to a motion at constant |ξ|, 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
|ξ|, hence to linear order the electronic eigenvalues are:
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− (ξ) has a circular valley of radius ξmin , 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 + ) ω0 + Aj 2 , (2.3)
2
corresponding to an oscillation of frequency ω0 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 φ with half a flux
quantum φ0 (a.k.a. π 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 1963—much earlier than topology
became fashionable in electronic structure—by 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 φ = 0 and φ = φ0 /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”.
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 τ , 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 τ c
ne2 τ
jx = Ex − ωc τ jy
m
jy = ωc τ jx . (2.7)
ne2 τ
jx = σ0 Ex , σ0 = , (2.8)
m
while for B 6= 0 the conductivity tensor is:
σ0
jx = Ex = σxx Ex
1 + (ωc τ )2
ωc τ σ0
jy = Ex = σyx Ex . (2.9)
1 + (ωc τ )2
15
provides a remarkably simple expression for the longitudinal and transverse
resistivity
m mωc 1
ρxx = 1/σ0 = 2 , ρxy = = B. (2.11)
ne τ 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 (τ ≫ 1/ωc ) both ρxx and σxx vanish.
If we write n as N/A (number of carriers per unit area), then
1 φ
ρxy = B= , (2.12)
nec N ec
where φ = AB is the magnetic flux through area A. Although we are still at a
purely classical level, it is instructive to multiply and divide ρxy by h. We may
thus identically write
1 h N φ0
ρxy = , ν= . (2.13)
ν e2 φ
Here φ0 = hc/e is the flux quantum, as defined above. The dimensionless
quantity ν, called the filling factor, equals the ratio between the number of
electrons N and the number of flux quanta φ/φ0 . 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)
ν 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 ν are expected to be very different from the simple monothonical
form of Eq. (2.13).
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.
17
plateau, where Rxx = 0 and Rxy = 6453.3 ± 0.1, corresponding to the fill-
ing factor ν = 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 ρxx and ρxy 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 ν = 1 value is achieved above
≃ 10 tesla; at low field (high ν) the system becomes dissipative (ρxx > 0), while
the classical linear behavior of ρxy is recovered; the slope depends on electron
concentration n.
∂U (ϕ)
Ix = −c , (2.17)
∂ϕ
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 (ϕ) 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 stated 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 φ/φ0 . 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 σxx = 0 (the fluid is a “quantum
Hall insulator”), and ρxx = 0 (transport is dissipationless).
We now imagine to adiabatically increase the vector potential by an amount
∆A = φ0 /L, where φ0 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 (ϕ + φ0 ) −
U (ϕ) ≃ −φ0 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 −ν such integer number, the relationship is then
φ0 1 h
φ0 Ix /c = νeVy ; RH = Vy /Ix = = . (2.18)
νce ν e2
The flux ϕ acts therefore as a charge pump; the pump cycle is one flux quantum
φ0 .
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 ν cannot change, even if the number of current carry-
ing states does obviously change. The identification of ν 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
ψnk (x, y) = eikx χn (y − y0 ), y0 = ℓ2 k − . (2.19)
ωc B
For a finite L, the allowed k’s are integer multiples of 2π/L and the correspond-
ing centers y0 are spaced by 2πℓ2 /L = Lφ0 /φ. Threading a flux ϕ shifts y0
linearly in ϕ; when ϕ equals one flux quantum each eigenfunction goes over to
the next. Therefore one carrier is shifted for each n; the integer index ν 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
21
and its complement, i.e.
P̂ (ξ) = |Ψ0 (ξ)ihΨ0 (ξ)|; Q̂(ξ) = 1̂ − P̂ (ξ). (3.2)
Both P̂ (ξ) and Q̂(ξ) are gauge-invariant (for a fixed Hamiltonian).
22
Figure 3.2: A closed path joining four
states in ξ-space.
It is now clear that all the gauge-arbitrary phases cancel in pairs, such as to
make the overall phase γ a gauge–invariant quantity. The above simple–minded
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 Berry’s 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
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 ∆ξ:
−i∆ϕ ≃ hΨ0 (ξ)|∇ξ Ψ0 (ξ)i · ∆ξ. (3.9)
In the limiting case of a set of points which becomes dense on the continuous
path, the total phase difference γ converges to a circuit integral:
M
X I
γ= ∆ϕs,s+1 −→ A(ξ) · dξ, (3.10)
s=1 C
23
Since the state vectors are assumed to be normalized at any ξ, the connection
is real; we can therefore equivalently write
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 Σ.
For d 6= 3 the Berry curvature is conveniently written as the d × d antisym-
metric matrix
Ωαβ (ξ) = −2 Im h∂α Ψ0 (ξ)|∂β Ψ0 (ξ)i; (3.15)
Greek subscripts are Cartesian coordinates throughout, and ∂α = ∂/∂ξα . The
Stokes theorem can still be applied, generalizing Eq. (3.13) to
1
Z
γ= dξ α ∧ dξ β Ωαβ (ξ). (3.16)
2 Σ
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 ξ-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 γ, defined as the integral over a closed curve C of the
connection, is gauge invariant only modulo 2π. This indeterminacy is resolved
by Eqs. (3.13) and (3.16) whenever the curve C is the boundary ∂Σ of a surface
Σ where the curvature is regular. In fact, the curvature is gauge-invariant and
has no modulo 2π indeterminacy.
We notice that ∂S+ = ∂S− = C, but the surface normals n have opposite
orientations. From Stokes theorem, Eq. (3.13), we get:
Z I
Ω(ξ) · n dσ = ± A± (ξ) · dξ (3.19)
S± C
Z I I
Ω(ξ) · n dσ = A+ (ξ) · dξ − A− (ξ) · dξ . (3.20)
S2 C C
The two upper and lower Berry connections A± (ξ) may only differ by a gauge
transformation; the rhs of Eq. (3.20) is the difference of two Berry phases on
25
the same path and is necessarily a multiple of 2π. 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
Dξ ,ξ+dξ = gαβ (ξ)dξα dξβ , (3.21)
α,β=1
the projector Q̂(ξ) 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 Q̂(ξ) is irrelevant in the latter, i.e.
It is therefore clear that gαβ and Ωαβ 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 Fαβ in the following:
26
X′ hΨn (ξ)|∂α H(ξ)|Ψ0 (ξ)i
|∂α Ψ0 (ξ)i = |Ψn (ξ)i . (3.27)
E0 (ξ) − En (ξ)
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 ξ. 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 ξ-dependent phase factor is le-
gitimate, and must not modify any physical result, while the Berry connection
is instead affected. Nonetheless, since our Fαβ (ξ) 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
Fαβ (ξ) (3.28)
X′ hΨ0 (ξ)|∂α H(ξ)|Ψn (ξ)ihΨn (ξ)|∂β H(ξ)|Ψ0 (ξ)i
= .
[E0 (ξ) − En (ξ)]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.
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 π; 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 ξ 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.
Within loss of generality, we may address the case which is most interesting in
electronic structure. We assume that the n states |ψj (ξ)i are spin orbitals, and
that |Ψ(ξ)i is the many-body wavefunction built as their Slater determinant:
1
Ψ(ξ)i = √ |ψ1 (ξ)ψ2 (ξ) . . . ψN (ξ)|. (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
e−iγ = exp A(ξ) · dξ, (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 ′ (ξ) = −Im hψj (ξ)|∇ξ ψj ′ (ξ)i, (3.35)
and the Berry phase factor of Eq. (3.34) generalizes to the unitary matrix
I
e−iΓ = P exp A (ξ) · dξ. (3.36)
C
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 e−iγ is gauge-invariant, the Wilczek-Zee
unitary matrix e−iΓ is only gauge-covariant, i.e. a gauge transformation yields
a matrix unitarily equivalent to e−iΓ ; the gauge can be fixed by choosing the
vectors |ψj (ξ)i at one point of the path. The eigenvalues γj of the Hermitian
matrix Γ, defined modulo 2π, are gauge-invariant and in principle individually
observable. It is easily proved that their sum, i.e. the trace of Γ, coincides with
the Berry phase γ 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
Fαβ (ξ) = h∂α Ψ0 (ξ)|∂β Ψ0 (ξ)i − h∂α Ψ0 (ξ)|P̂ (ξ)|∂β Ψ0 (ξ)i. (3.37)
Next, we need now to explicitate the Cartesian indices α, β 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.
Fαβ,jj′ (ξ) = h∂α ψj (ξ)|∂β ψj ′ (ξ)i − h∂α ψj (ξ)|P (ξ)|∂β ψj ′ (ξ)i, (3.38)
where now P (ξ) is the single-particle projector; even this matrix is gauge-
covariant. If we rewrite the Cartesian components of Eq. (3.35) as Aα,jj ′ (ξ) =
−Im hψj (ξ)|∂α ψj ′ (ξ)i, the nonAbelian curvature becomes
Ωαβ,jj′ (ξ) = ∂α Aβ,jj ′ (ξ) − ∂β Aα,jj′ (ξ) − i[Aα (ξ), Aβ (ξ)]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 αβ
Cartesian indices, while at fixed αβ 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
Fαβ,jj (ξ)
j=1
Xn n
X
= h∂α ψj (ξ)|∂β ψj (ξ)i − h∂α ψj (ξ)|ψj ′ (ξ)ihψj ′ (ξ)|∂β ψj (ξ)i
j=1 jj ′ =1
= Fαβ (ξ). (3.40)
The proof of the last line of Eq. (3.40) is tedious, but straightforward.
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 |ψjk i = eik·r |ujk i, where |ujk i are
BZ-periodical, and are eigenfunctions of the Hamiltonian Hk = e−ik·r Heik·r .
While the |ψjk i at different k’s 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 k’s are arbitrary. Whenever possible, it is cus-
tomary to enforce the so-called periodic gauge |ψjk+G i = |ψjk i, which implies
|ujk+G i = e−iG·r |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
γ=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 γ = 0 and γ = π (modulo 2π).
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 |Ψ(k)i is the Slater determinant of the n occupied |ujk i, then
hΨ(k1 )|Ψ(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 = e−iG·r |ujk1 i. The discretized formula is then
Z
γ = i dk · hΨ(k)|∇k Ψ(k)i → −Im log ΠM s=1 hΨ(ks )|Ψ(ks+1 )i
C
= −Im log det ΠM
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
Fαβ (k) = h∂α ujk |∂β ujk i − h∂α ujk |uj ′ k ihuj ′ k |∂β 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
Fαβ,jj′ (k) = h∂α ujk |∂β uj ′ k i − h∂α ujk |uj”k ihuj”k |∂β uj ′ k i
j”=1
∞
X
= h∂α ujk |usk ihusk |∂β uj ′ k i. (3.48)
s=n+1
The many-band curvature obtains from either Eq. (3.47) or Eq. (3.49) as
n
X
Ωαβ = −2 Im h∂α ujk |∂β ujk i. (3.50)
j=1
31
Chapter 4
H(ξ) = ξ · ~σ
= ξ (sin ϑ cos ϕ σx + sin ϑ sin ϕ σy + cos ϑ σz ), (4.1)
where σα are the three Pauli matrices. The spectrum is non degenerate for
ξ 6= 0, and the lowest eigenvalue is −ξ. Upon symmetry arguments, we can
already guess the curvature to be isotropic.
sin ϑ2 e−iϕ
|ψ(ϑ, ϕ)i = .
− cos ϑ2
Aϑ = ihψ|∂ϑ ψi = 0
ϑ
Aϕ = ihψ|∂ϕ ψi = sin2
2
1
Ω = ∂ϑ Aϕ − ∂ϕ Aϑ = sin ϑ. (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
(ϑ = π); 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 singularity—often called
“obstruction”—can be moved but not removed. The algebra is the same as for
Dirac’s 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.
33
Figure 4.2: Discretization of the domain [0, π] ×
[0, 2π]; 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.
The Berry curvature is the Berry phase per unit (ϑ, ϕ) area. In this simple,
analytically soluble, case we know the exact value; Eq. (4.2) implies for Eq. (4.5)
γ = 12 sin ϑ ∆ϑ∆ϕ modulo 2π. 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 2π indeterminacy in Eq. (4.5)? The
size of ∆ϑ∆ϕ is very small, and each contribution γ to the sum is also small
(proportional to ∆ϑ∆ϕ), although Eq. (4.5) is in principle arbitrary modulo 2π.
It should be now pretty clear that the right solution is in choosing the “Im log”
branch with values in [−π, π].
Last but not least: where is the obstruction? In the continuous formulation,
any gauge choice yields a singularity at a single (ϑ, ϕ) point. In the discrete for-
mulation, there is no way to locate the singularity: in some sense, the singularity
is everywhere since the gauge is in principle erratic at all mesh points.
34
we will identify the ξ parameter with the box position R. Because of transla-
tional invariance, the R-dependence of the state vectors is
quantum
system
35
transformation, which changes A while leaving B invariant. In fact, in this
example B is essentially the Berry curvature. The Berry phase is
e 2π
I I
γ=− A(R) · dR = − A(R) · dR, (4.12)
~c C φ0 C
where φ0 is the flux quantum. Therefore γ 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.
We then multiply by the electronic eigenbra hΨ(ξ)| on the left, thus integrating
over the electronic degrees of freedom. We get the effective nuclear momentum
π acting on Φ as:
π Φ(ξ) = [ p − i~hΨ(ξ)|∇ξ Ψ(ξ)i ] Φ(ξ) = [ p − ~A(ξ) ] Φ(ξ), (4.15)
where we easily recognize the Berry connection. The momentum π is the kine-
matical (also called covariant, or mechanical) momentum, to be distinguished
from p = −i~∇ξ , 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 ξ in H(ξ, [x]) as a
classical parameter. The electronic eigenvalue E(ξ) 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 Φ(ξ) is then:
d
1 X
Heff = M−1
αβ πα πβ + E(ξ). (4.16)
2
α,β=1
36
In the molecular physics literature the extra term in Eq. (4.15) is seldom
mentioned, and π 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 ξ; the tradeoff is that—in some
cases—the 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 2π); the values al-
lowed by time-reversal symmetry are 0 and π; 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 Newton’s 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 present—and the Hamiltonian
is no longer time-reversal-invariant—is qualitatively different in this respect, see
Sec. 4.3 below.
37
i.e. γ = π mod 2π: 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.
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
σxy = 2 Im dk . (4.22)
~ (2π)2 BZ j=1 ′ (ǫjk − ǫj ′ k )2
j =n+1
38
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 ′
(ǫjk − ǫj ′ k )2
j =n+1
n
X 1
= Im h∂x ujk |∂x ujk i = − Ωxy (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 2π times an integer, the (first)
Chern number C. The milestone TKNN discovery is that Hall conductivity is
a Chern number when expressed in klitzing−1 :
e2
σxy = − 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
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 ν using the same argu-
ments as in Sec. 2.4.4; the integer can only be varied by crosssing a conducting
state.
39
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 π includes
now the two different vector potentials—yields 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 naı̈ve Born–Oppenheimer approximation—where Berry phases are
neglected—the 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 ξ with the nuclear coordinate R. If the atom is
subject only to a magnetic field, the complete Hamiltonian H and the electronic
Hamiltonian H are
1 h e i2
H(R, r) = p − A(R) + H(R, r); (4.25)
2M c
1 h e i2 e2
H(R, r) = −i~∇r + A(r) − . (4.26)
2m c |r − R|
As explained above, the nuclear kinematical momentum of Eq. (4.15) becomes
e
π = p − ~A(R) − A(R) (4.27)
c
The case of a constant B field can be dealt with analytically. We choose the
central gauge A(r) = 21 B × r. If φ(r) is the exact ground eigenfunction when
the proton sits at R = 0, the eigenfunction at a generic R is:
ie
hr|ψ(R)i = e− 2~c r·B×R φ(r − R), (4.28)
with an R-independent eigenvalue. The Berry connection is clearly
e e
A(R) = ihψ(R)|∇R ψ(R)i = − hψ(R)|B × r|ψ(R)i = − B×R
2~c 2~c
e
= − A(R), (4.29)
~c
since the R-derivative of φ(r − R) does not contribute. Replacing Eq. (4.29)
into Eq. (4.27) we find π = p, as it must be: the nucleus travels at constant
speed, and is not deflected by a Lorentz force.
40
Remarkably, the “magnetic-like” field due to the Berry phase is—in this sim-
ple example—exactly 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].
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
σxy = 2 Im dk h∂x ujk |∂x ujk i; (4.31)
~ (2π)d j=1 ǫjk <ǫF
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. (4.31) are the (periodic parts of) genuine
Bloch orbitals.
Given that the Fermi surface is symmetrical under k → −k, the symmetry
considerations of Sec. 3.7 show that Eq. (4.31) 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
41
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. (4.31); 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-
dau’s 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. (4.31), 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].
42
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) = − h∇k uk | × (Hk − ǫk ) |∇k uk i. (4.35)
2~c
Because of this, the band structure acquires a B-dependent term
1 ∂ǫ̃k
ṙ = − k̇ × Ω(k)
~ ∂k
1
~k̇ = −e (E + ṙ × B), (4.37)
c
where Ω(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.
where φ(r) is the potential of the electric field E(r), and m(k) is the wave
packet’s moment, Eq. (4.35). Then the EoM are
1 ∂H (r, k)
ṙ = − k̇ × Ω(k)
~ ∂k
1 ∂H (r, k) e
k̇ = − − ṙ × B(r), . (4.39)
~ ∂r ~c
In order to realize the relationship of Eq. (4.39) to the standard symplectic form
of Hamilton’s equations, it is expedient to adopt atomic units, and furthermore
43
to redefine B by including the 1/c factor: we get
ṙ = ∇p H (r, p) − ṗ × Ω(p)
ṗ = −∇r H (r, p) − ṙ × B(r). (4.40)
Notice that the translational momentum is coupled to B via the explicit Lorentz
force ṙ × 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. (4.40) conserves the energy:
dH
= ∇r H · ṙ + ∇p H · ṗ = 0 (4.41)
dt
In the simple case where Ω and B are identically vanishing, H becomes the
standard classical Hamiltonian and Eq. (4.40) are the corresponding Hamilton’s
equations. If I is the 3 × 3 identity, the standard 6 × 6 symplectic matrix is
0 −I
Θ= , (4.42)
I 0
When either B or Ω (or both) are nonzero, one recasts identically Eq. (4.40) in
the similar form
ṙ ∇r H
Θ̃ = , (4.44)
ṗ ∇p H
where Θ̃ is a deformed symplectic matrix defined as
~
B(r) −I
Θ̃ = ~ , (4.45)
I Ω(p)
~
and the 3 × 3 antisymmetric matrices B(r) ~
and Ω(p) are
0 −Bz By 0 −Ωz Ωy
B~ = Bz 0 −Bx , Ω ~ = Ωz 0 −Ωx . (4.46)
−By Bx 0 −Ωy Ωx 0
~ and the
Finally, we notice that if we set to zero both the Berry curvature Ω
moment m(k) in Eq. (4.38), then Eq. (4.44) provides an alternative form for
Hamilton’s 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.
44
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 · ṙ + ∇p · ṗ
∆V ∂t
= ∇r · [ ∇p H − ṗ × Ω(p) ] + ∇p · [ −∇r H − ṙ × B(r) ]
= −∇r · [ ṗ × Ω(p) ] − ∇p · [ ṙ × B(r) ]. (4.47)
In order to arrive at Eqs. (4.48) and (4.49) two special features of the matrix
Θ̃(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 Ω(p) are both zero.
Within ordinary statistical mechanics, the density of states in phase space is
1/hd (in dimension d). The number of states in volume ∆V = ∆r∆p is ∆V /h3 ,
and—owing to Liouville’s theorem—this 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 2π
−→ d [ 1 + B(r) · Ω(p/~) ], (4.50)
hd h φ0
where φ0 is the flux quantum.
45
density n depends on B. At zero temperature (for each spin channel)
1 2π
Z
n = dk 1 + B · Ω(k) ϑ(µ − ǫk )
(2π)d BZ φ0
∂n 1
Z
= dk Ω(k) ϑ(µ − ǫk ) (4.51)
∂B µ (2π)d−1 φ0 BZ
The latter assumes a perspicuous meaning in 2d, when µ is in a gap:
∂n 1 C1 1
Z
= dk Ω(k) = = − σxy . (4.52)
∂B µ 2πφ0 BZ φ0 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.
Since the adiabatic density matrix commutes with H(t), the time evolution
is
d
[H(t), ∆ρ(t)] = i~ρ̇(t) ≃ i~ |ψ0 (t)ihψ0 (t)|
dt
= i~ ( |ψ̇0 (t)ihψ0 (t)| − |ψ0 (t)ihψ̇0 (t)| ), (4.55)
where we are neglecting a term of higher order in the adiabaticity parameter.
We now take the matrix elements between hψ0 (t)| and |ψn (t)i:
(E0 − En )hψ0 (t)| ∆ρ(t) |ψn (t)i = i~ ( hψn (t)|ψ̇0 (t)i − hψ̇0 (t)|ψn (t)i ). (4.56)
The term with n = 0 in the rhs vanishes because of norm conservation; replace-
ment into Eq. (4.54) yields
v(t) = hψ0 (t)| v |ψ0 (t)i
" #
X hψ̇0 (t)|ψn (t)ihψn (t)| v |ψ0 (t)i
+i~ − c.c. . (4.57)
E0 − En
n6=0
The first term on the rhs is zero in the special case where—at all times—the
Hamiltonian H(t) is time-reversal invariant and the state is nondegenerate.
46
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 |ψ0k i = eik·r |u0k i is
" #
X hψ̇0k |ψjk ihψjk | v |ψ0k i
vk (t) = hψ0k | v |ψ0k i + i~ − c.c.
ǫ0k − ǫjk
j6=0
X hu̇0k |ujk ihujk | v |u0k i
= hu0k | v |u0k i + i~ − c.c. , (4.58)
ǫ0k − ǫjk
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 (4.59)
~ ~ ǫ0k − ǫjk
j6=0
1
vk (t) = ∇k ǫ0k + i ( hu̇0k |∇k u0k i − h∇k u0k |u̇0k i ). (4.60)
~
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 ( hu̇0k |∇k u0k i − h∇k u0k |u̇0k i ); (4.61)
(2π)d BZ
the Bloch states are normalized to one in the crystal cell (as everywhere in the
present Notes).
If the time evolution of the Hamiltonian is cyclic H(T ) = H(0), then the second
line in Eq. (4.62) 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 (4.63)
0
47
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
Ref. [87].
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).
48
Chapter 5
Macroscopic polarization
49
z
+ + + + + +
− − − − − −
x
boundary condition for the integration of Poisson’s 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 [104, 105]).
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 [89]. 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 theoretician’s 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 [88]: 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 = −4πP is the
depolarization field when P is normal to the slab (longitudinal polarization):
see Fig. 5.1.
50
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).
51
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).
52
It is essential that the gap does not close, i.e. the system remain insulating, for
all λ values.
The expression in Eq. (5.3) can be integrated with respect to λ to obtain
n Z
2ie X
P(λ) = − dk hujk |∇k ujk i : (5.4)
(2π)3 j=1 BZ
this is the (by now famous) King-Smith and Vanderbilt formula [94], 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. (4.62):
n Z
" Z #
2e X π/a
Z π/a π/a
Pz (λ) = − dkx dky i dkz hujk |∂kz ujk i , (5.5)
(2π)3 j=1 −π/a −π/a −π/a
53
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 [107], 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 [108]. A compre-
hensive review appeared as Ref. [19], and a public-domain implementation is in
wannier90 [109]. If the crystal is metallic, the WFs can still be technically
useful [110], 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 [111].
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 [112], 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 eik·R |ψjk i . (5.6)
(2π)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) (λ) = − 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 |ψjk 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 [113]: 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.
54
In the latter case, in fact, we may consider the set of WFs centered on a
given molecule; their total charge distribution coincides—in the weakly inter-
acting limit—with 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 [114]. 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 [115].
The case of alkali halides—where the model is often phenomenologically
used—deserves 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. [116].
55
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 been published in 2011 by Stengel [120].
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. [119].
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. [119]. 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.
56
7
[+-]
[NN]
6
3
Figure 5.5: Quantization of the
2
end charges in polyacetylene.
1 Dipole per monomer as a func-
0 tion of the number of monomers
-1
in the chain, for the two differ-
1 10 100
ent terminations.
(2π)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 π Z 2π
L L
γz = i dkz hujk |∂kz ujk i = i dkz hujk |∂kz ujk i → −Im log det S(k1 , k2 ),
π
−L 0
(5.10)
where k1 = (0, 0, 0) and k2 = (0, 0, 2π/L). In Eq. (5.10) we have used the
discretized Berry phase, Eq. (3.46), with only one factor in the matrix product.
2πz
Then, as in Sec. 3.9, we notice that |ujk2 i = e−i L |ujk1 i: therefore the overlap
matrix in Eq. (5.10) becomes
2πz
Sjj ′ (k1 , k2 ) = huj |e−i L |uj ′ i, (5.11)
where all the orbitals |uj i = |ψj 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) = − γz = Im log det S. (5.12)
πL2 πL2
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)
πL L m
57
This the expression currently used in computing power spectra and infrared
spectra [121], and, more generally, whenever a single k point is used in the
first-principle simulations [122, 114].
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. [122, 114],
or when performing Car-Parrinello simulations [121]. 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.
58
and the polarization formula obtains by replacing γz in Eq. (5.13), or equiva-
lently
2π
(det S)2 → Im log hΨ0 |e−i L i zi |Ψ0 i.
P
(5.16)
e 2π
Pz(el) = Im log hΨ0 |e−i L i zi |Ψ0 i.
P
(5.17)
2πL2
The proof of Eq. (5.17) is provided in the original paper [96], as well as in some
review papers [3, 6, 7, 8].
Here we content ourselves to prove that Eq. (5.17) coincides with Eq. (5.12)
in the special case where |Ψ0 i is the Slater determinant of the (doubly occu-
pied) k = 0 orbitals |uj i = |ψj i. The key observation is that the many-body
wavefunction
2π
|Ψ̃0 i = e−i L i zi |Ψ0 i
P
(5.18)
2πz
is the Slater determinant built from the orbitals |ũj i = e−i L |uj i, hence
2π
P
hΨ0 |e−i L i zi
|Ψ0 i = hΨ0 |Ψ̃0 i = ( dethuj |ũj ′ i )2 = (det S)2 , (5.19)
59
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 [124]—not reviewed here—or 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.
60
Chapter 6
61
tion space.
Kohn’s theory remained little visited for many years [133] 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 [134] in 1999
provided a definition of many-electron localization rather different from Kohn’s,
deeply rooted in the theory of polarization, and therefore based on geometrical
concepts. Their program was completed soon after by Souza, Wilkens and
Martin [135] (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.
As in Sec. 1.6.2 the flux is easily “gauged away”: the state e−iκ·r̂ |Ψ0 i coincides
with the ground eigenstate |Ψ0 (κ)i of the twisted Hamiltonian, Eq. (6.1). It
62
is legitimate to multiply this eigenvector by any κ-dependent (and position-
independent) phase factor; our choice is then
where d = hΨ0 |r̂|Ψ0 i is the electronic dipole of the molecular system. It follows
that the κ-derivative needed in our metric-curvature tensor, Eq. (3.24) is
63
or the Hamiltonian is κ-independent but the boundary conditions are “twisted”
in a κ-dependent way.
Within PBCs the tensor Fαβ (0) cannot be simplified—as e.g. in Eq. (6.5)—
and must be therefore addressed in its original form by actually evaluating the
κ derivatives:
Fαβ (κ) = h∂α Ψ0 (κ)|Q̂(κ)|∂β Ψ0 (κ)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
Fαβ (0)/N . We also observe that for time-reversal symmetric systems Fαβ (0) is
real symmetric, and coincides therefore with the metric tensor gαβ (0).
where the rhs is nothing else than ~ times the velocity operator v̂; Eq. (6.8)
becomes then
1 X′ hΨ0 |v̂α |Ψn ihΨn |v̂β |Ψ0 i
hrα rβ ic =
~2 N (E0 − En )2
n6=0
1 X′ hΨ0 |v̂α |Ψn ihΨn |v̂β |Ψ0 i
= 2 , (6.10)
N ω0n
n6=0
64
[135], where additional results are found, most notably relating localization to
conductivity. Since then, most authors (including the present one) have adopted
the notation hrα rβ ic for the localization tensor, where “c” stays for cumulant.
The formulation within OBCs, using the language and the notations familiar in
quantum chemistry, dates since 2006 [111].
The tensor hrα rβ 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
hrα rβ ic = Fαβ (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
65
whose response can be written, using the compact notations due to Zubarev
[67, 137, 138], as:
1
δhB̂i = ( hhB̂|Âiiω e−iωt + hhB̂|Âii−ω eiωt ). (6.16)
2
The quantity hhB̂|Âiiω is by definition the linear response induced by the per-
turbation  at frequency ω on the expectation value hB̂i. Straightforward
first-order perturbation theory provides its explicit expression as:
where ω0n are the excitation frequencies of the unperturbed system, and the
positive infinitesimal η ensures causality. Expressions of the kind of Eq. (6.17)
go under the name of Kubo formulae.
6.3.3 Conductivity
The conductivity tensor σαβ (ω) measures the current linearly induced by an
electric field: jα = σαβ Eβ . We therefore identify  with the potential of an
electric field along β, i.e. Â = eE rˆβ , and B̂ with the current operator −ev̂α /L3 .
An important detail must be stressed at this point. The macroscopic field
inside the sample includes by definition screening effects due to the electronic
system, while the perturbation δ Ĥ entering Eq. (6.1)—via the  operator—is
the “bare”, or unscreened one. This point will be discussed below (Sect. 6.3.5);
for the time being we simply identify screened and unscreened fields.
The Kubo formula for conductivity is therefore
e2
σαβ (ω) = − hhv̂α |r̂β iiω ; (6.18)
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. (6.17). As usual, we may
exploit the identity hΨ0 |r̂|Ψn i = ihΨ0 |v̂|Ψn i/ω0n . The Kubo formula becomes
then
ie2 X′ 1 hΨ0 |v̂α |Ψn ihΨn |v̂β |Ψ0 i
σαβ (ω) = lim
~L3 η→0+ ω0n ω − ω0n + iη
n6=0
hΨ0 |v̂β |Ψn ihΨn |v̂α |Ψ0 i
+ . (6.19)
ω + ω0n + iη
We introduce a compact notation for the real and imaginary parts of the
numerators in Eq. (6.19), i.e.
66
which are symmetric and antisymmetric, respectively. Using then
1 1
lim = P − iπδ(x), (6.22)
F →0+ x + iη x
and omitting the principal part, we separate for ω > 0 the symmetric and
antisymmetric parts in the conductivity tensor as
1 X ′ Rn,αβ
Re hrα rβ ic = 2
N ω0n
n6=0
1 X ′ In,αβ
Im hrα rβ ic = 2 . (6.24)
N ω0n
n6=0
67
entering Schrödinger 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. (6.25)
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 hrα rβ 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
[141]; hence Eqs. (6.25) and (6.26) are correct as they stand within PBCs. This
no longer holds within OBCs: in this case Eq. (6.25) needs to be modified, while
Im hrα rβ 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 hrα rβ 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. (6.25), assumes a different form:
this is discussed in Refs. [111, 141]. 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) [111,
141]. Within such theories, therefore, the sum rules hold in the simple form
of Eqs. (6.25) and (6.26); the conductivity therein is the independent-particle
conductivity (“uncoupled” response in quantum-chemistry jargon).
68
The f -sum rule yields
∞ ωp2 πe2 N
Z
dω Re σαα (ω) = = , (6.28)
0 8 2me L3
where ωp is the plasma frequency. Therefore the integral in Eq. (6.27) always
converges at ∞; its convergence/divergence is dominated by the small-ω behav-
ior of Re σαα (ω).
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 ω < Eg /~, and Eq. (6.28) yields
Z ∞
~L3 dω
Re hrα rα ic = 2
Re σαα (ω)
πe N Eg /~ ω
Z ∞
~2 L 3
< dω Re σαα (ω)
πe2 N Eg 0
~2
= . (6.29)
2me Eg
This inequality is due to SWM and clearly proves that Re hrα rβ 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 Kohn’s 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-ω
behavior of Re σαα (ω) is the result of a competition between numerators and
denominators in the Kubo formula, Eq. (6.19). Since we aim at a continuous
function of ω, the singularities in Eq. (6.23) must be smoothed: this can be done
by keeping the “dissipation” η finite while performing the thermodynamic limit
first [150]. For a band metal the localization tensor diverges (see below). Ac-
cording to SWM, a gapless material is insulating whenever Re σαα (ω) → 0 like
a positive power of ω, and metallic otherwise. The only example of gapless in-
sulator considered so far is a model Anderson insulator in 1d [149]. Simulations
prove indeed that hx2 ic is finite therein (Sect. 6.6.4).
69
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 |Ψ0 (κ)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
ρ(r, r′ ) = 2P (r, r′ ) = 2 ϕj (r)ϕ∗j (r′ ), (6.30)
j=1
where a singlet ground state is assumed, and ϕj (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) [111]:
1
Z
hrα rβ ic = drdr′ (r − r′ )α (r − r′ )β |P (r, r′ )|2 . (6.31)
N
If we define the complementary projector
where Nc is the number of electrons per crystal cell. According to the discussion
in Sec. 6.3.5, we need not to worry about shape issues in taking the limit; we
also stress that the density matrix, Eq. (6.30), 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. (6.34), while instead in
Eq. (6.33) 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.
70
6.4.2 Band insulators and band metals
As observed, Eq. (6.34) 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 [151, 152, 153, 154] 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 [142, 155] for 1d, 2d, and 3d. In all these
cases, the inner integral in Eq. (6.34) diverges like the linear dimension of the
system.
It is therefore clear that the localization tensor, when expressed in the form
of Eq. (6.34), measures in a perspicuous way the “nearsightedness”[156] 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. (6.30),
ϕj (r) → ψjk (r) = eik·r ujk (r), (6.35)
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 ψjk (r)ψjk∗
(r′ ) (6.36)
(2π)3 j=1 BZ
n Z
Vcell X ′
= dk eik·(r−r ) ujk (r)u∗jk (r′ ),
(2π)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. (6.36), 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
hrα rβ ic = dk Fαβ (k)
(2π)3 n BZ
X n n
X
Fαβ (k) = h∂α ujk |∂β ujk i − h∂α ujk |uj ′ k ihuj ′ k |∂β ujk i. (6.37)
j=1 jj ′ =1
The proof is given in Refs. [4, 142], and will not be repeated here.
Within OBCs the localization tensor is always real: this is perspicuous
in Eqs. (6.31) and (6.34). Instead the imaginary part of the PBC hrα rβ ic ,
Eq. (6.37), is the BZ integral of the Bloch Berry curvature, discussed below,
Sec. 7.1
71
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. (6.37) and the like. But these orbitals deserve a discussion of their own;
we limit ourselves to the crystalline case.
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 |ψjk i. (6.39)
(2π)3 BZ
72
The average quadratic spread in the x direction is by definition
N/2
2 X
λ2xx = (hϕi |x2 |ϕi i − hϕi |x|ϕi i2 ). (6.42)
N i=1
The gauge–invariant term in Eq. (6.44) coincides with the xx element hx2 ic of
the localization tensor hrα rβ ic , Eq. (6.33).
If we look for the orbitals which minimize the average spread in the x di-
rection, the solution, after Eq. (6.44), 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 χs ky kz (r) for a generic eigenstate of P xP , then it obeys the Bloch
theorem in the form
a glance at Eq. (6.39) 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.
73
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.
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.
where n is the number of occupied bands. As said above, the trace in Eq. (6.47)
is a lower bound (and not a minimum in dimension d > 1) for the spherical
second (cumulant) moment—a.k.a. quadratic spread—of the WFs, averaged
over the sample.
We stress that Eqs. (6.36) and (6.37) make sense only insofar the Fermi level
falls in a gap, in which case hrα rβ ic is always finite. If we vary the Hamiltonian
continuously, allowing the gap to close, then hrα rβ ic diverges; the quadratic
spread of the Wannier functions diverges as well [111].
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.
74
6.6 Localization in different kinds of insulators
6.6.1 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.
The tensor hrα rβ 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 [111].
The small-N version of the main concepts of the present review (in their
OBCs flavour [111]) has been adopted in quantum chemistry by Àngyàn
[158, 159]. Besides providing HF calculations of hrα rβ ic for a sample of small
molecules, Àngyàn even provides experimental values drawn from compilations
of the dipole oscillation-strength distributions: basically, from Eq. (6.25).
75
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. [142])
The points corresponding to Si
and Ge with the theoretical
gaps are out of scale. From Ref.
[142].
which clearly implies divergence of Eq. (6.50). At any finite N within OBCs
Eq. (6.50) leads to a finite hx2 ic value; however Eq. (6.51) 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 [149].
Other simulations [145] have addressed dimerized chains, i.e. ∆ = 0 but
alternant hoppings in Eq. (6.48). 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 [142]. 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 [143], and some model Hamiltonians in 1d and
2d [144].
76
Figure 6.3: Squared localiza-
tion length for the Hamiltonian
in Eq. (6.52) 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. [134].
77
Figure 6.4: Density of
states (arbitrary units) for
a model binary alloy in 1d.
The crystalline (band) case
corresponds to the Hamilto-
nian of Eq. (6.48) with ∆ =
0.25 and t = 1. The disor-
dered (Anderson) case cor-
responds to a random choice
of the anion/cation distri-
bution.
spectrum is gapless [130, 162]. The density of states for both the ordered and
disordered systems are shown in Fig. 6.4, and confirm the expected features.
The band structure of Eq. (6.49) yields obviously a gapped density of states; at
the band
√ edges it shows van Hove singularities, which in 1d have the character
of 1/ ǫ 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 [128]. More than fifty years of literature
have been devoted to investigate Anderson insulators under the most diverse
aspects. [130, 162, 163, 164].
At variance with such wisdom, Ref. [149] addressed this paradigmatic An-
derson insulator from the nonconventional viewpoint of the modern theory of
the insulating state. In the spirit of Kohn’s 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
computed within OBCs from Eq. (6.50), 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 but“locked” by the Pauli principle if the Fermi level
lies in the gap.
78
Chapter 7
Vcell 2Vcell
Z Z
Im hrα rβ ic = dk Im F αβ (k) = − dk Ω(k)
(2π)3 n BZ (2π)3 n BZ
n
Vcell
Z X
= dk Im h∂α ujk |∂β ujk i. (7.1)
(2π)3 n BZ j=1
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 [165], 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 [166, 167]; this is at variance with nor-
mal insulators where localized Wannier functions exist, and can even be chosen
as exponentially localized [19, 108, 168].
79
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.
γ = −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
γ = −Im log det S(k1 , k2 )S(k2 , k3 )S(k3 , k4 )S(k4 , k1 ), (7.4)
where S is the overlap matrix
see Eq. (3.46). One has to choose the “Im log” branch with γ in [−π, π].
The Chern number C1 is the sum of all the γs, 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
γs are summed.
80
We start rewriting Eq. (7.2) for a square lattice as
Z π/L Z π/L
1
C1 = dkx dky Ω(k), (7.6)
2π −π/L −π/L
81
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.
The uk orbitals are clearly lattice-periodical and normalized over the unit cell;
their k derivative yields
X
∂1 uk (r) = −i e−ikx (x−ma) (x − ma) χ̃ky (x − ma, y),
m
X ′ ∂
∂2 uk (r) = e−ikx (x−m a) χ̃k (x − m′ a, y),
∂ky y
m′
X ∂
h∂1 uk |∂2 uk i = i e−ikx ma hχ̃m ky |(x − ma)| χ̃0 ky i, (7.10)
m
∂ky
where “h i′′ means r integral in (−∞, ∞) × (0, a). The Berry curvature and the
Chern number are therefore
82
2π/a
1 dx0 1
Z
C1 = − dky = − [ x0 (2π/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 + 2π/a.
More generally, not only the center but even the function χ̃m ky (x, y) enjoys an
analogous property.
As for the k space periodicity of the Bloch functions, Eqs. (7.9) and (7.13)
yield
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 .
and the cumulant second moment is clearly finite. The trace hx2 ic + hy 2 ic of
the localization tensor is in fact equal to ℓ2 , the squared magnetic length, and
the minimum quadratic spread in the x direction is ℓ2 /2.
The LLL orbitals in the Landau gauge are labeled by a one-dimensional wave
vector q ∈ (−∞, ∞) and have the form
The quadratic spread of this function is indeed ℓ2 /2; the ψq (r) are therefore
eigenstates of P xP , and are the magnetic analogue of the maximally localized
hermaphrodite orbitals. Notice that in Eq. (7.16) ψq (r) has a plane-wave-like
normalization over y.
83
To make contact
√ with our notations for hermaphrodite orbitals it is enough
to define a = 2π ℓ and
84
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 h∂1 Ψ0 (0)|∂2 Ψ0 (0)i. (7.21)
N
In the large-L limit, we may write
Im h∂1 Ψ0 (0)|∂2 Ψ0 (0)i (7.22)
2 Z 2π Z 2π
L L L
= Im dκ1 dκ2 h∂1 Ψ0 (κ)|∂2 Ψ0 (κ)i.
2π 0 0
The dimensionless integral is over a closed surface (a torus), given the magnetic
boundary conditions. It is therefore equal to −πC1 , where C1 is the Chern
number of the first class (Sect. 3.4); hence
L2
Im hxyic = − C1 , (7.23)
4πN
and the transverse conductivity, Eq. (7.20) becomes
e2 e2
Re σ12 (0) = − C1 = − C1 . (7.24)
2π~ h
We have thus arrived at the famous Niu-Thouless-Wu formula [170] 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. [136] (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 ℓ = (~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.
[136] for details): the system is not insulating, and its transverse conductivity
is not quantized.
85
Figure 7.3: Four unit cells of crystalline Halda-
nium [165]. 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 eiφ . Arrows
indicate sign of the phase φ for second-neighbor hop-
ping.
7.3.1 Haldanium
The archetype of all TIs is the Haldane model Hamiltonian [165] (“Haldanium”):
its hallmark is quantized Hall conductance in absence of a macroscopic magnetic
field. Haldanium is the paradigm for Chern insulators, and has been used as
a workhorse in many simulations, providing invaluable insight into topological
features of the electronic wavefunction [165, 169, 166, 171, 172, 173], as well as
into features of orbital magnetization [174, 175, 169, 176].
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 e±iφ , 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 φ, this system
undergoes a transition from a normal insulator (C1 = 0) to Chern insulator
86
(|C1 | = 1). Its phase diagram is shown in Fig. 7.4. On the vertical φ = 0 axis
the Hamiltonian is time-reversal invariant and Haldanium becomes a model for
hexagonal boron nitride; the center of the plot (∆ = 0, φ = 0) corresponds
to graphene. Notice also that on the horizontal ∆ = 0 axis Haldanium is a
nonpolar and centrosymmetric insulator (except at φ = 0).
Like in the QH case discussed in Sect. 7.2: (i) the imaginary part of the
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. [166], where the actual value of ΩI , Eq. (6.47), is computed. We remind
that, despite ΩI 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 [166] vs. Gaussian in the (noninteracting)
QH case, Eq. (7.15).
No microscopic realization of a Chern insulator (in absence of a macroscopic
magnetic field) is known to date, although some unpublished work in this direc-
tion was announced at the APS 2012 March meeting. Chern insulators remained
a curiosity of academic interest only for many years. In 2005 it was realized that
the Haldane model is the forebear of a completely new class of topological in-
sulators, the time-reversal symmetric ones [51, 10, 11].
87
in the Z2 -odd case.
Acell
Z
Im hrα rβ ic = dk Im h∂α ujk |∂β ujk i, (7.25)
(2π)2 BZ
88
whence the Chern number is related to the imaginary part of the localization
tensor as
4π 4π
C1 = − Im hrα rβ ic = − Im Trcell{P xQy}, (7.26)
Acell Acell
where the second equality owes to Eq. (6.33).
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
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. [173] 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
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 [89]).
We stress that the microscopic function C(r) itself, before macroscopic averaging,
bears no physical meaning.
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 [173].
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 (2013).
89
Chapter 8
Orbital magnetization
where ρ(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 ρ(micro) (r) and j(micro) (r)
[180]. 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 [175, 181], and is
still (2013) partly work on progress [181, 182, 183, 169, 176]. A couple of reviews
have appeared [8, 18], while only a few first-principle implementations appeared
so far (2013) [184, 185].
90
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 electric field B(micro) (r), which fluctuates at the atomic scale [89].
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 theoretician’s 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 [88]: 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 = 4πM 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).
91
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 P—as stressed in Sec.
5.2—the 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 [94, 118] (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.
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 [175, 181, 182], 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 µ is in the
bulk gap, the value of M is µ-independent in the normal case and µ-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 µ.
The macroscopic magnetization at fixed chemical potential is M =
−(1/A) ∂G/∂B, where A is the system area and G = U − µN is the grand
thermodynamic potential at T = 0:
1 ∂U µ ∂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 = µ ∂n/∂B; in the thermodynamic limit we make contact with Streda’s
formula [84], Eq. (4.52):
∂n eC1 C1
= = , (8.3)
∂B 2π~c φ0
where φ0 = 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= εj . (8.4)
ǫj <µ
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
B → 0. We may therefore invoke the Hellmann-Feynman theorem to get
∂U X ∂H e X
− =− hϕj | |ϕj i = − hϕj | r × v |ϕj i; (8.5)
∂B ǫ <µ
∂B 2c ǫ <µ
j j
92
not surprisingly, M1 coincides then with the 2d analogue of Eq. (8.1). Using
v = i[H, r]/~ we have
ie X
M1 = − hϕn | r × [H, r] |ϕn i (8.6)
2~cA ǫ <µ
j
ie X
= − ( hϕn | xHy |ϕn i − hϕn | yHx |ϕn i ).
2~cA ǫ <µ
j
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)
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)
93
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 rQ—like P itself—can 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 |ψnk ihψnk |,
(2π)2 n=1 BZ
∞
Acell X
Z
Q = dk′ |ψn′ k′ ihψn′ k′ |. (8.12)
(2π)2 ′
n =Nc +1
94
expressed in terms of Bloch orbitals, as shown in Sec. 8.2.1:
Nc ∞
Acell X X Z
P xQHQyP = dk ×
(2π)2 n=1 BZ
n′ =Nc +1
× |ψnk ih∂1 unk |un′ k i εn′ k hun′ k |∂2 unk ihψnk |
Nc ∞
Acell X X Z
QxP HP yQ = dk ×
(2π)2 n=1 ′ BZ
n =Nc +1
× |ψn′ k ihun′ k |∂1 unk i εnk h∂2 unk |un′ k ihψn′ k | (8.16)
Nc
X ∞
X Z
Im Trcell {P xQHQyP − QxP HP yQ} = dk ×
n=1 n′ =Nc +1 BZ
× [ εn′ k h∂1 unk |un′ k ihun′ k |∂2 unk i − εnk h∂2 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 h∂1 unk | (Hk + εnk ) |∂2 unk i, (8.17)
~c(2π)2 BZ
where Hk is the effective Hamiltonian acting on the u’s, i.e. Hk = e−ik·r Heik·r .
In order to get the total magnetization we still have to add the second term
M2 in Eq. (8.2). Using Streda’s 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 h∂1 unk | (Hk + εnk − 2µ) |∂2 unk i. (8.18)
~c(2π)2 BZ
This is the ultimate formula for orbital magnetization at zero temperature. The
finite temperature formula is in Ref. [182], 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, and—without
proof here—to 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 [174, 175, 169]. In the Chern
insulating case Eq. (8.18) was heuristically guessed and numerically validated
in Ref. [175].
95
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. [176].
96
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 physical meaning.
97
98
Bibliography
[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).
99
[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, Berry’s Phase and Geometric Quantum Distance: Macroscopic
Polarization and Electron Localization, Lecture Notes for the “Troisième
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 Aharonov–Bohm 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. Öpik, 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).
100
[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).
101
[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] L. D. Landau and E. M. Lifshitz, Electrodynamics of Continuous Media
(Pergamon Press, Oxford, 1984).
[89] J. D. Jackson, Classical Electrodynamics (Wiley, New York, 1975).
[90] C. Kittel, Introduction to Solid State Physics, 7th. edition (Wiley, New
York, 1996).
102
[91] R. M. Martin, Phys. Rev. B 9, 1998 (1974).
[92] M. Posternak, A. Baldereschi, A. Catellani and R. Resta, Phys. Rev. Lett.
64, 1777 (1990).
[93] R. Resta, Ferroelectrics 136, 51 (1992).
[94] R. D. King-Smith and D. Vanderbilt, Phys. Rev. B 47, 1651 (1993).
[95] G. Ortı́z and R. M. Martin, Phys. Rev. B 43, 14202 (1994).
[96] R. Resta, Phys. Rev. Lett. 80, 1800 (1998).
[97] http://www.uam.es/departamentos/ciencias/fismateriac/siesta/.
[98] http://www.cpmd.org/.
[99] R. Resta, Modelling Simul. Mater. Sci. Eng. 11, R69 (2003).
[100] W. H. Duan and Z. R. Liu, Curr. Opin. Solid State Mater. Sci. 10, 40
(2006).
[101] 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).
[102] M.P. Marder, Condensed Matter Physics (Wiley, New York, 2000).
[103] G. Grosso and G. Pastori-Parravicini, Solid State Physics (Academic, San
Diego, 2000).
[104] I. Souza, J. Ìñı̀guez, and D. Vanderbilt, Phys. Rev. Lett. 89, 117602
(2002).
[105] P. Umari and A. Pasquarello, Phys. Rev. Lett. 89, 157602 (2002).
[106] R. Resta, M. Posternak, and A. Baldereschi, Phys. Rev. Lett. 70, 1010
(1993).
[107] G. H. Wannier, Phys. Rev. 52, 191 (1937).
[108] N. Marzari and D. Vanderbilt, Phys. Rev. B 56, 12847 (1997).
[109] A. A. Mostofi, Y.-S. Lee, I. Souza, D. Vanderbilt, and N. Marzari, Comput.
Phys. Commun. 178, 685 (2008).
[110] J. R. Yates, C. J. Pickard, and F. Mauri, Phys. Rev. B 76, 024401 (2007).
[111] R. Resta, J. Chem. Phys. 124, 104104 (2006).
[112] E. I. Blount, in Solid State Physics, edited by H. Ehrenreich, F. Seitz and
D. Turnbull, vol 13 (Academic, New York, 1962), p. 305.
[113] O. F. Mossotti, Memorie di Matematica e di Fisica della Società Ital-
iana delle Scienze Residente in Modena, 24, 49 (1850); R. Clausius, Die
Mechanische Behandlung der Electrica (Vieweg, Berlin, 1879).
[114] A. Pasquarello and R. Resta, Phys. Rev. B 68, 174302 (2003).
103
[115] I. Dabo, B. Kozinsky, N. E. Singh-Miller, and N. Marzari, Phys. Rev. B
77 , 115139 (2008); ibid. 84, 159910(E) (2011).
[116] 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.
[117] Q. Niu, Phys. Rev. 33, 5368 (1986).
[118] D. Vanderbilt and R. D. King-Smith, Phys. Rev. B 48, 4442 (1993).
[119] K. N. Kudin, R. Car, and R. Resta, J. Chem. Phys. 127, 194902 (2007).
[120] M. Stengel, Phys. Rev. B 84, 205432 (2011).
[121] P. L. Silvestrelli, M. Bernasconi, and M. Parrinello, Chem. Phys. Lett.
277, 478 (1997).
[122] A. Pasquarello and R. Car, Phys. Rev. Lett. 79, 1766 (1997).
[123] X. Gonze, Ph. Ghosez, and R. W. Godby, Phys. Rev. Lett. 74, 4035
(1995).
[124] S. Baroni, S. de Gironcoli, A. Dal Corso, and P. Giannozzi, Rev. Mod.
Phys. 73, 515 (2001).
104
[139] 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).
[140] D. Forster, Hydrodynamic Fluctuations, Broken Symmetry, and Correla-
tion Functions (Benjamin, Reading, 1975).
[141] R. Resta, Phys. Rev. Lett. 96, 137601 (2006).
[142] C. Sgiarovello, M. Peressi, and R. Resta, Phys. Rev. 64, 115202 (2001).
[143] M. Veithen, X. Gonze, and Ph. Ghosez, Phys. Rev. B 66, 235113 (2002).
[144] N. D. M. Hine and W. M. C. Foulkes, J. Phys: Condens. Matter 19,
506212 (2007).
[145] A. Monari, G. L. Bendazzoli, and S. Evangelisti, J. Chem. Phys. 129,
134104 (2008).
[146] C. Aebischer, D. Baeriswyl, and R. M. Noack , Phys. Rev. Lett. 86, 468
(2001).
[147] V. Vetere, A. Monari, G.L. Bendazzoli, S. Evangelisti, and B. Paulus, J.
Chem. Phys. 128, 214701 (2008).
[148] L. Stella, C. Attaccalite, S. Sorella, and A. Rubio, Phys. Rev. B 84, 245117
(2011).
[149] G. L. Bendazzoli, S. Evangelisti, A. Monari, and R. Resta, J. Chem. Phys.
133, 064703 (2010).
[150] E. Akkermans, J. Math. Phys. 38, 1781 (1997).
[151] W. Kohn, Phys. Rev. Lett. 2, 393 (1959).
[152] J. des Cloizeaux, Phys. Rev. 135, A685 (1964); ibid. 135, A697 (1964).
[153] S. Ismail-Beigi and T.A. Arias, Phys. Rev. Lett. 82, 2127 (1999).
[154] L. He and D. Vanderbilt, Phys. Rev. Lett. 86, 5341 (2001).
[155] G. F. Giuliani and G. Vignale, Quantum Theory of the Electron Liquid
(Cambridge University Press, Cambridge, 2005).
[156] W. Kohn, Phys. Rev. Lett. 76, 3168 (1996).
[157] S.F. Boys, Rev. Mod. Phys. 32, 296 (1960); J.M. Foster and S.F. Boys,
ibid. 300.
[158] J. G. Àngyàn, Int. J. Quantum Chem. 109, 2340 (2009).
[159] J. G. Àngyàn, Curr. Org. Chem. 15, 3609 (2011).
[160] R. Resta and S. Sorella, Phys. Rev. Lett. 87, 4738 (1995).
[161] T. Wilkens and R. M. Martin, Phys. Rev. B 63, 235108 (2001).
[162] B. Kramer and A. MacKinnon, Rep. Prog. Phys. 56, 1469 (1993).
105
[163] D. J. Thouless, Phys. Rep. 13, 93 (1974).
[164] A. Lagendijk, B. van Tiggelen, and D. S. Wiersma, Phys. Today 62(8),
24 (2009).
[165] F. D. M. Haldane, Phys. Rev. Lett. 61, 2015 (1988).
[166] T. Thonhauser and D. Vanderbilt, Phys. Rev. B 74, 235111 (2006).
[167] D. J. Thouless, J. Phys. C 17, L325 (1984).
[168] C. Brouder, G. Panati, M. Calandra, Ch. Mourougane, and N. Marzari,
Phys. Rev. Lett. 98, 046402 (2007).
[169] D. Ceresoli and R. Resta, Phys. Rev. B 76, 012405 (2007).
[170] Q. Niu, D. J. Thouless, and Y. S. Wu, Phys. Rev. B 31, 3372 (1985).
[171] N. Hao et al., Phys. Rev. B 78, 075438 (2008).
[172] S. Coh and D. Vanderbilt, Phys. Rev. Lett. 102, 107603 (2009).
[173] R. Bianco and R. Resta, Phys. Rev. B 84, 241106(R) (2011).
[174] T. Thonhauser, D. Ceresoli, D. Vanderbilt, and R. Resta, Phys. Rev. Lett.
95, 137205 (2005).
[175] D. Ceresoli, T. Thonhauser, D. Vanderbilt, and R. Resta, Phys. Rev. B
74, 024408 (2006).
[176] R. Bianco and R. Resta, Phys. Rev. Lett. 110, 087202 (2013).
[177] A. A. Soluyanov and D. Vanderbilt, Phys. Rev, B 83, 035108 (2011).
106