You are on page 1of 135

Theory of Superconductivity

c Carsten Timm 2017


Version: October 4, 2017

Wintersemester 2017/2018 TU Dresden Institute of Theoretical Physics


Version: October 4, 2017
LATEX & Figures: S. Lange and C. Timm
Contents

1 Introduction 5
1.1 Scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Books . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2 Basic experiments 7
2.1 Conventional superconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Superfluid helium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3 Unconventional superconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.4 Bose-Einstein condensation in dilute gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

3 Bose-Einstein condensation 13

4 Normal metals 18
4.1 Electrons in metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.2 Semiclassical theory of transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

5 Electrodynamics of superconductors 23
5.1 London theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
5.2 Rigidity of the superfluid state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
5.3 Flux quantization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
5.4 Nonlocal response: Pippard theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

6 Ginzburg-Landau theory 30
6.1 Landau theory of phase transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
6.2 Ginzburg-Landau theory for neutral superfluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
6.3 Ginzburg-Landau theory for superconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
6.4 Type-I superconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
6.5 Type-II superconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

7 Superfluid and superconducting films 52


7.1 Superfluid films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
7.2 Superconducting films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

8 Origin of attractive interaction 70


8.1 Reminder on Green functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
8.2 Coulomb interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
8.3 Electron-phonon interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
8.4 Effective interaction between electrons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

3
9 Cooper instability and BCS ground state 82
9.1 Cooper instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
9.2 The BCS ground state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

10 BCS theory 89
10.1 BCS mean-field theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
10.2 Isotope effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
10.3 Specific heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
10.4 Density of states and single-particle tunneling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
10.5 Ultrasonic attenuation and nuclear relaxation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
10.6 Ginzburg-Landau-Gorkov theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

11 Josephson effects 107


11.1 The Josephson effects in Ginzburg-Landau theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
11.2 Dynamics of Josephson junctions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
11.3 Bogoliubov-de Gennes Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
11.4 Andreev reflection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

12 Unconventional pairing 120


12.1 The gap equation for unconventional pairing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
12.2 Cuprates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
12.3 Pnictides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
12.4 Triplet superconductors and He-3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

4
1

Introduction

1.1 Scope
Superconductivity is characterized by a vanishing static electrical resistivity and an expulsion of the magnetic field
from the interior of a sample. We will discuss these basic experiments in the following chapter, but mainly this
course is dealing with the theory of superconductivity. We want to understand superconductivity using methods
of theoretical physics. Experiments will be mentioned if they motivate certain theoretical ideas or support or
contradict theoretical predictions, but a systematic discussion of experimental results will not be given.
Superconductivity is somewhat related to the phenomena of superfluidity (in He-3 and He-4) and Bose-Einstein
condensation (in weakly interacting boson systems). The similarities are found to lie more in the effective low-
energy description than in the microscopic details. Microscopically, superfluidity in He-3 is most closely related
to superconductivity since both phenomena involve the condensation of fermions, whereas in He-4 and, of course,
Bose-Einstein condensates it is bosons that condense. We will discuss these phenomena briefly.
The course assumes knowledge of the standard material from electrodynamics, quantum mechanics I, and
thermodynamics and statistics. We will also use the second-quantization formalism (creation and annihilation
operators), which are usually introduced in quantum mechanics II. A prior course on introductory solid state
physics would be useful but is not required. Formal training in many-particle theory is not required, necessary
concepts and methods will be introduced (or recapitulated) as needed.

1.2 Overview
This is a maximal list of topics to be covered, not necessarily in this order:
Basic experiments
Review of Bose-Einstein condensation
Electrodynamics: London and Pippard theories
Ginzburg-Landau theory, Anderson-Higgs mechanism
Vortices, Kosterlitz-Thouless transition
Origin of the electron-electron attraction
Cooper instability and BCS theory
Consequences of Cooper pairing: Thermodynamics, tunneling, nuclear relaxation
Josephson effects
Bogoliubov-de Gennes equation for inhomogeneous superconductors
Unconventional superconductivity, cuprate and pnictide superconductors
Andreev scattering and Andreev bound states
Topological superconductors

5
1.3 Books
There are many textbooks on superconductivity and it is recommended to browse a few of them. None of them
covers all the material of this course. M. Tinkhams Introduction to Superconductivity (McGraw-Hill, 2nd edition,
1996) is well written and probably has the largest overlap with this course.
A much broader, deeper, and more modern book is Superfluid States of Matter by B. Svistunov, E. Babaev,
and N. Prokofev (CRC Press, 2015). It certainly goes beyond this course in many places.
A classic book is J. R. Schrieffers Theory of Superconductivity (Addison-Wesley, 1983). It is deep than
Tinkhams book. Schrieffer uses methods from many-particle theory, which are, however, introduced in the book.
The presentation is not always pedagogical.
Many books on solid-state theory or on many-particle theory contain chapters on superconductivity at various
levels. One of them is Many-Body Quantum Theory in Condensed Matter Physics by H. Bruus and K. Flensberg
(Oxford University Press, 2004). The authors use a modern, pedagogical style.

6
2

Basic experiments

In this chapter we will review the essential experiments that have established the presence of superconductivity,
superfluidity, and Bose-Einstein condensation in various materials classes. Experimental observations that have
helped to elucidate the detailed properties of the superconductivity or superfluidity in specific systems are not
covered; some of them are discussed in later chapters.

2.1 Conventional superconductors


After H. Kamerlingh Onnes had managed to liquify Helium, it became for the first time possible to reach tem-
peratures low enough to achieve superconductivity in some chemical elements. In 1911, he found that the static
(dc) resistivity of mercury abruptly fell to zero at a critical temperature Tc of about 4.1 K.

normal metal

superconductor
0
0 Tc T

In a normal metal, the resistivity decreases with decreasing temperature but saturates at a finite value for T 0.
The most stringent bounds on the resisitivity can be obtained not from direct measurement but from the decay
of persistant currents, or rather from the lack of decay. A current set up (by induction) in a superconducting ring
is found to persist without measurable decay after the electromotive force driving the current has been switched
off.
B

7
Assuming exponential decay, I(t) = I(0) et/ , a lower bound on the decay time is found. From this, an upper
bound of . 1025 m has been extracted for the resistivity. For comparison, the resistivity of copper at room
temperature is Cu 1.7 108 m.
The second essential observation was that superconductors not only prevent a magnetic field from entering
this will be discussed in Sec. 5.1but actively expel the magnetic field from their interior. This was observed by
W. Meiner and R. Ochsenfeld in 1933 and is now called the Meiner or Meiner-Ochsenfeld effect.

B=0

From the materials relation B = H with the permeability = 1 + 4 and the magnetic suceptibility (note
that we are using Gaussian units) we thus find = 0 and = 1/4. Superconductors are diamagnetic since
< 0. What is more, they realize the smallest (most diamagnetic) value of consistent with thermodynamic
stability. The field is not just diminished but completely expelled. They are thus perfect diamagnets.
It costs energy to make the magnetic field nonuniform although the externally applied field is uniform. It is
plausible that at some externally applied magnetic field Hc (T ) Bc (T ) this cost will be so high that there is no
advantage in forming a superconducting state. For typical conventional superconductors, the experimental phase
diagram in the temperature-magnetic field plane looks like this:
H
normal metal
Hc (T= 0)
1s
to
rd
er

superconductor 2nd order

Tc T

To specify which superconductors discovered after Hg are conventional, we need a definition of what we want
to call conventional superconductors. There are at least two inequivalent but often coinciding definitions:
Conventional superconductors
show a superconducting state of trivial symmetry (essentially, the superconductor does not break any
symmetry beyond a global U(1) phase symmetry),
result from an attractive interaction between electrons in which phonons play a dominant role.
Conventional superconductivity was observed in quite a lot of elements at low temperatures. The record critical
temperature for elements are Tc = 9.3 K for Nb under ambient pressure and Tc = 20 K for Li under high
pressure. Superconductivity is in fact rather common in the periodic table, 53 pure elements show it under some
conditions. Many alloys and intermetallic compounds were also found to show conventional superconductivity
according to the above criteria. Of these, for a long time Nb3 Ge had the highest known Tc of 23.2 K. But it is
now thought that MgB2 (Tc = 39 K, discovered to be superconducting in 2001) and a few related compounds are
also conventional superconductors in the above sense. They nevertheless show some interesting properties. The
rather high Tc = 39 K of MgB2 is interesting since it is on the order of the maximum Tcmax 30 K expected for
phonon-driven superconductivity. To increase Tc further, the interaction between electrons and phonons would

8
have to be stronger, which however would make the material unstable towards a charge density wave. MgB2
would thus be an optimal conventional superconductor.
Superconductivity with comparably high Tc has also been found in fullerites, i.e., compounds containing
fullerene anions. The record Tc in this class is at present Tc = 38 K for body-centered cubic Cs3 C60 under
pressure. Superconductivity in fullerites was originally thought to be driven by phonons with strong molecular
vibration character but there is recent evidence that it might be unconventional (not phonon driven).
The newest development is the discovery of superconductivity in hydrogen sulfide at up to 203 K under a
high pressure of above 150 GPa (corresponding to about 1.5 million standard atmospheres), see Drozdov et al.,
Nature 525, 73 (2015). The responsible compound is thought to be H3 S forming a body-centered cubic lattice.
Amazingly, superconductivity in this system appears to be phonon driven and in this sense conventional. The
presence of the light hydrogen atoms is crucial for this. There is an ongoing search for high-Tc superconductivity
in other hydrogen compounds.

2.2 Superfluid helium


In 1937, P. Kapiza and independently Allen and Misener discovered that helium shows a transition at Tc = 2.17 K
under ambient pressure, below which it flows through narrow capilaries without resistance. The analogy to
superconductivity is obvious but here it was the viscosity instead of the resistivity that dropped to zero. The
phenomenon was called superfluidity. It was also observed that due to the vanishing viscosity an open container
of helium would empty itself through a flow in the microscopically thin wetting layer.

He

On the other hand, while part of the liquid flows with vanishing viscosity, another part does not. This was shown
using torsion pendulums of plates submerged in helium. For T > 0 a temperature-dependent normal component
oscillates with the plates.

He

Natural atmospheric helium consists of 99.9999 % He-4 and only 0.0001 % He-3, the only other stable isotope.
The observed properties are thus essentially indistinguishable from those of pure He-4. He-4 atoms are bosons
since they consist of an even number (six) of fermions. For weakly interacting bosons, A. Einstein predicted in
1925 that a phase transition to a condensed phase should occur (Bose-Einstein condensation). The observation
of superfluidity in He-4 was thus not a suprisein contrast to the discovery of superconductivitybut in many
details the properties of He-4 were found to be different from the predicted Bose-Einstein condensate. The reason
for this is that the interactions between helium atoms are actually quite strong. For completeness, we sketch the
temperature-pressure phase diagram of He-4:

9
p (10 6 Pa)
solid
3 phases

normal
superfluid liquid
1

ambient critical endpoint


gas
0
0 1 2 3 4 5 T (K)

The other helium isotope, He-3, consists of fermionic atoms so that Bose-Einstein condensation cannot take
place. Indeed no superfluid transition was observed in the temperature range of a few Kelvin. It then came
as a big suprise when superfluidity was finally observed at much lower temperatures below about 2.6 mK by D.
Lee, D. Osheroff, and R. Richardson in 1972. In fact they found two new phases at low temperatures. (They
originally misinterpreted them as possible magnetic solid phases.) Here is a sketch of the phase diagram, note
the temperature scale:

p (10 6 Pa)
solid phases
superfluid
3 phase
A

2 superfluid
phase B normal
liquid

ambient
0
0 1 2 3 T (mK)

Superfluid He-3 shows the same basic properties as He-4. But unlike in He-4, the superfluid states are sensitive
to an applied magnetic field, suggesting that the states have non-trivial magnetic properties.

2.3 Unconventional superconductors


By the late 1970s, superconductivity seemed to be a more or less closed subject. It was well understood based on
the BCS theory and extensions thereof that dealt with strong interactions. It only occured at temperatures up to
23.2 K (Nb3 Ge) and thus did not promise widespread technological application. It was restricted to non-magnetic
metallic elements and simple compounds. This situation started to change dramatically in 1979. Since then,
superconductivity has been observed in various materials classes that are very different from each other and from
the typical low-Tc superconductors known previously. In many cases, the superconductivity was unconventional
and often Tc was rather high. We now give a brief and incomplete historical overview.

10
In 1979, Frank Steglich et al. observed superconductivity below Tc 0.5 K in CeCu2 Si2 . This material is not
a normal metal in its normal state. Instead is is a heavy-fermion metal. The electrons at the Fermi energy
have strong Ce f -orbital character. The very strong Coulomb repulsion between electrons in the f -shell
leads to a high effective mass m  me at the Fermi energy, hence the name. Since then, superconductivity
has been found in various other heavy-fermion compounds. BCS theory cannot explain superconductivity
in these highly correlated metals. Nuclear magnetic resonance (discussed below) and other experimental
techniques have shown that many of these heavy-fermion superconductors show unconventional symmetry
of the superconducting state.
Also in 1979, D. Jrome et al. (Klaus Bechgaards group) observed superconductivity in an organic salt
called (TMTSF)2 PF2 with Tc = 1.1 K. Superconductivity has since been found in various organic materials
with a maximum Tc of about 18 K. The symmetry of the superconducting state is often unconventional.
(We do not include fullerites under organic compounds since they lack hydrogen atoms.)
While the previously mentioned discoveries showed that superconductivity can occur in unexpected materials
classes and probably due to unconventional mechanisms, the Tc values did not surpass the Tc 23 K of
Nb3 Ge. In 1986, J. G. Bednorz and K. A. Mller observed superconductivity in La2x Bax CuO4 (the layered
perovskite cuprate La2 CuO4 with some Ba substituted for La) with Tc on the order of 35 K. In the following
years, many other superconductors based on the same type of nearly flat CuO2 planes sketched below were
discovered. The record transition temperatures for cuprates are Tc = 138 K for Hg0.8 Tl0.2 Ba2 Ca2 Cu3 O8+
at ambient pressure and Tc = 164 K for HgBa2 Ca2 Cu3 O8+ under high pressure. The high Tc values as
well as many experimental probes show that the cuprates are unconventional superconductors. High-Tc
superconductivity in the cuprates was historically important since many advanced methods of many-body
theory have been developed motivated by the desire to understand this phenomenon. We will come back to
this materials class below.

Cu O

In 1991, A. F. Hebard et al. found that the fullerite K3 C60 = (K )3 C3


60 became superconducting below
Tc = 18 K. Tc in this class has since been pushed to Tc = 33 K for Cs2 RbC60 at ambient pressure and
Tc = 38 K for (b.c.c., while all the other known superconducting fullerites are f.c.c.) Cs3 C60 under high
pressure. The symmetry of the superconducting state appears to be trivial but, as noted above, there is an
ongoing debate on whether the pairing is phonon-mediated.
In 2001, Nagamatsu et al. reported superconductivity in MgB2 with Tc = 39 K. The high Tc and the
layered crystal structure, reminiscant of cuprates, led to the expectation that superconductivity in MgB2 is
unconventional. However, most experts now think that it is actually conventional, as noted above.
A new series of important discoveries started in 2006, when Kamihara et al. (H. Hosonos group) observed
superconductivity with Tc 4 K in LaFePO, another layered compound. While this result added a new
materials class based on Fe2+ to the list of superconductors, it did not yet cause much excitement due to
the low Tc . However, in 2008, Kamihara et al. (the same group) found superconductivity with Tc 26 K
in LaFeAsO1x Fx . Very soon thereafter, the maximum Tc in this iron-pnictide class was pushed to 55 K.
Superconductivity was also observed in several related materials classes, some of them not containing oxygen
(e.g., LiFeAs) and some with the pnictogen (As) replaced by a chalcogen (e.g., FeSe). The common structural
element is a flat, square Fe2+ layer with a pnictogen or chalcogen sitting alternatingly above and below
the centers of the Fe squares. Superconductivity is thought to be unconventional. The mechanism is not
necessarily the same for all members of this diverse family.

11
Fe As

In 2015, Drozdov et al. reported superconductivity in H3 S under high pressure with a maximum critical
temperature of Tc = 203 K, as noted above. Although not completely understood yet, superconductivity is
thought to be due to phonons and in this sense conventional.

2.4 Bose-Einstein condensation in dilute gases


An important related breakthrough was the realization of a Bose-Einstein condensate (BEC) in a highly diluted
and very cold gas of atoms. In 1995, Anderson et al. (C. E. Wiman and E. A. Cornells group) reported conden-
sation in a dilute gas of Rb-87 below Tc = 170 nK (!). Only a few months later, Davis et al. (W. Ketterles group)
reported a BEC of Na-23 containing many more atoms. About a year later, the same group was able to create
two condensates and then merge them. The resulting interference effects showed that the atoms where really in a
macroscopic quantum state, i.e., a condensate. All observations are well understood from the picture of a weakly
interacting Bose gas. Bose-Einstein condensation will be reviewed in the following chapter.

12
3

Bose-Einstein condensation

In this short chapter we review the theory of Bose-Einstein condensation. While this is not the correct theory
for superconductivity, at least in most superconductors, it is the simplest description of a macroscopic quantum
condensate. This concept is central also for superconductivity and superfluidity.
We consider an ideal gas of indistinguishable bosons. Ideal means that we neglect any interaction and also
any finite volume of the particles. There are two cases with completely different behavior depending on whether
the particle number is conserved or not. Rb-87 atoms are bosons (they consist of 87 nucleons and 37 electrons)
with conserved particle number, whereas photons are bosons with non-conserved particle number. Photons can
be freely created and destroyed as long as the usual conservation laws (energy, momentum, angular momentum,
. . . ) are satisfied. Bosons without particle-number conservation show a Planck distribution,
1
nP (E) = (3.1)
eE 1
with := 1/kB T , for a grand-canonical ensemble in equilibrium. Note the absence of a chemical potential, which
is due to the non-conservation of the particle number. This distribution function is an analytical function of
temperature and thus does not show any phase transitions.
The situation is different for bosons with conserved particle number. We want to consider the case of a given
number N of particles in contact with a heat bath at temperature T . This calls for a canonical description (N, T
given). However, it is easier to use the grand-canonical ensemble with the chemical potential given. For large
systems, fluctuations of the particle number become small so that the descriptions are equivalent. However,
must be calculated from the given N .
The grand-canonical partition function is
Y  Y 1
Z= 1 + e(i ) + e2(i ) + = , (3.2)
i i
1 e(i )

where i counts the single-particle states of energy i in a volume V . The form of Z expresses that every state
can be occupied not at all, once, twice, etc. For simplicity, we assume the volume to be a cube with periodic
boundary conditions. Then the states can be enumerated be wave vectors k compatible with these boundary
conditions. Introducing the fugacity
y := e , (3.3)
we obtain
Y 1
Z= (3.4)
1 yek
k
X 1 X
ln 1 yek .

ln Z = ln 
= (3.5)
1 ye k
k k

13
The fugacity has to be chosen to give the correct particle number
!
X X 1 X 1
N= hnk i = = . (3.6)
e(k ) 1 y 1 ek 1
k k k

Since hnk i must be non-negative, must satisfy

k k. (3.7)

For a free particle, the lowest possible eigenenergy is k = 0 for k = 0 so that we obtain 0.
For a large volume V , the allowed vectors k become dense and we can replace the sums over k by integrals
according to
Z
d3 k
Z
X 2V 3/2
V 3
= 3 (2m) d  (3.8)
(2) h
k 0

In the last equation we have used the density of states (DOS) of free particles in three dimensions. This replace-
ment contains a fatal mistake, though. The DOS for  = 0 vanishes so that any particles in the state with k = 0,
k = 0 do not contribute to the results. But that is the ground state! For T = 0 all bosons should be in this
state. We thus expect incorrect results at low temperatures.
Our mistake was that Eq. (3.8) does not hold if the fraction of bosons in the k = 0 ground state is macroscopic,
i.e., if N0 /N := hn0 i /N remains finite for large V . To correct this, we treat the k = 0 state explicitly (the same
would be necessary for any state with macroscopic occupation). We write
Z
X 2V
(2m)3/2 d  + (k = 0 term). (3.9)
h3
k 0

Then
Z Z
2V  by parts 2V 2 3/2
ln Z = 3 (2m)3/2 (2m)3/2

d  ln 1 ye ln(1 y) = d ln(1 y)
h h3 3 y 1 e 1
0 0
(3.10)
with
Z Z
2V 1 1 2V  y
N = 3 (2m)3/2 d  1  + = 3 (2m)3/2 d + . (3.11)
h y e 1 y 1 1 h y 1 e 1 1 y
0 0

Defining
Z
1 xn1
gn (y) := dx (3.12)
(n) y 1 ex1
0

for 0 y 1 and n R, and the thermal wavelength


s
h2
:= , (3.13)
2mkB T

we obtain
V
ln Z = g5/2 (y) ln(1 y), (3.14)
3
V y
N = 3 g3/2 (y) + . (3.15)
| {z } | {z y}
1
=: N =: N0

14
We note the identity

X yk
gn (y) = , (3.16)
kn
k=1

which implies

gn (0) = 0, (3.17)

X 1
gn (1) = = (n) for n > 1 (3.18)
kn
k=1

with the Riemann zeta function (x). Furthermore, gn (y) increases monotonically in y for y [0, 1[.
gn ( y)
g3/2

2
g5/2

1 goo

0
0 0.5 1 y

We now have to eliminate the fugacity y from Eqs. (3.14) and (3.15) to obtain Z as a function of the particle
number N . In Eq. (3.14), the first term is the number of particles in excited states (k > 0), whereas the second
term is the number of particles in the ground state. We consider two cases: If y is not very close to unity
(specifically, if 1 y  3 /V ), N0 = y/(1 y) is on the order of unity, whereas N is an extensive quantity. Thus
N0 can be neglected and we get
V
N = N = 3 g3/2 (y). (3.19)

Since g3/2 (3/2) 2.612, this equation can only be solved for the fugacity y if the concentration satisfies

N (3/2)
. (3.20)
V 3
To have 1 y  3 /V in the thermodynamic limit we require, more strictly,
N (3/2)
< . (3.21)
V 3
Note that 3 T 3/2 increases with decreasing temperature. Hence, at a critical temperature Tc , the inequality
is no longer fulfilled. From
N ! (3/2)
= 3/2 (3.22)
V h2
2mkB Tc

we obtain 2/3
h2

1 N
kB Tc = 2/3
. (3.23)
[(3/2)] 2m V
If, on the other hand, y is very close to unity, N0 cannot be neglected. Also, in this case we find
V V
N = g3/2 (y) = 3 g3/2 (1 O(3 /V )), (3.24)
3

15
where O(3 /V ) is a correction of order 3 /V  1. Thus, by Taylor expansion,

V V
N = g3/2 (1) O(1) = 3 (3/2) O(1). (3.25)
3
The intensive term O(1) can be neglected compared to the extensive one so that

V
N
= 3 (3/2). (3.26)

This is the maximum possible value at temperature T . Furthermore, N0 = y/(1 y) is solved by

N0 1
y= = . (3.27)
N0 1 1 + 1/N0

For y to be very close to unity, N0 must be N0  1. Since


V
N0 = N N
= N 3 (3/2) (3.28)

must be positive, we require

N (3/2)
> (3.29)
V 3
T < Tc . (3.30)

We conclude that the fraction of particles in excited states is


3/2
3 (Tc )

N V T
= (3/2) = = . (3.31)
N N 3 3 (T ) Tc

The fraction of particles in the ground state is then


 3/2
N0 T
=1 . (3.32)
N Tc

In summery, we find in the thermodynamic limit


(a) for T > Tc :
N N0
= 1,  1, (3.33)
N N
(b) for T < Tc :
 3/2  3/2
N T N0 T
= , =1 . (3.34)
N Tc N Tc

1
N0 / N

N / N

0
0 0.5 1 T / Tc

16
We find a phase transition at Tc , below which a macroscopic fraction of the particles occupy the same single-
particle quantum state. This fraction of particles is said to form a condensate. While it is remarkable that
Bose-Einstein condensation happens in a non-interacting gas, the BEC is analogous to the condensate in strongly
interacting superfluid He-4 and, with some added twists, in superfluid He-3 and in superconductors.
We can now use the partition function to derive equations of state. As an example, we consider the pressure
kB T
p= =+ kB T ln Z = 3 g5/2 (y) (3.35)
V V
( is the grand-canonical potential). We notice that only the excited states contribute to the pressure. The term
ln(1 y) from the ground state drops out since it is volume-independant. This is plausible since particles in
the condensate have vanishing kinetic energy.
For T > Tc , we can find y and thus p numerically. For T < Tc we may set y = 1 and obtain
kB T
p= (5/2) T 5/2 . (3.36)
3
Remember that for the classical ideal gas at constant volume we find

p T. (3.37)

For the BEC, the pressure drops more rapidly since more and more particles condense and thus no longer
contribute to the pressure.

17
4

Normal metals

To be able to appreciate the remarkable poperties of superconductors, it seems useful to review what we know
about normal conductors.

4.1 Electrons in metals


Let us ignore electron-electron Coulomb interaction and deviations from a perfectly periodic crystal structure
(due to defects or phonons) for now. Then the exact single-particle states are described by Bloch wavefunctions

k (r) = uk (r) eikr , (4.1)

where uk (r) is a lattice-periodic function, is the band index including the spin, and ~k is the crystal momentum
in the first Brillouin zone. Since electrons are fermions, the average occupation number of the state |ki with
energy k is given by the Fermi-Dirac distribution function
1
nF (k ) = . (4.2)
e(k ) + 1
If the electron number N , and not the chemical potential , is given, has to be determined from
X 1
N= , (4.3)
k
e(k ) +1

cf. our discussion for ideal bosons. In the thermodynamic limit we again replace

d3 k
X Z
V . (4.4)
(2)3
k

Unlike for bosons, this is harmless for fermions, since any state can at most be occupied once so that macroscopic
occupation of the single-particle ground state cannot occur. Thus we find

N X Z dk 3 1
= 3 e(k ) + 1
. (4.5)
V
(2)

If we lower the temperature, the Fermi function nF becomes more and more step-like. For T 0, all states with
energies k EF := (T 0) are occupied (EF is the Fermi energy), while all states with k > EF are empty.
This Fermi sea becomes fuzzy for energies k EF at finite temperatures but remains well defined as long as
kB T  EF . This is the case for most materials we will discuss.
The chemical potential, the occupations nF (k ), and thus all thermodynamic variables are analytic functions
of T and N/V . Thus there is no phase transition, unlike for bosons. Free fermions represent a special case with
only a single band with dispersion k = ~2 k 2 /2m. If we replace m by a material-dependent effective mass, this

18
gives a reasonable approximation for simple metals such as alkali metals. Qualitatively, the conclusions are much
more general.
Lattice imperfections and interactions result in the Bloch waves k (r) not being exact single-particle eigen-
states. (Electron-electron and electron-phonon interactions invalidate the whole idea of single-particle states.)
However, if these effects are in some sense small, they can be treated pertubatively in terms of scattering of
electrons between single-particle states |ki.

4.2 Semiclassical theory of transport


We now want to derive an expression for the current in the presence of an applied electric field. This is a question
about the response of the system to an external perturbation. There are many ways to approach this type of
question. If the perturbation is small, the response, in our case the current, is expected to be a linear function
of the perturbation. This is the basic assumption of linear-response theory. In the framework of many-particle
theory, linear-response theory results in the Kubo formula (see lecture notes on many-particle theory). We here
take a different route. If the external perturbation changes slowly in time and space on atomic scales, we can use
a semiclassical description. Note that the following can be derived cleanly as a limit of many-particle quantum
theory.
The idea is to consider the phase space distribution function (r, k, t). This is a classical concept. From
quantum mechanics we know that r and p = ~k are subject to the uncertainty principle rp ~/2. Thus
distribution functions that are localized in a phase-space volume smaller than on the order of ~3 violate quantum
mechanics. On the other hand, if is much broader, quantum effects should be negligible.
The Liouville theorem shows that satisfies the continuity equation
d
+ r + k =0 (4.6)
t r k dt
(phase-space volume is conserved under the classical time evolution). Assuming for simplicity a free-particle
dispersion, we have the canonical (Hamilton) equations
H p ~k
r = = = , (4.7)
p m m
1 1 H 1 1
k = p = = V = F (4.8)
~ ~ r ~ ~
with the Hamiltonian H and the force F. Thus we can write
 
~k F
+ + = 0. (4.9)
t m r ~ k
This equation is appropriate for particles in the absence of any scattering. For electrons in a uniform and time-
independent electric field we have
F = eE. (4.10)
Note that we always use the convention that e > 0. It is easy to see that
 
eE
(r, k, t) = f k + t (4.11)
~
is a solution of Eq. (4.9) for any differentiable function f . This solution is uniform in real space (/r 0)
and shifts to larger and larger momenta ~k for t . It thus describes the free acceleration of electrons in an
electric field. There is no finite conductivity since the current never reaches a stationary value. This is obviously
not a correct description of a normal metal.
Scattering will change as a function of time beyond what as already included in Eq. (4.9). We collect all
processes not included in Eq. (4.9) into a scattering term S[]:
 
~k F
+ + = S[]. (4.12)
t m r ~ k

19
This is the famous Boltzmann equation. The notation S[] signifies that the scattering term is a functional of .
It is generally not simply a function of the local density (r, k, t) but depends on everywhere and at all times,
to the extent that this is consistent with causality.
While expressions for S[] can be derived for various cases, for our purposes it is sufficient to employ the
simple but common relaxation-time approximation. It is based on the observation that (r, k, t) should relax to
thermal equilibrium if no force is applied. For fermions, the equilibrium is 0 (k) nF (k ). This is enforced by
the ansatz
(r, k, t) 0 (k)
S[] = . (4.13)

Here, is the relaxation time, which determines how fast relaxes towards 0 .
If there are different scattering mechanisms that act independently, the scattering integral is just a sum of
contributions of these mechanisms,
S[] = S1 [] + S2 [] + . . . (4.14)
Consequently, the relaxation rate 1/ can be written as (Matthiessens rule)
1 1 1
= + + ... (4.15)
1 2
There are three main scattering mechanisms:
Scattering of electrons by disorder: This gives an essentially temperature-independent contribution, which
dominates at low temperatures.
Electron-phonon interaction: This mechanism is strongly temperature-dependent because the available
phase space shrinks at low temperatures. One finds a scattering rate
1
T 3. (4.16)
e-ph

However, this is not the relevant rate for transport calculations. The conductivity is much more strongly
affected by scattering that changes the electron momentum ~k by a lot than by processes that change it
very little. Backscattering across the Fermi sea is most effective.

ky

k = 2 kF

kx

Since backscattering is additionally suppressed at low T , the relevant transport scattering rate scales as
1
trans T 5. (4.17)
e-ph

Electron-electron interaction: For a parabolic free-electron band its contribution to the resistivity is actually
zero since Coulomb scattering conserves the total momentum of the two scattering electrons and therefore
does not degrade the current. However, in a real metal, umklapp scattering can take place that conserves the
total momentum only modulo a reciprocal lattice vector. Thus the electron system can transfer momentum
to the crystal as a whole and thereby degrade the current. The temperature dependence is typically
1
umklapp
T 2. (4.18)
e-e

20
We now consider the force F = eE and calculate the current density
d3 k ~k
Z
j(r, t) = e (r, k, t). (4.19)
(2)3 m
To that end, we have to solve the Boltzmann equation
 
~k eE 0
+ = . (4.20)
t m r ~ k

We are interested in the stationary solution (/t = 0), which, for a uniform field, we assume to be spacially
uniform (/r = 0). This gives
eE 0 (k) (k)
(k) = (4.21)
~ k
eE
(k) = 0 (k) + (k). (4.22)
~ k
We iterate this equation by inserting it again into the final term:
 
eE eE eE
(k) = 0 (k) + 0 (k) + (k). (4.23)
~ k ~ k ~ k
To make progress, we assume that the applied field E is small so that the response j is linear in E. Under this
assumption we can truncate the iteration after the linear term,
eE
(k) = 0 (k) + 0 (k). (4.24)
~ k
By comparing this to the Taylor expansion
 
eE eE
0 k + = 0 (k) + 0 (k) + . . . (4.25)
~ ~ k
we see that the solution is, to linear order in E,
 
eE
(k) = 0 k + nF (k+eE /~ ). (4.26)
~

Thus the distribution function is simply shifted in k-space by eE /~. Since electrons carry negative charge, the
distribution is shifted in the direction opposite to the applied electric field.

E
0

eE /h

The current density now reads


d3 k ~k d3 k ~k d3 k ~k eE 0
Z   Z Z
eE
j = e 0 k + = e 0 (k) e . (4.27)
(2)3 m ~ (2)3 m (2)3 m ~ k
| {z }
=0

21
The first term is the current density in equilibrium, which vanishes. In components, we have

e2 X d3 k 0 by parts e2 X d3 k k e2 d3 k
Z Z Z
j = E k = + E 0 = E 0 . (4.28)
m (2)3 k m (2)3 k m (2)3
|{z} | {z }
= =n

Here, the integral is the concentration of electrons in real space, n := N/V . We finally obtain

e2 n !
j= E = E (4.29)
m
so that the conductivity is
e2 n
= . (4.30)
m
This is the famous Drude formula. For the resistivity = 1/ we get, based on our discussion of scattering
mechanisms, !
m m 1 1 1
= 2 = 2 + transport + umklapp . (4.31)
e n e n dis e-ph e-e

large T :
T 5 mostly due
to electron
phonon scattering
residual resistivity
due to disorder

22
5

Electrodynamics of superconductors

Superconductors are defined by electrodynamic propertiesideal conduction and magnetic-field expulsion. It is


thus appropriate to ask how these materials can be described within the formal framework of electrodynamics.

5.1 London theory


In 1935, F. and H. London proposed a phenomenological theory for the electrodynamic properties of superconduc-
tors. It is based on a two-fluid picture: For unspecified reasons, the electrons from a normal fluid of concentration
nn and a superfluid of concentration ns , where nn + ns = n = N/V . Such a picture seemed quite plausible based
on Einsteins theory of Bose-Einstein condensation, although nobody understood how the fermionic electrons
could form a superfluid. The normal fluid is postulated to behave normally, i.e., to carry an ohmic current

jn = n E (5.1)

governed by the Drude law


e2 nn
n = . (5.2)
m
The superfluid is assumed to be insensitive to scattering. As noted in section 4.2, this leads to a free acceleration
of the charges. With the supercurrent
js = e ns vs (5.3)
and Newtons equation of motion
d F eE
vs = = , (5.4)
dt m m
we obtain
js e2 ns
= E. (5.5)
t m
This is the First London Equation. We have assumed that nn and ns are both uniform (constant in space) and
stationary (constant in time). This is a serious restriction of London theory, which will only be overcome by
Ginzburg-Landau theory.
Note that the curl of the First London Equation is

e2 ns e2 ns B
js = E= . (5.6)
t m mc t
This can be integrated in time to give
e2 ns
js = B + C(r), (5.7)
mc
where the last term represents a constant of integration at each point r inside the superconductor. C(r) should
be determined from the initial conditions. If we start from a superconducting body in zero applied magnetic

23
field, we have js 0 and B 0 initially so that C(r) = 0. To describe the Meiner-Ochsenfeld effect, we have
to consider the case of a body becoming superconducting (by cooling) in a non-zero applied field. However, this
case cannot be treated within London theory since we here assume the superfluid density ns to be constant in
time.
To account for the flux expulsion, the Londons postulated that C 0 regardless of the history of the system.
This leads to
e2 ns
js = B, (5.8)
mc
the Second London Equation.
Together with Ampres Law
4 4
B= js + jn (5.9)
c c
(there is no displacement current in the stationary state) we get

4e2 ns 4 4e2 ns 4 B
B= 2
B+ n E = B n . (5.10)
mc c mc2 c t
We drop the last term since we are interested in the stationary state and use an identity from vector calculus,

4e2 ns
( B) + 2 B = B. (5.11)
mc2
Introducing the London penetration depth s
mc2
L := , (5.12)
4e2 ns
this equation assumes the simple form
1
2 B = B. (5.13)
2L
Let us consider a semi-infinite superconductor filling the half space x > 0. A magnetic field Bapl = Hapl = Bapl y
is applied parallel to the surface. One can immediately see that the equation is solved by

B(x) = Bapl y ex/L for x 0. (5.14)

The magnetic field thus decresases exponentially with the distance from the surface of the superconductor. In
bulk we indeed find B 0.
B

Bapl

outside inside

0 L x

The Second London Equation


c
js = B (5.15)
42L
and the continuity equation
js = 0 (5.16)
can now be solved to give
c
js (x) = Bapl z ex/L for x 0. (5.17)
4L

24
Thus the supercurrent flows in the direction parallel to the surface and perpendicular to B and decreases into
the bulk on the same scale L . js can be understood as the screening current required to keep the magnetic field
out of the bulk of the superconductor.
The two London equations (5.5) and (5.8) can be summarized using the vector potential:

e2 ns
js = A. (5.18)
mc
This equation is evidently not gauge-invariant since a change of gauge

A A + (5.19)

changes the supercurrent. (The whole London theory is gauge-invariant since it is expressed in terms of E and
B.) Charge conservation requires js = 0 and thus the vector potential must be transverse,

A = 0. (5.20)

This is called the London gauge. Furthermore, the supercurrent through the surface of the superconducting region
is proportional to the normal conponent A . For a simply connected region these conditions uniquely determine
A(r). For a multiply connected region this is not the case; we will return to this point below.

5.2 Rigidity of the superfluid state


F. London has given a quantum-mechanical justification of the London equations. If the many-body wavefunction
of the electrons forming the superfluid is s (r1 , r2 , . . . ) then the supercurrent in the presence of a vector potential
A is
Z
1 X h  e   e  i
js (r) = e d3 r1 d3 r2 (r rj ) s pj + A(rj ) s + s pj + A(rj ) s . (5.21)
2m j c c

Here, j sums over all electrons in the superfluid, which have position rj and momentum operator
~
pj = . (5.22)
i rj
Making pj explicit and using the London gauge, we obtain
Z  
e~ X
js (r) = d3 r1 d3 r2 (r rj ) s s s s
2mi j rj rj
e2 X Z
A(r) d3 r1 d3 r2 (r rj ) s s
mc j

e2 ns
Z  
e~ X
3 3
= d r1 d r2 (r rj ) s s s s A(r). (5.23)
2mi j rj rj mc

Now London proposed that the wavefunction s is rigid under the application of a transverse vector potential.
More specifically, he suggested that s does not contain a term of first order in A, provided A = 0. Then,
to first order in A, the first term on the right-hand side in Eq. (5.23) contains the unperturbed wavefunction one
would obtain for A = 0. The first term is thus the supercurrent for A 0, which should vanish due to Ampres
law. Consequently, to first order in A we obtain the London equation
e2 ns
js = A. (5.24)
mc
The rigidity of s was later understood in the framework of BCS theory as resulting from the existence of a
non-zero energy gap for excitations out of the superfluid state.

25
5.3 Flux quantization
We now consider two concentric superconducting cylinders that are thick compared to the London penetration
depth L . A magnetic flux Z
= d2 rB (5.25)

penetrates the inner hole and a thin surface layer on the order of L of the inner cylinder. The only purpose of
the inner cylinder is to prevent the magnetic field from touching the outer cylinder, which we are really interested
in. The outer cylinder is completetly field-free. We want to find the possible values of the flux .
y

Although the region outside of the inner cylinder has B = 0, the vector potential does not vanish. The relation
A = B implies I ZZ
ds A = d2 r B = . (5.26)

By symmetry, the tangential part of A is



A = . (5.27)
2r
The London gauge requires this to be the only non-zero component. Thus outside of the inner cylinder we have,
in cylindrical coordinates,

A= = . (5.28)
2r 2
Since this is a pure gradient, we can get from A = 0 to A = (/2r) by a gauge transformation
A A + (5.29)
with

= . (5.30)
2
is continuous but multivalued outside of the inner cylinder. We recall that a gauge transformation of A must
be accompanied by a transformation of the wavefunction,
 
ieX
s exp (rj ) s . (5.31)
~c j

This is most easily seen by noting that this guarantees the current in Sec. 5.2 to remain invariant under gauge
transformations. Thus the wavefuncion at = 0 (A = 0) and at non-zero flux are related by
   
i e X j e X
s = exp 0
s = exp i 0
j s , (5.32)
~ c j 2 hc j

26
where j is the polar angle of electron j. For 0
s as well as s to be single-valued and continuous, the exponential
factor must not change for j j + 2 for any j. This is the case if

e hc
Z =n with n Z. (5.33)
hc e
We find that the magnetic flux is quantized in units of hc/e. Note that the inner cylinder can be dispensed
with: Assume we are heating it enough to become normal-conducting. Then the flux will fill the whole interior
of the outer cylinder plus a thin (on the order of L ) layer on its inside. But if the outer cylinder is much thicker
than L , this should not affect s appreciably, away from this thin layer.
The quantum hc/e is actually not correct. Based in the idea that two electrons could form a boson that
could Bose-Einstein condense, Onsager suggested that the relevant charge is 2e instead of e, leading to the
superconducting flux quantum
hc
0 := , (5.34)
2e
which is indeed found in experiments.

5.4 Nonlocal response: Pippard theory


Experiments often find a magnetic penetration depth that is significantly larger than the Londons prediction
L , in particular in dirty samples with large scattering rates 1/ in the normal state. Pippard explained this
on the basis of a nonlocal electromagnetic response of the superconductor. The underlying idea is that the
quantum state of the electrons forming the superfluid cannot be arbitrarily localized. The typical energy scale of
superconductivity is expected to be kB Tc . Only electrons with energies  within kB Tc of the Fermi energy can
contribute appreciably. This corresponds to a momentum range p determined by

kB Tc !  p
= = = vF (5.35)
p p =EF m =EF

kB Tc
p = (5.36)
vF
with the Fermi velocity vF . From this, we can estimate that the electrons cannot be localized on scales smaller
than
~ ~vF
x = . (5.37)
p kB Tc
Therefore, Pippard introduced the coherence length
~vF
0 = (5.38)
kB Tc
as a measure of the minimum extent of electronic wavepackets. is a numerical constant of order unity. BCS
theory predicts 0.180. Pippard proposed to replace the local equation

e2 ns
js = A (5.39)
mc
from London theory by the nonlocal

3 e2 ns ( A(r0 )) /0
Z
js (r) = d3 r 0 e (5.40)
40 mc 4
with
= r r0 . (5.41)

27
The special form of this equation was motivated by an earlier nonlocal generalization of Ohms law. The main
point is that electrons within a distance 0 of the point r0 where the field A acts have to respond to it because of
the stiffness of the wavefunction. If A does not change appreciably on the scale of 0 , we obtain

3 e2 ns 3 0 ( A(r)) /0 3 e2 ns ( A(r)) /0
Z Z

js (r) = d r e = d3 e . (5.42)
40 mc 4 40 mc 4

The result has to be parallel to A(r) by symmetry (since it is a vector depending on a single vector A(r)). Thus

3 e2 ns ( A(r))2 /0
Z
js (r) = A(r) d3 e
40 mc 4
Z1 Z
3 e2 ns 2 cos2 /0
= A(r) 2 d(cos ) d 2  4 e
40 mc 


1 0
2 2
3 e ns 2 e ns
= A(r) 2 0 = A(r). (5.43)
40 mc 3 mc
We recover the local London equation. However, for many conventional superconductors, 0 is much larger than
. Then A(r) drops to zero on a length scale of  0 . According to Pippards equation, the electrons respond
to the vector potential averaged over regions of size 0 . This averaged field is much smaller than A at the surface
so that the screening current js is strongly reduced and the magnetic field penetrates much deeper than predicted
by London theory, i.e.,  L .
The above motivation for the coherence length 0 relied on having a clean system. In the presence of strong
scattering the electrons can be localized on the scale of the mean free path l := vF . Pippard phenomenologically
generalized the equation for js by introducing a new length where
1 1 1
= + (5.44)
0 l

( is a numerical constant of order unity) and writing

3 e2 ns ( A(r0 )) /
Z
js = d3 r 0 e . (5.45)
40 mc 4
Note that 0 appears in the prefactor but in the exponential. This expression is in good agreement with
experiments for series of samples with varying disorder. It is essentially the same as the result of BCS theory.
Also note that in the dirty limit l  0 , L we again recover the local London result following the same argument
as above, but since the integral gives a factor of , which is not canceled by the prefactor 1/0 , the current is
reduced to
e2 ns e2 ns l
js = A= A. (5.46)
mc 0 mc 0
Taking the curl, we obtain

e2 ns l
js = B (5.47)
mc 0
4e2 ns l
B= B (5.48)
mc2 0
4e2 ns l
2 B = B, (5.49)
mc2 0
in analogy to the derivation in Sec. 5.1. This equation is of the form
1
2 B = B (5.50)
2

28
with the penetration depth s s s
mc2 0 0
= = L . (5.51)
4e2 ns l l
p
Thus the penetration depth is increased by a factor of order 0 /l in the dirty limit, l  0 .

29
6

Ginzburg-Landau theory

Within London and Pippard theory, the superfluid density ns is treated as given. There is no way to understand
the dependence of ns on, for example, temperature or applied magnetic field within these theories. Moreover, ns
has been assumed to be constant in time and uniform in spacean assumption that is expected to fail close to
the surface of a superconductor.
These deficiencies are cured by the Ginzburg-Landau theory put forward in 1950. Like Ginzburg and Landau
we ignore complications due to the nonlocal electromagnetic response. Ginzburg-Landau theory is developed as
a generalization of London theory, not of Pippard theory. The starting point is the much more general and very
powerful Landau theory of phase transitions, which we will review first.

6.1 Landau theory of phase transitions


Landau introduced the concept of the order parameter to describe phase transitions. In this context, an order
parameter is a thermodynamic variable that is zero on one side of the transition and non-zero on the other. In
ferromagnets, the magnetization M is the order parameter. The theory neglects fluctuations, which means that
the order parameter is assumed to be constant in time and space. Landau theory is thus a mean-field theory. Now
the appropriate thermodynamic potential can be written as a function of the order parameter, which we call ,
and certain other thermodynamic quantities such as pressure or volume, magnetic field, etc. We will always call
the potential the free energy F , but whether it really is a free energy, a free enthalpy, or something else depends
on which quantities are given (pressure vs. volume etc.). Hence, we write
F = F (, T ), (6.1)
where T is the temperature, and further variables have been suppressed. The equilibrium state at temperature
T is the one that minimizes the free energy. Generally, we do not know F (, T ) explicitly. Landaus idea was to
expand F in terms of , including only those terms that are allowed by the symmetry of the system and keeping
the minimum number of the simplest terms required to get non-trivial results.
For example, in an isotropic ferromagnet, the order parameter is the three-component vector M. The free
energy must be invariant under rotations of M because of isotropy. Furthermore, since we want to minimize F
as a function of M, F should be differentiable in M. Then the leading terms, apart from a trivial constant, are

F
= M M + (M M)2 + O (M M)3 .

(6.2)
2
Denoting the coefficients by and /2 is just convention. and are functions of temperature (and pressure
etc.).
What is the corresponding expansion for a superconductor or superfluid? Lacking a microscopic theory,
Ginzburg and Landau assumed based on the analogy with Bose-Einstein condensation that the superfluid part is
described by a single one-particle wave function s (r). They imposed the plausible normalization
Z
2
d3 r |s (r)| = Ns = ns V ; (6.3)

30
Ns is the total number of particles in the condensate. They then chose the complex amplitude of s (r) as the
order parameter, with the normalization
2
|| ns . (6.4)
Thus the order parameter in this case is a complex number. They thereby neglect the spatial variation of s (r)
on an atomic scale.
The free energy must not depend on the global phase of s (r) because the global phase of quantum states is
not observable. Thus we obtain the expansion
2 4 6
F = || + || + O(|| ). (6.5)
2
Only the absolute value || appears since F is real. Odd powers are excluded since they are not differentiable at
= 0. If > 0, which is not guaranteed by symmetry but is the case for superconductors and superfluids, we
can neglect higher order terms since > 0 then makes sure that F () is bounded from below. Now there are two
cases:
If 0, F () has a single minimum at = 0. Thus the equilibrium state has ns = 0. This is clearly a
normal metal (nn = n) or a normal fluid.
If < 0, F () has a ring of minima with equal amplitude (modulus) || but arbitrary phase. We easily see
r
F 3
= 2 || + 2 || = 0 || = 0 (this is a maximum) or || = (6.6)
||

Note that the radicand is positive.


F
>0 =0

<0

0 Re

Imagine this figure rotated around the vertical axis to find F as a function of the complex . F () for < 0
is often called the Mexican-hat potential. In Landau theory, the phase transition clearly occurs when = 0.
Since then T = Tc by definition, it is useful to expand and to leading order in T around Tc . Hence,


= 0 (T Tc ), 0 > 0, (6.7)
const.
= (6.8)

Then the order parameter below Tc satisfies


s s
0 (T Tc ) 0 p
|| = = T Tc . (6.9)

Tc T

31
Note that the expansion of F up to fourth order is only justified as long as is small. The result is thus limited
to temperatures not too far below Tc . The scaling || (T Tc )1/2 is characteristic for mean-field theories. All
solutions with this value of || minimize the free energy. In a given experiment, only one of them is realized. This
eqilibrium state has an order parameter = || ei with some fixed phase . This state is clearly not invariant
under rotations of the phase. We say that the global U(1) symmetry of the system is spontaneously broken since
the free energy F has it but the particular equilibrium state does not. It is called U(1) symmetry since the group
U(1) of unitary 1 1 matrices just contains phase factors ei .

Specific heat
Since we now know the mean-field free energy as a function of temperature, we can calculate further thermody-
namic variables. In particular, the free entropy is
F
S= . (6.10)
T
Since the expression for F used above only includes the contributions of superconductivity or superfluidity, the
entropy calculated from it will also only contain these contributions. For T Tc , the mean-field free energy is
F ( = 0) = 0 and thus we find S = 0. For T < Tc we instead obtain

2 1 2 2
r   

S= F = + =
T T 2 T 2
(0 )2 (0 )2 (0 )2
= (T Tc )2 = (T Tc ) = (Tc T ) < 0. (6.11)
T 2
We find that the entropy is continuous at T = Tc . By definition, this means that the phase transition is continuous,
i.e., not of first order. The heat capacity of the superconductor or superfluid is
S
C=T , (6.12)
T
which equals zero for T Tc but is
(0 )2
C= T (6.13)

for T < Tc . Thus the heat capacity has a jump discontinuity of

(0 )2
C = Tc (6.14)

at Tc . Adding the other contributions, which are analytic at Tc , the specific heat c := C/V is sketched here:
c

tion
con tribu
no rmal
Tc T

Recall that Landau theory only works close to Tc .

32
6.2 Ginzburg-Landau theory for neutral superfluids
To be able to describe also spatially non-uniform situations, Ginzburg and Landau had to go beyond the Landau
description for a constant order parameter. To do so, they included gradients. We will first discuss the simpler
case of a superfluid of electrically neutral particles (think of He-4). We define a macroscopic condensate wave
function (r), which is essentially given by s (r) averaged over length scales large compared to atomic distances.
We expect any spatial changes of (r) to cost energythis is analogous to the energy of domain walls in magnetic
systems. In the spirit of Landau theory, Ginzburg and Landau included the simplest term containing gradients
of and allowed by symmetry into the free energy
Z  
2 4
F [] = d3 r || + || + () , (6.15)
2
where we have changed the definitions of and slightly. We require > 0 so that the system does not
spontaneously become highly non-uniform. The above expression is a functional of , also called the Landau
functional. Calling it a free energy is really an abuse of language since the free energy proper is only the value
assumed by F [] at its minimum.
If we interpret (r) as the (coarse-grained) condensate wavefunction, it is natural to identify the gradient
term as a kinetic energy by writing
Z " 2 #
1 ~
F []
2 4
= d3 r || + || +

, (6.16)
2 2m i

where m is an effective mass of the particles forming the condensate.


From F [] we can derive a differential equation for (r) minimizing F . The derivation is very similar to the
derivation of the Lagrange equation (of the second kind) from Hamiltons principle S = 0 known from classical
mechanics. Here, we start from the extremum principle F = 0. We write
(r) = 0 (r) + (r), (6.17)
where 0 (r) is the as yet unknown solution and (r) is a small displacement. Then
F [0 + ] = F [0 ]
~2 ~2
Z  
3
+ d r 0 + 0 + 0 0 0 + 0 0 0 + (0 ) + () 0
2m 2m
+ O(, )2
by parts
= F [0 ]
~2 ~2 2
Z  
3 2
+ d r 0 + 0 + 0 0 0 + 0 0 0 ( 0 ) 0
2m 2m
+ O(, )2 . (6.18)
If 0 (r) minimizes F , the terms linear in and must vanish for any (r), (r). Noting that and are
linearly independent, this requires the prefactors of and of to vanish for all r. Thus we conclude that
~2 2 2 ~2 2
0 + 0 0 0 0 = 0 + |0 | 0 0 = 0. (6.19)
2m 2m
Dropping the subscript and rearranging terms we find
~2 2
2 + + || = 0.
(6.20)
2m
This equation is very similar to the time-independent Schrdinger equation but, interestingly, it contains a non-
linear term. We now apply it to find the variation of (r) close to the surface of a superfluid filling the half space
x > 0. We impose the boundary condition (x = 0) = 0 and assume that the solution only depends on x. Then
~2 00 2
(x) + (x) + |(x)| (x) = 0. (6.21)
2m

33
Since all coefficients are real, the solution can be chosen real. For x we should obtain the uniform solution
r

lim (x) = . (6.22)
x
Writing r

(x) = f (x) (6.23)

we obtain
~2
f 00 (x) + f (x) f (x)3 = 0. (6.24)
| 2m
{z }
>0

This equation contains a characteristic length with


~2 ~2
2 = = 0
> 0, (6.25)
2m 2m (Tc T )
which is called the Ginzburg-Landau coherence length. It is not the same quantity as the the Pippard coherence
length 0 or . The Ginzburg-Landau has a strong temperature dependence and actually diverges at T = Tc ,
whereas the Pippard has at most a weak temperature dependence. Microscopic BCS theory reveals how the
two quantities are related, though. Equation (6.24) can be solved analytically. It is easy to check that
x
f (x) = tanh (6.26)
2
is a solution satisfying the boundary conditions at x = 0 and x . (In the general, three-dimensional case, the
solution can only be given in terms of Jacobian elliptic functions.) The one-dimensional tanh solution is sketched
here:
(x)

0 x

Fluctuations for T > Tc


So far, we have only considered the state 0 of the system that minimizes the Landau functional F []. This is
the mean-field state. At nonzero temperatures, the system will fluctuate about 0 . For a bulk system we have

(r, t) = 0 + (r, t) (6.27)

with uniform 0 . We consider the cases T > Tc and T < Tc separately.


For T > Tc , the mean-field solution is just 0 = 0. F [] = F [] gives the energy of excitations, we can thus
write the partition function Z as a sum over all possible states of Boltzmann factors containing this energy,
Z
Z = D2 eF []/kB T . (6.28)

The notation D2 expresses that the integral is over uncountably many complex variables, namely the values of
(r) for all r. This means that Z is technically a functional integral. The mathematical details go beyond the

34
scope of this course, though. The integral is difficult to evaluate. A common approximation is to restrict F []
4
to second-order terms, which is reasonable for T > Tc since the fourth-order term (/2) || is not required to
stabilize the theory (i.e., to make F for || ). This is called the Gaussian approximation. It allows Z
to be evaluated by Fourier transformation: Into
Z    
3 1 ~ ~
F [] = d r (r)(r) + (r) (r) (6.29)
2m i i
we insert
1 X ikr
(r) = e k , (6.30)
V k
which gives
Z  
1 X 3 ikr+ik0 r 1 0
F [] = d re k k0 + (~kk ) ~k k0
V 2m
kk0 | {z }
V kk0
X ~2 k 2 ~2 k 2
 X 

= k k + k = + k k . (6.31)
2m k 2m
k k

Thus
!
X 2 2
Z Y 
1 ~ k
Z
= d2 k exp + k k
kB T 2m
k k
Y Z ~2 k 2
   
2 1
= d k exp + k k . (6.32)
kB T 2m
k

The integral is now of Gaussian type (hence Gaussian approximation) and can be evaluated exactly:
Y kB T
Z
= 2 2 . (6.33)
k
+ ~2mk

From this, we can obtain the thermodynamic variables. For example, the heat capacity is
S 2F 2 2 X kB T
C=T = T = T k B T ln Z = kB T T ln 2 2 . (6.34)
T T 2 T 2 T 2 + ~2mk
k

We only retain the term that is singular at Tc . It stems from the temperature dependence of , not from the
explicit factors of T . This term is
2 X ~2 k 2 0
 
X
Ccrit
= kB T 2 2 ln +
= kB T 2 2 2
T
k
2m T
k
+ ~2mk
 2
2
X (0 )2 2 2m 0 2
X 1
= kB T 2 = k B T 2
( ) 2
. (6.35)
~2 k 2 k + 2m
 ~ 2
k + 2m k ~2

Going over to an integral over k, corresponding to the thermodynamic limit V , we obtain


 2
d3 k
Z
2 2m 0 2 1
Ccrit = kB T ( ) V
~2 (2)3 k 2 + 2m2 2

~
kB T 2 (m )2 0 2 1 kB T 2 (m )3/2 0 2 1
= 4
( ) V q = 3
( ) V
2 ~
2m 2 2 ~
2 ~
1
. (6.36)
T Tc

35
the mean-field level, C just showed a step at Tc . Including fluctuations, we obtain a divergence
Recall that at
of the form 1/ T Tc for T Tc from above. This is due to superfluid fluctuations in the normal state. The
system notices that superfluid states exist at relatively low energies, while the mean-field state is still normal.
We note for later that the derivation has shown that for any k we have two fluctuation modes with dispersion

~2 k 2
k = + . (6.37)
2m
The two modes correspond for example to the real and the imaginary part of k . Since = 0 (T Tc ) > 0, the
dispersion has an energy gap .
k

0 kx

In the language of field theory, one also says that the superfluid above Tc has two degenerate massive modes,
with the mass proportional to the energy gap.

Fluctuations for T < Tc


Below pTc , the situation is a bit more complex due to the Mexican-hat potential: All states with amplitude
|0 | = / are equally good mean-field solutions.

F
Im

Re

ringshaped
minimum

It is plausible that fluctuations of the phase have low energies since global changes of the phase do not increase
F . To see this, we write
= (0 + ) ei , (6.38)
where 0 and are now real. The Landau functional becomes
Z 

F []
= d3 r (0 + )2 + (0 + )4
2
   
1 ~ i i ~ i i
+ ()e + ~( 0 + )()e ()e + ~(0 + )()e . (6.39)
2m i i

36
As above, we keep only terms up to second order in the fluctuations (Gaussian approximation), i.e., terms
proportional to 2 , , or 2 . We get
Z "  
1 ~ ~
F [, ] 3 (3 (( 2 3 2
= d r 2
((0(
( +( 2
(
0 + + 30 +
2m i i
  ( (( ! #
1 ~ ( (((~ 1
+ ( ~ (+ (~0 ) +
( ( (~0 ) ~0 (+ const). (6.40)
2m i ( (0( i 2m
((((
Note that the first-order terms cancel since we are expanding about a minimum. Furthermore, up to second order
there are no terms mixing amplitude fluctuations and phase fluctuations . We can simplify the expression:

~2
Z   
3 2 1 ~ ~
F [, ] = d r 2 + ()
2m i i 2m
~2 k 2 ~2 k 2
X   

= 2 + k k k , (6.41)
|{z} 2m 2m k
k >0 | {z }
>0

in analogy to the case T > Tc . We see that amplitude fluctuations are gapped (massive) with an energy gap
2 = 0 (Tc T ). They are not degenerate. Phase fluctuations on the other hand are ungapped (massless) with
quadratic dispersion
~2 k 2
k = . (6.42)
2m
The appearance of ungapped so-called Goldstone modes is characteristic for systems with spontaneously
broken
continuous symmetries. We state without derivation that the heat capacity diverges like C Tc T for
T Tc , analogous to the case T > Tc .

6.3 Ginzburg-Landau theory for superconductors


To describe superconductors, we have to take the charge q of the particles forming the condensate into account.
We allow for the possibility that q is not the electron charge e. Then there are two additional terms in the
Landau functional:
The canonical momentum has to be replaced by the kinetic momentum:
~ ~ q
A, (6.43)
i i c
where A is the vector potential.
The energy density of the magnetic field, B 2 /8, has to be included.
Thus we obtain the functional
"  2 #
2
Z 
1 ~ q B
F [, A] 3 2 4
= d r || + || + A + . (6.44)
2 2m i c 8

Minimizing this free-energy functional with respect to , we obtain, in analogy to the previous section,
 2
1 ~ q 2
A + + || = 0. (6.45)
2m i c

37
To minimize F with respect to A, we write A(r) = A0 (r) + a(r) and inset this into the terms containing A:
Z     
3 1 ~ q q
F [, A0 + a] = F [, A0 ] + d r A 0 a
2m i c c
  
1  q  ~ q 1
+ a A0 + ( A0 ) ( a) + O(a2 )
2m c i c 4 | {z }
= a(A0 ) ((A0 )a)
Z    
3 q ~ ~
= F [, A0 ] + d r + a
2m c i i
2

q 2 1
+ 2 || A0 a + ( B0 ) a + O(a2 ), (6.46)
m c 4

where we have used A0 = B0 and Gauss theorem. At a minimum, the coefficient of the linear term must
vanish. Dropping the subscript we obtain

q~ q2 2 1
i
([ ] ) + 2 || A + B = 0. (6.47)
2m c m c 4
With Ampres Law we find

c q~ q2 2
j= B = i ([ ] ) || A, (6.48)
4 2m m c
where we have dropped the subscript s of j since we assume that any normal current is negligible. Equations
(6.45) and (6.48) are called the Ginzburg-Landau equations.
In the limit of uniform (r), Eq. (6.48) simplifies to
2
q 2 ||
j= A. (6.49)
m c
This should reproduce the London equation
e2 ns
j= A, (6.50)
mc
which obviously requires
2
q 2 || e2 ns
= . (6.51)
m m
As noted in Sec. 5.3, based on flux-quantization experiments and on the analogy to Bose-Einstein condensation,
it is natural to set q = 2e. For m one might then insert twice the effective electron (band) mass of the metal.
However, it turns out to be difficult to measure m independently of the superfluid density ns and it is therefore
common to set m = 2m 2me . All system-specific properties are thus absorbed into
2 2
m q 2 || m e2 || 2
ns = 2
= 2 = 2 || . (6.52)
e m e 2m
With this, we can write the penetration depth as
s s s
mc2 mc2 mc2
= = = . (6.53)
2
4e ns 8e2 ||
2 8e2 (
)

We have found two characteristic length scales:

: penetration of magnetic field,


: variation of superconducting coarse-grained wave function (order parameter).

38
Within the mean-field theories we have so far employed, both quantities scale as
1
, (6.54)
Tc T
close to Tc . Thus their dimensionless ratio
(T )
:= (6.55)
(T )
is roughly temperature-independent. It is called the Ginzburg-Landau parameter and turns out to be very im-
portant for the behaviour of superconductors in an applied magnetic field. For elemental superconductors, is
small compared to one.

Fluctuations and the Anderson-Higgs mechanism


When discussing fluctuations about the mean-field state of a superconductor, we have to take the coupling to the
electromagnetic field into account. We proceed analogously to the case of a neutral superfluid and employ the
Gaussian approximation throughout. For T > Tc , we note that the kinetic term
  2
1 ~ q

A (6.56)
2m i c

is explicitly of second order in the fluctuations = so that electromagnetic field fluctuations appear only
in higher orders. Thus to second order, the order-parameter fluctuations decouple from the electromagnetic
fluctuations,
" 2 #
Z
1 ~ B2 X ~2 k 2
 X B Bk
3
k
F [, B] = d r (r)(r) + + = + k k + . (6.57)
2m i 8 2m 8
k k

The superconducting fluctuations consist of two degenerate massive modes with dispersion k = + ~2 k 2 /2m ,
as for the neutral superfluid.
The case T < Tc is more interesting. Writing, as above, = (0 + ) ei , we obtain to second order
Z "  
3 (3 (( 2 2 2 1 ~ ~
F [, , A] = d r ( 2
(0( + 20 + + 30 +
( ( (
2m i i
  ( ( ( !
1 ~ (()(((~ 1
+
~ 0(( +((~ 0 + (~0 ) ~0
2m i((( ( ( ( i 2m
(
(
   ( (((
1 ~ q   q
( ( (  ~
+ ( (0(+((
A A0
2m i (((((( c c i
((
1   q  q  

+
(~0 ) A0 + A0 ~0
2m c c #
1  q   q  1
+ A0 A0 + ( A) ( A) (+ const)
2m c c 8
Z   
1 ~ ~
= d3 r 2 2 +
2m i i
2

~  q   q  1
A A + ( A) ( A) . (6.58)
2m ~c ~c 8

Note that the phase of the macroscopic wave function and the vector potential appear in the combination

39
(q/~c)A. Physics is invariant under the gauge transformation

A A + , (6.59)
1
, (6.60)
c
eiq/~c , (6.61)

where is the scalar electric potential and (r, t) is an arbitrary scalar field. We make use of this gauge invariance
by choosing
~c
= . (6.62)
q
Under this tranformation, we get
~c
A A =: A0 , (6.63)
q
= (0 + ) ei (0 + ) ei(+) = 0 + . (6.64)

The macroscopic wave function becomes purely real (and positive). The Landau functional thus tranforms into
Z   
1 ~ ~
F [, A0 ] d3 r 2 2 +

2m i i
2

q 0 0 1 0 0
(A ) A + ( A ) ( A ) (6.65)
2m c2 8

(note that A0 = A). Thus the phase no longer appears in F ; it has been absorbed into the vector potential.
Furthermore, dropping the prime,
X  ~2 k 2 q2
 
1
F [, A] = 2 + k k A Ak + (k Ak ) (k Ak )
2m 2m c2 k 8
k
X  ~2 k 2 q2 k2
 
1
= 2 + k k A Ak + A Ak (k Ak )(k Ak ) . (6.66)
2m 2m c2 k 8 k 8
k

Obviously, amplitude fluctuations decouple from electromagnetic fluctuations and behave like for a neutral super-
fluid. We discuss the electromagnetic fluctuations further. The term proportional to is due to superconductivity.
Without it, we would have the free-field functional
X 1 
k 2 Ak Ak (k Ak )(k Ak ) .

Ffree [A] = (6.67)
8
k

Decomposing A into longitudinal and transverse components,

Ak = k(k Ak ) + Ak k(k Ak ) (6.68)


| {z } | {z }
k
=: Ak =: A
k

with k := k/k, we obtain


X 1 h k k

k

k
i
Ffree [A] = k 2 Ak Ak + k 2 A
k Ak k Ak k Ak
8
k
X 1 h k k k k
i X 1
Ak + k 2 A
k 2 A
 
= k 2Ak k 2
k Ak  Ak Ak = k Ak . (6.69)
8  8
k k

Thus only the transverse components appearonly they are degrees of freedom of the free electromagnetic field.
They do not have an energy gap.

40
Including the superconducting contribution, we get
X  q2 k k q2 1 2


F [A] = A Ak A Ak + k Ak Ak . (6.70)
2m c2 k 2m c2 k 8
k

All three components of A appear now, the longitudinal one has been introduced by absorbing the phase (r).
Even more importantly, all components obtain a term with a constant coefficient, i.e., a mass term. Thus the
electromagnetic field inside a superconductor becomes massive. This is the famous Anderson-Higgs mechanism.
The same general idea is also thought to explain the masses of elementary particles, although in a more complicated
way. The Higgs bosons in our case are the amplitude-fluctuation modes described by . Contrary to what is
said in popular discussions, they are not responsible for giving mass to the field A. Rather, they are left over
when the phase fluctuations are eaten by the field A.
The mass term in the superconducting case can be thought of as leading to the Meiner effect (finite penetration
depth ). Indeed, we can write
X ~2 k 2
 X 1 1 
2
F [, A] = 2 + k k + A Ak + k Ak Ak (k Ak )(k Ak ) . (6.71)
2m 8 2 k
k k

The photon mass is proportional to 1/. (To see that the dispersion relation is ~2 2 = m2 c4 + p2 c2 , we would
have to consider the full action including a temporal integral over F and terms containing time-derivatives.)

Elitzurs theorem
One sometimes reads that in the superconducting state gauge symmetry is broken. This is not correct. Gauge
symmetry is the invariance under local gauge transformations. S. Elitzur showed in 1975 that a local gauge
symmetry cannot be spontaneously broken. Rather, superconductors and superfluids spontaneously break a
global U(1) symmetry in that the ordered state has a prefered macroscopic phase, as noted above.

6.4 Type-I superconductors


Superconductors with small Ginzburg-Landau parameter are said to be of type I. The exact condition is
1
< . (6.72)
2
It turns out that these superconductors have a uniform state in an applied magnetic field, at least for simple
geometries such as a thin cylinder parallel to the applied field.
The appropriate thermodyamic potential for describing a superconductor in an applied magnetic field is the
Gibbs free energy G (natural variable H) and not the Helmholtz free energy F (natural variable B), which we
have used so far. They are related by the Legendre transformation
HB
Z
G = F d3 r . (6.73)
4

For a type-I superconductor, the equilibrium state inpthe bulk is uniform (we will see later why this is not a
trivial statement). The order parameter is then || = / and the magnetic induction B as well as the vector
potential (in the London gauge) vanish. Thus the Gibbs free-energy density is

2 4 2 2 2
gs = fs = || + || = + = . (6.74)
2 2 2
On the other hand, in the normal state vanishes, but the field penetrates the system and B H so that

HB H2 H2 H2
gn = fn = = . (6.75)
4 8 4 8

41
The system will only become superconducting if this reduces the Gibbs free energy, i.e., if

2 H2
gs gn = + < 0. (6.76)
2 8
Thus in an applied magnetic field of H Hc , with the critical field
s
2
Hc (T ) := 4 , (6.77)

superconductivity does not occur.


We can use the relation (6.77) and

mc2 mc2
2 (T ) = 2 = (6.78)
8e2 || 8e2

to express the phenomenological parameters and in terms of the measurable quantities and Hc :

2e2 2 2
= Hc , (6.79)
mc2
 2 2
2e
= 4 4 Hc2 . (6.80)
mc2

Domain-wall energy and intermediate states


We can now calculate the energy per area of a domain wall between superconducting and normal regions. We
assume that (r) is only a function of x and impose the boundary conditions
( p
:= / for x ,
(x) (6.81)
0 for x .

What are reasonable boundary conditions for the local field B(x)? In the superconductor we have

B(x) 0 for x . (6.82)

At the other end, if we have limx B(x) > Hc , the bulk superconductor has a positive Gibbs free-energy density,
as seen above. Thus superconductivity is not stable. For limx B(x) < Hc , the Gibbs free-energy density of
the superconductor is negative and it eats up the normal phase. Thus a domain wall can only be stable for
s
2
B(x) Hc = 4 for x . (6.83)

Technically, we have to solve the Ginzburg-Landau equations under these boundary conditions, which can only
be done numerically. The qualitative form of the solution is clear, though:

B(x) ~
(x)

0 x
~

42
The Gibbs free energy of the domain wall, per unit area, will be denoted by . To derive it, first note that for the
given boundary conditions, gs (x ) = gn (x ), i.e., the superconducting free-energy density deep inside
the superconductor equals the normal free-energy density deep inside the normal conductor. This bulk Gibbs
free-energy density is, as derived above,

Hc2 H2
gs (x ) = gn (x ) = fn (x ) = c. (6.84)
8 8
The additional free-energy density due to the domain wall is

H2 H2
 
gs (x) c = gs (x) + c . (6.85)
8 8
The corresponding free energy per area is
Z Z
Hc2 Hc2
   
B(x)Hc
= dx gs (x) + = dx fs (x) +
8 4 8

Z  2
B2 Hc2
  
2 4 1 ~ q BHc
= dx || + || + i x c A(x) + 8 4 + 8 .
(6.86)
2 2m

| {z }
= (BHc )2 /8

We can simplify this by multiplying the first Ginzburg-Landau equation (6.45) by and integrating over x:
Z "  2 #
2 1 4 ~ q
dx || + || + A(x)
2m i x c

Z      
by parts 2 4 1 ~ q ~ q
= dx || + || + A(x) (x) A(x)
2m i x c i x c

Z "   2 #
2 1 ~
4 q
= dx || + || +
A(x) = 0. (6.87)
2m i x c

Thus
Z Z " 2 #
(B Hc )2 Hc2
  
4 4 B
= dx || + = dx 2 || + 1
2 8 8 Hc

Z " 2 4
#
Hc2 B ||
= dx 1 4
, (6.88)
8 Hc

where we have drawn out the characteristic energy density Hc2 /8. The domain wall energy is given by the differ-
ence of the energy cost of expelling the magnetic field and the energy gain due to superconducting condensation.
For strong type-I superconductors,  ; there is a region of thickness > 0 in which the first term is
already large, while the second only slowly approaches its bulk value (see sketch above). Thus > 0 for type-I
superconductors. They therefore tend to minimize the total area of domain walls.
Note that even in samples of relatively simple shape, magnetic flux will penetrate for non-zero applied field. It
will do so in a way that minimizes the total Gibbs free energy, only one contribution to which is the domain-wall
energy. For example in a very large slab perpendicular to the applied field, the flux has to penetrate in some
manner, since going around the large sample would cost much energy. A careful analysis shows that this will
usually happen in the form of normal stripes separated by superconducting stripes, see Tinkhams book. Such a
state is called an intermediate state. It should not be confused with the vortex state to be discussed later.

43
S N S N S N S

z
x

6.5 Type-II superconductors


Type-II superconductors are defined by a large Ginzburg-Landau parameter
1
> . (6.89)
2
The analysis in the previous section goes through. But now domain walls have a region where the condensate is
nearly fully developed but the flux is not completely expelled.

B(x) (x)

0 x
~

Therefore, the domain-wall energy is negative, < 0. Hence, the system tends to maximize the total area of
domain walls. This tendency should be counterbalanced by some other effect, otherwise the system would become
inhomogeneous on microscopic (atomic) length scales and Ginzburg-Landau theory would break down. The effect
in question is flux quantizationthe penetrating magnetic flux cannot be split into portions smaller than the flux
quantum 0 = hc/2e, see Sec. 5.3.

Fluxoid quantization
We revisit the quantization of magnetic flux. Consider an arbitrary closed path S forming the edge of a surface
S, where at least the edge S must lie inside a superconductor,

44
The magnetic flux through this loop is
Z Z I
Stokes
= da B = da ( A) = ds A. (6.90)
S S S

With the second Ginzburg-Landau equation

q~ q2 2
j=i ([ ] ) || A (6.91)
2m m c
we obtain
( )
m c j
I I  
~c mc ~c
= ds 2 +i 2 ([ ] ) = ds 2 j + i (2i)
q 2 || 2q || e ns 2q
S S
I I
mc ~c
= ds vs ds , (6.92)
e 2e
S S

where vs := j/ens is the superfluid velocity. The last term contains the charge of the phase of the macroscopic
wave function along the path, which must be an integer multiple of 2 for (r) to be continuous:
I
ds = 2n, n Z. (6.93)
S

(The minus sign is conventional.) Thus we find for the so-called fluxoid
I
mc ~c hc
0 := ds vs = 2n = n = n0 , n Z. (6.94)
e 2e 2e
S

We see that it is not the flux but the fluxoid 0 that is quantized. However, deep inside a superconducting region
the current vanishes and 0 equals .

Vortices
The smallest amount of fluxoid that can penetrate a superconductor is one flux quantum 0 . It does so as a
vortex (or vortex line).
vortex core
B

Following the above arguments, the phase of changes by 2 as one circles a vortex in the positive direction,
where, by convention, the direction of a vortex is the direction of the magnetic field B. Thus in the center of the
vortex line (the vortex core), the phase is undefined. This is only consistent with a continuous if = 0 in the

45
vortex core. For a vortex along the z-axis we choose cylindrical coordinates %, , z. We then have to solve the
Ginzburg-Landau equations with the boundary conditions

(% = 0) = 0, (6.95)
|(% )| = 0 , (6.96)
B(% ) = 0. (6.97)

From symmetry, we have

(r) = || (%) ei = || (%) ei , (6.98)


vs (r) = vs (%), (6.99)
B(r) = zB(%) (6.100)

and we can choose


A(r) = A(%) (6.101)
where
1
B(%) = %A(%). (6.102)
% %
Choosing a circular integration path of radius r centered at the vortex core, the enclosed fluxoid is
I
mc mc !
0 (r) = (r) ds vs = 2rA(r) 2r vs (r) = 0
e e
S
mc 0
A(r) vs (r) = . (6.103)
e 2r
This relation follows from fluxoid quantization and thus ultimately from the second Ginzburg-Landau equation.
To obtain j(r), (r), and A(r), one also has to solve the first Ginzburg-Landau equation
 2
1 ~ 2e 2
+ A + + || = 0 (6.104)
4m i c

and Ampres Law


c
j= B. (6.105)
4
2
This cannot be done analytically because of the non-linear term || . For small distances % from the core,
one can drop this term, thereby linearizing the Ginzburg-Landau equation. The solution is still complicated, the
result is that || increases linearly in %. Numerical integration of the full equations gives the results sketched here
for a cut through the vortex core:

~ (r)

B(r)

0 x
~

46
Another useful quantity is the free energy per unit length of a vortex line (its line tension). An analytical
expression can be obtained in the strong type-II case of  1. We only give the result here: The vortex line
tension is
 2
0
v ln
4
H2
= c 4 2 ln . (6.106)
8
We can now calculate the field for which the first vortex enters the superconductor, the so-called lower critical field
Hc1 . This happens when the Gibbs free energy for a superconductor without vortices (Meiner phase) equals the
Gibbs free energy in the presence of a single vortex. We assume the sample to have cross section A and thickness
L, parallel to the applied field.
B

Then we have the condition

0 = Gone vortex Gno vortex


Z
1
=F
s + L v d3 r H B 
F
s
4
Z
Hc1
= Lv d3 r B(r)
4
Hc1 L
= Lv 0 . (6.107)
4
Thus
4v
Hc1 = . (6.108)
0
The line tension in Eq. (6.106) can also be written as

0 Hc H ln
v ln = 0 c (6.109)
4 2 4 2

so that
Hc ln
Hc1 = . (6.110)
2
Recall that this expression only holds for  1. Hc is the thermodynamic critical field derived above. In a
type-II superconductor, nothing interesting happens at H = Hc .

The Abrikosov vortex lattice


We have considered the structure of an isolated vortex line. How does a finite magnetic flux penetrate a type-II
superconductor? Based on Ginzburg-Landau theory, A. A. Abrikosov proposed in 1957 that flux should enter as
a periodic lattice of parallel vortex lines carrying a single flux quantum each. He proposed a square lattice, which
was due to a small mistake. The lowest-free-energy state is actually a triangular lattice.

47
y
B

As noted above, the magnetic flux starts to penetrate the superconductor at the lower critical field Hc1 . Further-
more, since flux expulsion in type-II superconductors need not be perfect, they can withstand stronger magnetic
fields than type-I superconductors, up to an upper critical field Hc2 > Hc , Hc1 .
H
Hc2 (T )
normal metal
Shubnikov
(vortex)

2n
phase

do
rde
r
Hc1 (T )
2nd
Meiner ord
er
phase
Tc T

We will now review the basic ideas of Abrikosovs approach. Abrikosovs results are quantitatively valid only
close to Hc2 since he assumed the magnetic flux density B to be uniform, which is valid for

 l, (6.111)

where r
0
l= (6.112)
B
is the typical distance between vortices (B/0 is the two-dimensional concentration of vortex lines). For B =
B z = const = H z we can choose the gauge A = y Hx. Then the first Ginzburg-Landau equation becomes (note
m = 2m)
 2
1 ~ 2eH 2
+ yx + + || = 0. (6.113)
2m i c
Slightly below Hc2 , || is expected to be small (this should be checked!) so that we can neglect the non-linear
term. Introducing the cyclotron frequency of a superconductor,
2eH
c := , (6.114)
m c
we obtain
~2 2
 
1 2 2
i~c x + m c x (r) = (r). (6.115)
2m y 2 |{z}
>0

This looks very much like a Schrdinger equation and from the derivation it has to be the Schrdinger equation
for a particle of mass m and charge q = 2e in a uniform magnetic field H. This well-known problem is solved
by the ansatz
(x, y) = eiky y f (x). (6.116)

48
We obtain
~2 2
 
iky y 1 2 2 iky y
e i~c x + m c x e f (x)
2m y 2
~2 d2 f ~2 ky2 1
= 2 + f i~c x iky f + m c2 x2 f
2m dx 2m 2 !
~2 d2 f 1 2 2 ~ky ~2 ky2
= 2 + m c x + 2 + f = f. (6.117)
2m dx 2 m c (m )2 c2

Defining x0 := ~ky /m c , we get

~2 d2 f 1
+ m c2 (x x0 )2 f = f. (6.118)
2m dx2 2
This is the Schrdinger equation for a one-dimensional harmonic oscillator with shifted minimum. Thus we obtain
solutions fn (x) as shifted harmonic-oscillator eigenfunctions for
 
1
= ~c n + , n = 0, 1, 2, . . . (6.119)
2
Solution can only exist if
~c ~eH
= 0 (T Tc ) = 0 (Tc T ) = . (6.120)
2 m c
Keeping T Tc fixed, a solution can thus only exist for H Hc2 with the upper critical field
m c
Hc2 = . (6.121)
~e
Using 2 = ~2 /2m , we obtain
~c 0
Hc2 (T ) = 2
= . (6.122)
2e (T ) 2 2 (T )

Note that since 1/ Tc T close to Tc , Hc2 (T ) sets in linearly, as shown in the above sketch. Hc2 should be
compared to the thermodynamic critical field Hc ,
 2
2 mc2 1 ~2 c2 1 1 hc 1
Hc = 2 2 = 2 2 2
= 2
2e 8e 8 2e 2
2

Hc2 Hc2
Hc = 0 = =
2 2 2 2

Hc2 = 2 Hc . (6.123)

For = 1/ 2, Hc2 equals Hc . This is the transition between a type-II and a type-I superconductor.
So far, our considerations have not told us what the state for H . Hc2 actually looks like. To find out,
one in principle has to solve the full, not the linearized, Ginzburg-Landau equation. (We have seen that the
linearized equation is equivalent to the Schrdinger equation for a particle in a uniform magnetic field. The
solutions are known: The eigenfunctions have uniform amplitude and the eigenenergies form discrete Landau
levels, very different from what is observed in the Shubnikov phase.) Abrikosov did this approximately using a
variational approach. He used linear combinations of solutions of the linearized equation as the variational ansatz
and assumed that the solution is periodic in two dimensions, up to a plane wave.
We call the periods in the x- and y-directions ax and ay , respectively. The function

n (x, y) = eiky y fn (x) (6.124)

has the period ay in y if


2
ky = q, q Z. (6.125)
ay

49
Then the harmonic oscillator is centered at
~ 2 0
x0 = q= q. (6.126)
m c ay Hay
Since the lowest-energy solution is obtained for n = 0 (the ground state of the harmonic oscillator), Abrikosov
only considered the n = 0 solutions
2 !
1 m c
    
2y 2y 0
0 (x, y) = exp iq f0 (x) = C exp iq exp x+ q . (6.127)
ay |{z} ay 2 ~ Hay
normalization | {z }
x0

In the Gauss function we find the quantity


~ ~c ~c ~2
= = = 2 (T ), (6.128)
m c 2eH 2e Hc2 2m
as long as H Hc2 . Thus
   2 !
2y 1 0
0 (x, y) = C exp iq exp 2 x + q . (6.129)
ay 2 Hay

This is a set of functions enumerated by q Z. Abrikosov considered linear combinations


   2 !
X 2y 1 0
(x, y) = Cq exp iq exp 2 x + q . (6.130)
q=
ay 2 Hay

To be periodic in x with period ax , up to a plane wave (the corresponding discussion in Tinkhams book is not
fully correct), this ansatz has to satisfy
   2 !
X 2y 1 0
(x + ax , y) = Cq exp iq exp 2 x + ax + q
q=
ay 2 Hay
    2 !
X 2y 1 0 Hay
= Cq exp iq exp 2 x + q+ ax
q=
ay 2 Hay 0
!
(x, y) x, y. (6.131)
This requires
Hax ay
=: N. (6.132)
0
(Note that this quantity is positive.) Then
   2 !
X 2y 1 0
(x + ax , y) = Cq exp iq exp 2 x + (q + )
q=
ay 2 Hay
  X    2 !
2y 2y 1 0
= exp i Cq exp iq exp 2 x + q . (6.133)
ay q= ay 2 Hay

This equals (x, y) up to a plane-wave factor if


Cq = Cq q, (6.134)
i.e., if Cq is periodic. Abrikosov considered the case = 1, which leads to a square lattice. The lowest-free-energy
solution is obtained for = 2 and C1 = iC0 , which gives a triangular lattice, sketched above. Note that
Hax ay
= (6.135)
0

50
has a simple interpretation: It is the number of flux quanta passing through a rectangular unit cell of the vortex
lattice.
The vortex lattice is a rather complex system: It is a lattice of interacting lines with a line tension v . At
non-zero temperatures, the vortices fluctuate, which can lead to the melting of the vortex lattice. The resulting
vortex liquid can be pictured as a pot of boiling spaghetti, with the constraint that the vortex lines must either
terminate at the surface or form closed loops (!). Moving vortices lead to ohmic resistance, even though most of
the sample is still superconducting. To complicate matters, interaction of vortices with defects (pinning) plays
an important role. There is an extensive research literature on vortex matter, which we cannot review here.

51
7

Superfluid and superconducting films

Generally, fluctuations are stronger in systems of lower dimension. Indeed, they change the properties of two-
dimensional superfluid and superconducting films qualitatively compared to three-dimensional samples. We will
consider such films within the Ginzburg-Landau theory.

7.1 Superfluid films


We start from the two-dimensional Landau functional
Z  
2 2 4
F [] = d r || + || + () . (7.1)
2

We consider temperatures T < Tc . As shown in Sec. 6.2, fluctuations of the amplitude || then have an energy
gap, whereas fluctuations of the phase are ungapped. Not too close to Tc , phase fluctuations will thus dominate
and we negelect amplitude fluctuations, writing
p
(r) = 0 ei(r) with 0 = /. (7.2)

Thus, up to an irrelevant constant,


Z   X  

F [] = 2
d r () = k 2 k k . (7.3)

| {z } k
>0

We can now calculate the correlation function


Z
D E 1
h(r) (0)i = 02 ei(r) ei(0) = 02 D ei(r)+i(0) eF []/kB T . (7.4)
Z

52
This is a Gaussian average since F [] is bilinear in k . Thus we can write

X 1
h(r) (0)i = 02 (i)n h((r) (0))n i
n=0
n!

X 1
= 02 (i)n 1 3 5 (n 1) ((r) (0))2


n=0
n!
n even

n=2m 2
X 1 3 5 (2m 1)

(1)m (0))2

= 0 ((r)
m=0
1 2 3 (2m)

X 1
= 02 (1)m ((r) (0))2


m=0
2 4 (2m)
| {z }
= 2m m!
 m  
X 1 1 1

= 02 ((r) (0))2 = 02 exp ((r) (0))2 .




(7.5)
m=0
m! 2 2

Herein, we have
1 X ikr 0
((r) (0))2 = 1)(eik r 1) hk k0 i


(e
V
kk0
1 1 X ikr 0 kB T 1
= (e 1)(eik r 1) kk0
V 2 0 k 2
kk
1 1X kB T 1
= (2 2 cos k r) . (7.6)
V 2 k 2
k

Going over to the continuum, we obtain

d2 k 1
Z
kB T
((r) (0))2 =

(1 cos k r)
(2)2 k 2
Z2 Z
kB T 1 1
= d dk (1 cos(kr cos ))
(2)2 k
0 0
Z
kB T 1 1
= dk (1 J0 (kr) ). (7.7)
2 k | {z }
0 Bessel function

The k-integral is cut off at , which is the inverse of some microscopic length scale, = 1/r0 . The idea is that
physics at shorter length scales r < r0 have been integrated out to obtain the Landau functional. For large
distances, r  r0 , we approximate, rather brutally,
(
1 for kr < 1
J0 (kr) (7.8)
0 for kr 1

and obtain
1/r
Z 0
kB T 1 dk kB T 1 r
((r) (0))2

= = ln (7.9)
2 k 2 r0
1/r

and thus    
1 kB T 1 r r
h(r) (0)i

= 02 exp ln = 02 (7.10)
2 2 r0 r0

53
with
1 kB T
= > 0. (7.11)
4
Thus the correlation function of the order parameters decays like a power law of distance in two dimensions. We do
not find long-range order, which would imply limr h(r) (0)i = const. This agrees with the Mermin-Wagner
theorem, which forbids long-range order for the two-dimensional superfluid. The power-law decay characterizes
so-called quasi-long-range order (short range order would have an even faster, e.g., exponential, decay).

Isolated vortices
We have argued that fluctuations in the amplitude are less important because they have an energy gap proportional
to 2 > 0. This is indeed true for small amplitude fluctuations. However, there exist variations of the amplitude
that are, while energetically costly, very stable once they have been created. These are vortices. In two dimensions,
a vortex is a zero-dimensional object; the order parameter goes to zero at a single point at its center. The simplest
form of a vortex at the origin can be represented by

(r) = |(r)| ei(r) = |(r)| ei(0 ) , (7.12)

where r and are (planar) polar coordinates of r. An antivortex would be described by

(r) = |(r)| ei(0 ) . (7.13)

In both cases, limr0 |(r)| = 0. Note that we have changed the convention for the sign in the exponent compared
to superconducting vortices.
In the presence of vortices, the phase (r) of the order parameter is multivalued and, of course, undefined at
the vortex centers. On the other hand, the phase gradient v = is single-valued (but still undefined at the
vortex centers). For any closed loop C not touching any vortex cores, we have
I
ds v = total change in phase along C = 2 NC , (7.14)
C

where NC Z is the enclosed winding number or vorticity. The vorticity can be written as the sum of the
vorticities Ni = 1 of all vortices and antivortices inside the loop,
X
NC = Ni . (7.15)
i

N1 = +1
N2 = +1

N3 = 1

Defining the vortex concentration by X


nv (r) := Ni (r Ri ), (7.16)
i

where Ri is the location of the center of vortex i, we obtain



v vy vx = 2 nv (r). (7.17)
x y

54
Note that in two dimensions the curl is a scalar.
Now it is always possible to decompose a vector field into a rotation-free and a divergence-free component,

v = vph + vv (7.18)

with

vph = 0, (7.19)
vv = 0. (7.20)

Since the vortex concentation associated with vph vanishes, the component vph does not contain any vortices.
Alternatively, note that the first equation implies that there exists a single-valued scalar field (r) so that

vph = . (7.21)

is a single-valued component of the phase, which cannot be due to vortices. This is the contribution from small
phase fluctuations, which we have already discussed. Conversely, vv is the vortex part, for which

vv = 2 nv , (7.22)
vv = 0. (7.23)

This already suggests an electrodynamical analogy, but to formulate this analogy it is advantageous to rescale
and rotate the field vv : r

E(r) := 2 z vv (r). (7.24)

| {z }
>0

Then the energy density far from vortex cores, where ||


p
= /, is
 1
1
w = (v ) v = vv vv = 2 E E = E E. (7.25)
2

Also, we find r r

E = 2 ( vv ) = 2 2 nv (7.26)

and
E = 0. (7.27)
These equations reproduce the fundamental equations of electrostatics if we identify the charge density with
r

v = 2 nv . (7.28)

(The factor in Gauss law is 2 instead of 4 since we are considering a two-dimensional system.) We can now
derive the pseudo-electric field E(r) for a single vortex,
I r

da E = 2r E = 2 2 (7.29)

q
2

E(r) = (7.30)
r
and thus q
2

E(r) = r. (7.31)
r

55
From this, we obtain the energy of a single vortex,
Z Z Z
1 1 1 1
E1 = Ecore + d2 r E E = Ecore 2 d2 r 2 = Ecore 2 dr . (7.32)
2 2 r r
2
Now we note that the derivation does not hold for small distances from the vortex center since there || is not
close to /. Thus we cut off the radial integral at the lower end at some vortex core radius r0 and put all
energy contributions from the core into Ecore . r0 is on the order of the coherence length . But the integral still
diverges; if our system has a characteristic size of L, we obtain

ZL
dr L
E1 = Ecore 2 = Ecore 2 ln . (7.33)
r r0
r0 | {z }
>0

Thus the energy of a single, isolated vortex diverges logarithmically with the system size. This suggests that
isolated vortices will never be present as thermal fluctuations as long as < 0. This is not true, though. The
probability of such vortices should be
 
1 1 2 L
p1 2 eE1 /kB T = 2 eEcore /kB T exp ln
r0 r0 kB T r0
  k2T   21
1 L B 1 L
= 2 eEcore /kB T = 2 eEcore /kB T . (7.34)
r0 r0 r0 r0

The typical area per vortex is 1/p1 and the total number of vortices will be, on average,
1
2 2
L2

L
Nv = = L2 p1 = eEcore /kB T . (7.35)
1/p1 r0

For > 1/4, Nv diverges in the thermodynamic limit so that infinitely many vortices are present. For < 1/4,
Nv 0 for L and according to our argument, which is essentially due to Kosterlitz and Thouless, there are
no vortices. It is plausible and indeed true that free vortices destroy quasi-long-range order and in this sense also
superfluidity. Note that
1 kB T ! 1
= = (7.36)
4 (T ) 4
is an equation for a critical temperature for the appearance of free vortices. We thus find that the critical
temperature in superfluid films should be reduced from the point where = 0 ( = ) to the one where = 1/4
due to vortices appearing as fluctuations of the order parameter. While qualitatively true, our description is still
incomplete, though, since we have so far neglected interactions between vortices.

Vortex interaction
The energy of two vortices with vorticities 1 can easily be obtained from the electrostatic analogy. We assume
that core regions do not overlap, i.e., the separation is R 2r0 . The pseudo-electric field of the two vortices,
assumed to be located at R/2 = R x/2, is

r R/2
r r
r + R/2
E(r) = 2 2 2
|r R/2| |r + R/2|2
2 2
|r + R/2| (r R/2) |r R/2| (r + R/2)
r
= 2 2 2 . (7.37)
|r R/2| |r + R/2|

56
Thus the energy is
Z
1
E2 = 2Ecore + d2 r
EE
2
!2
2 2
|r + R/2| (r R/2) |r R/2| (r + R/2)
Z
1 2
= 2Ecore 2 d r 2 2
2 |r R/2| |r + R/2|
Z Z

= 2Ecore 2 dx dy

0
4 2 4 2 2 2
|r + R/2| |r R/2| + |r R/2| |r + R/2| 2 |r + R/2| |r R/2| (r2 R2 /4)
4 4 . (7.38)
|r R/2| |r + R/2|

We introduce elliptic coordinates , according to


R
x= , (7.39)
2
Rp 2 p
y= 1 1 2, (7.40)
2
where [1, [, [1, 1]. Then
 2 Z
R 2 2
E2 = 2Ecore 2 d d
2 2 1 1 2
R 4
2 4 2
( + )4 R2 ( )2 + R2 ( )4 R2 ( + )2

2

R 4
4
( )4 R2 ( + )4

2
2 2 2
2 R2 ( + )2 R2 ( )2 R2 ( 2 2 + ( 2 1)(1 2 ) 1)

( + )2 + ( )2 2( 2 + 2 2)
Z
1
= 2Ecore 2 d d
2 1 1 2 2 2
Z
1 4
= 2Ecore 2 d d 2 2
2
1 1 2
Z
1
= 2Ecore 8 d . (7.41)
( 2 1)

We have to keep in mind that the integrals in real space have a lower cutoff r0 . The minimum separation from
the vortex at R x/2 is ( 1)R/2. For this separation to equal r0 , the lower cutoff for must be
2r0
0 = 1 + . (7.42)
R
With this cutoff, we obtain
1 (R + 2r0 )2
E2 = 2Ecore 8 ln , (7.43)
2 4r0 (R + r0 )
which for R  r0 becomes
R
E2 = 2Ecore 4 ln . (7.44)
4r0
We absorb an R-independent constant into Ecore and finally obtain
R
E2 2Ecore + Vint (R) = 2Ecore 4 ln . (7.45)
r0
| {z }
>0

57
Note that the energy of a vortex-antivortex pair remains finite for large system sizes, L . Also, the vortex-
antivortex interaction Vint (R) increases with R, i.e., vortex and antivortex attract each other. Conversely, one
can show that vortices with the same vorticity repel each other.
Vint

0 r0 R

One can also show that for arbitrary vorticities N1 , N2 Z, the interaction reads
R
Vint (R) = 4 N1 N2 ln . (7.46)
r0
| {z }
<0

Since the equations of (pseudo-) electrostatics are linear, the superposition principle applies and we do not have
additional 3-, 4-, etc. body interactions. The energy of a system of vortices is thus
X 1 X
E= Ecore + Ni Nj vint (|ri rj |) (7.47)
i
2
ij,i6=j

with
r
vint (r) := 4 ln , (7.48)
r0
P P
provided that i Ni = 0. If the total vorticity i Ni does not vanish, the energy diverges logarithmically with
the system size, as we have seen. As long as the total vorticity is zero, the energy per vortex is finite and thus
we expect a non-zero concentration of vortices for all temperatures T > 0. We now want to understand the
consequences of their presence.

Berezinskii-Kosterlitz-Thouless theory
The vortices in a two-dimensional superfluid behave like a Coulomb gasa gas of charged particles with Coulomb
interaction, which is logarithmic in 2D. We have seen that the total vorticity (charge) has to vanish. It is therefore
possible to group the vortices into vortex-antivortex pairs. We do this using the following simple algorithm:
1. find the vortex and the antivortex with the smallest separation r,
2. mark this vortex and this antivortex as a pair,
3. repeat these steps for all remaining vortices until none are left.

+
+
+

58
The energy of an isolated vortex-antivortex pair (we will use the term vortex pair from now on) of size r is
r
E2 (r) = 2Ecore 4 ln . (7.49)
r0
Thus the probability density of such pairs is
   2K0
1 E2 (r)/kB T 1 2 r 1 2 r
p2 (r) = 4 e = 4 y0 exp 2K0 ln = 4 y0 , (7.50)
r0 r0 r0 r0 r0

where
y0 := N0 eEcore /kB T (7.51)
is a vortex fugacity (N0 is a constant of order unity) or, more precisely, y02 is a vortex-pair fugacity, and
1
K0 := 2 > 0 (7.52)
kB T
is a dimensionless measure for the interaction strength in units of the thermal energy. K0 is called the stiffness.
The crucial idea is now that smaller pairs are polarized in the pseudo-electric field of the vortex and antivortex
forming a given pair. This leads to the screening of the vortex-antivortex interaction and thus reduces the energy
of large pairs. Formally, this can be described by renormalization-group (RG) theory (Kosterlitz 1974), which we
will summarize in the following. The grand-canonical partition function of the vortex-antivortex system is

Z 2 2
|r |
X 1 Z
d r 1 d r2N 1 X i r j
Z0 = y 2N exp 2 Ni Nj K0 ln , (7.53)
(N !)2 0 r02 r02 2 r0
N D1 D2N i6=j

where we have already implemented the constraint that the number of vortices, N , equals the number of antivor-
tices. The ranges of integration, Di , comprise the two-dimensional space R2 excluding disks of radius r0 centered
at all vortices (and antivortices) with numbers j < i. This means that the minimum separation is r0 . The idea
of RG theory is to perform the integrals for the smallest pairs of sizes between r0 and r0 + dr and rewrite the
result (approximately) in a form identical to Z0 but with changed (renormalized) parameters. Physically, we
thereby omit the smallest pairs and take their effect into account by renormalizing the parameters. In this way,
the partition function, the fugacity, and the stiffness become functions of the smallest length scale, which we now
denote by r. These running quantities are written as Z, y, and K, respectively. Thus

Z 2 2
|r |
X 1 Z
d r 1 d r2N 1 X i rj
Z= y 2N exp 2 Ni Nj K ln . (7.54)
(N !)2 r2 r2 2 r
N D1 D2N i6=j

We now perform the integrals over the smallest separations between r and r + dr. A crucial assumption is that
this only involves vortex-antivortex pairs, not pairs of equal vorticity. This is plausible since a vortex and an
antivortex attract each other. The integration starts by splitting the integrals,
Z Z Z Z
d2 r1 d2 r2N
= d2 r1 d2 r2N
D1 D2N D10 0
D2N
Z Z Z Z
1X
+ Ni ,Nj d2 r1 d2 r2N d2 rj d2 ri . (7.55)
2
i6=j D10 0
D2N Di,j r|ri rj |<r+dr
| {z }
excluding i,j

The Di0 correspond to the Di but with the minimum separation increased to r + dr. Di,j is the full R2 excluding
disks of radius r around all vortices n 6= i, j. Inserting this into the expression for Z, we obtain the integral, from

59
the second term,

|rn rj |
Z Z X
d2 rj d2 ri exp 2 Nn Nj K ln
r
n6=i,j
Di,j r|ri rj |<r+dr

X |rn ri | |rj ri |
2 Nn Nj K ln 2 K ln
r r
n6=i,j

2

Z

X (rm r j ) (r n r j ) r
= d2 rj 2r dr 1 + 2 K 2 Nm Nn 2 2

m,n6=i,j
|r m r j | |rn r j |
Di,j

(rm rj ) (rn rj )
X Z
= 2r dr v + 2 K 2 Nm Nn r2 d2 rj 2 . (7.56)

2
m,n6=i,j
|rm rj | |rn rj |
Di,j

Herein, we have
(rm rj ) (rn rj ) |rm rn |
Z
L
d2 rj 2 2 = 2 ln 2 (1 mn ) ln . (7.57)
|rm rj | |rn rj | r r
Di,j

The first term diverges for L but drops out when the sum over m, n is performed, due to overall vanishing
vorticity. The result is
3 2 2
X |rm rn |
= 2r dr v 2 K r N m N n ln (7.58)
r
m,n6=i,j
m6=n

and with the sum over i, j:



1X X |rm r |
n
= 2r dr N 2 V 2 3 K 2 r2 (N 1)2 Nm Nn ln , (7.59)
2 r
i6=j m6=n

neglecting some terms of order N 0 . Inserting everything into Z, we obtain two terms: The first corresponds to
Z with Di replaced by Di0 and the second reads
X 1  y 2N Z Z
2
d r1 d2 r2N 2r dr
(N !)2 r2
N
D10 0
D2N
| {z }
2N 2 integrals
! !
2 3 2 2
X
2 |rm rn | 1 X |rm rn |
N V 2 K r (N 1) Nm Nn ln exp 2K Ni Nj ln . (7.60)
r 2 r
m6=n i6=j
| {z }
(2N 2)(2N 3) terms

We rename the summation index N as N + 1 in this second term. Then both terms contain 2N integrals under
the sum over N . We also put all terms containing dr into the exponent using 1 + a dr = ea dr . We obtain
   X Z Z
y 2 1  y 2N
Z 0 = exp 2 2 r dr V d 2
r1 d2 r2N
r (N !)2 r2
N
D10 0
D2N

1 X dr

|r i r |
j
exp 2K + 8 4 y 2 K 2 Ni Nj ln . (7.61)
2 r r
i6=j

60
Next, we have to express Z 0 in terms of the new length scale r0 := r + dr. This is only relevant in expressions not
already linear in dr. This applies to, on the one hand,

1 1 1 + 2 drr0
= = , (7.62)
r2 (r0 dr)2 (r0 )2

and, on the other,



1 X 1 X
exp 2K Ni Nj ln r = exp 2K Ni Nj ln(r0 dr)
2 2
i6=j i6=j

1 X 1 X dr
= exp 2K Ni Nj ln r0 exp 2K Ni Nj 0
2 2 r
i6=j i6=j

 2N
1X dr
= exp 2K Ni Nj ln r0 1 K 0 , (7.63)
2 r
i6=j

neglecting terms of order N 0 compared to N . The renormalized partition function is finally


"  2 #
y X 1  y 2N  dr
2N
0 0
Z = exp 2 r dr V 1 + (2 K)
(r0 )2 (N !)2 (r0 )2 r0
| {z } N
=: Zpair

Z Z
1 X dr

|ri rj |
d2 r1 d2 r2N exp 2K + 8 4 y 2 K 2 0 Ni Nj ln . (7.64)
2 r r0
D10 0
D2N i6=j

Here, Zpair is the partition function of the small pair we have integrated out. It is irrelevant for the renormalization
of y and K. Apart from this factor, Z 0 is identical to Z if we set
 
dr
y(r0 ) = 1 + (2 K(r)) 0 y(r) (7.65)
r
dr
K(r0 ) = K(r) 4 3 y 2 K 2 (r) 0 . (7.66)
r
Introducing the logarithmic length scale
r dr
l := ln dl = , (7.67)
r0 r
we obtain the Kosterlitz RG flow equations
dy
= (2 K) y, (7.68)
dl
dK
= 4 3 y 2 K 2 . (7.69)
dl
The initial conditions are

y(l = 0) = y0 = eEcore /kB T , (7.70)


1
K(l = 0) = K0 = 2 , (7.71)
kB T

i.e., the parameters assume their bare values at r = r0 (l = 0).

61
We will now discuss the physics encoded by the RG flow equations. First, note that the quantity
2
C := 2 2 y 2 ln K (7.72)
K
is invariant under the RG flow:
dC dy 2 dK 1 dK
= 4 2 y 2
= 4 2 y 2 (2 K) 8 2 y 2 + 4 3 y 2 K = 0. (7.73)
dl dl K dl K dl
Thus C is a first integral of the flow equations. We can calculate C from the initial values y0 , K0 and obtain
 
2 2 2 2 2 1 1 K
2 y = 2 y0 + + ln (7.74)
K K0 K0
s  
2 1 1 1 1 K
y = y0 + 3 + 2 ln . (7.75)
K K0 2 K0
The RG flow is along curves decribed by this expression, where the curves are specified by y0 , K0 . These para-
meters change with temperature as given in Eqs. (7.70) and (7.71). These initial conditions are sketched as a
dashed line in the figure. We see that there are two distinct cases:
y
T

T = Tc
rix
arat
sep

0 2/ K

For T < Tc , K flows to some finite value K(l ) > 2/. This means that even infinitely large pairs feel
a logarithmic attraction, i.e., are bound. Moreover, the fugacity y flows to zero, y(l ) = 0. Thus large
pairs are very rare, which is consistent with their (logarithmically) diverging energy.
For T > Tc , K flows to K(l ) = 0. Thus the interaction between a vortex and an antivortex that
are far apart is completely screened. Large pairs become unbound. Also, y diverges on large length scales,
which means that these unbound vortices proliferate. This divergence is an artifact of keeping only the
leading order in y in the derivation. It is cut off at finite y if we count vortex-antivortex pairs consistently.
But the limit K 0 remains valid.
At T = Tc we thus find a phase transition at which vortex-antivortex pairs unbind, forming free vortices. It is called
the Berezinskii-Kosterlitz-Thouless (BKT) transition. In two-dimensional films, vortex interactions thus suppress
c
the temperature where free vortices appear and quasi-long-range order is lost from the point T = Tsingle vortex
where
1 kB T 1 1 ! 1 2
= = = K0 = (7.76)
4 2 K0 4
to the one where K(l ) = 2/ and y0 , K0 lie on the separatrix between the two phases,
2 ! 2 2 2
C = 2 2 y02 ln K0 = 0 1 ln + ln K0 = 1 + ln + 2 2 y02 . (7.77)
K0 K0
Clearly the two criteria agree if y0 = 0. This makes sense since for y0 = 0 there are no vortex pairs to screen the
interaction. In addition, a third temperature scale is given by the mean-field transition temperature TMF , where
!
= 0.

62
quasilongrange free vortices
order no condensate
c
0 Tc Tsingle vortex TMF T
K( l oo ) = 2/ K 0 = 2/ = 0
In the low-temperature phase, the largest pairs determine the decay of the correlation function h (r)(0)i for
large r. Thus we find
 
2 r
h (r)(0)i = 0 (7.78)
r0
with
1 1
= . (7.79)
2 K(l )
We note that the exponent changes with temperature (one could say that the whole low-temperature phase is
critical) but assumes a universal value at Tc : There,
2 1
lim K = (Tc ) = . (7.80)
l 4
Due to the screening of the vortex interaction, the condensate is less stiff than it would be in the absence of
vortices. This is described by the renormalization of K from K0 to K(l ). It is customary but somewhat
misleading to express this as a renormalization of the superfluid density ns , which is after all proportional to K0 .
Assuming that the temperature dependence of and thus of the bare superfluid density

n0s 02 = K0 (7.81)

is negligible close to Tc , the renormalized superfluid density
K 0
ns (T ) := n (7.82)
K0 s
obtains its temperature dependence exclusively from K/K0 . (The argument l is implied here and in the
following.) For T < Tc but close to the BKT transition we have
2
K= + K. (7.83)

2
The invariant C = 2 2 y02 K0 ln K0 is an analytic function of temperature and its value at Tc is
 
2 2
Cc = 2 2 y 2 ln K = 1 ln . (7.84)
K T =Tc
Thus we can write, close to Tc ,
2
C
= 1 ln + b (T Tc ) (7.85)

with some constant b := dC/dT |T =Tc . On the other hand,
 
2
C = 2 2 y 2 ln K
K
T 
2 2
= ln + K
2 + K


1 2  
= ln ln 1 + K
1 + 2 K 2
 2 2  1 2
= 1 + K  K 2 ln K
 + K 2
2 4 
2 2 4
2 2
= 1 ln K 2 . (7.86)
8

63
Thus
2 2 p
K 2
= b (T Tc ) K
= 2b Tc T . (7.87)
8 | {z }
const

Consequently, ns = (K/K0 ) n0s jumps to a finite value at Tc and then increases like a square root.
ns

Tc T

This behavior was indeed measured in tortion-pendulum experiments on He-4 films (Bishop and Reppy, 1980).

7.2 Superconducting films


In this section we concentrate on what is different for charged superconductors compared to neutral superfluids.
Recall that for a superfluid the gradient term in the Landau functional reads
Z
d2 r () . (7.88)

If we move around a vortex once, the phase has to change by 2, regardless of the distance from the vortex core.
This phase change leads to an unavoidable contribution to the gradient term of
Z Z  
2 1 i 1 i
d r () = dr d r e e
r r
Z   2 Z Z
1 1 dr
= dr d = dr d = 2 , (7.89)
r r r

which diverges logarithmically with system size. On the other hand, for a superconducting film the gradient term
reads Z   2
2 1 ~ q
d r
A . (7.90)
2m i c
Again, the phase of winds by 2 around a vortex, but the associated gradient can, in principle, be compensated
by the vector potential. Pearl (1964) showed that this indeed leads to a finite energy of a single vortex in a
superconducting film. In addition, the free energy contains the magnetic-field term

B 2 (r)
Z
d3 r . (7.91)
8
Note that the integral is three-dimensional the field is present in all space, also outside of the film.
We first note that the film has a new effective length scale: In the London gauge, Ampres law reads
4 4
A = ( A}) 2 A = j 2 A = j. (7.92)
| {z c c
=0

On the other hand, from the London equation we have


c
j= A, (7.93)
42

64
which is valid within the film and away from vortex cores so that ||
2
= /. However, the current is confined
to the thin film. We can write
j(r) = K(x, y) (z) (7.94)
with a surface current density K. Thus
Z Z  c 
K(x, y) = dz j(r) = dz A(r). (7.95)
42
film

If the thickness is d and A is approximately constant across the thickness, we have


c d
K(x, y) = A(x, y, 0) (7.96)
4 2
c d
j(r) = A(r) (z) (7.97)
4 2
and finally
d
2 A =A (z). (7.98)
2
This result exhibits the new length scale that controls the spatial variation of A and thus of the current j,

2
:= , (7.99)
d
which is large,  , for a thin film. We see that in thin films, assumes the role of the penetration depth
. Since is large for thin films, one could say that thin films are always effectively of type II.
We here do not discuss the full derivation of Pearl but only consider the far field for large r  and show
that it does not lead to a diverging
p free energy for a single vortex. We assume that most of the magnetic flux of
0 penetrates the film for % = x2 + y 2 . . Then the magnetic field above (and below) the film looks like a
monopole field for r  .

z B

This field is easy to obtain from symmetry:


Z
da B = 0 2r2 B = 0 (7.100)
half space
for z>0
0 0
B(r) = and B(r) = r for z > 0. (7.101)
2r2 2r2
By symmetry,
0
B(r) = sgn z r. (7.102)
2r2

65
This gives a contribution to the field energy of
Z
B2 20 20 20 1
Z Z
3 3 1 dr
d r = d r 4 = 4 = , (7.103)
8 32 3 r 32 3 r2 8 2

which is finite for an infinity film. The value of the lower cutoff does not matter for this. The cutoff has to be
present, since the monopole field is not a valid approximation for small r.
We next obtain the current from Ampres law in integral form:
r
B
d

I Z
4
dr B = da j (7.104)
c
r+r
Z
0 4
2 dr0 0 2
= r d j(r) (7.105)
2(r ) c
r
 
4 0 1 1 0 r
r d j(r) = = (7.106)
c r r + r r2
0 1
j(r) = (7.107)
4 2 r2 d
and with the vector character restored
0
j(r) = . (7.108)
4 2 r2 d
Note that the sheet current is thus
0
K(r) = . (7.109)
4 2 r2
p
For large r we have || = 0 = / (note that typically  ). Then the second Ginzburg-Landau
equation gives
q~ q2 q  q 
j = i (i i)02 02 A = 02 ~ A . (7.110)
2m m c m c
Thus we can rewrite the gradient term in the free energy as
2
2
m
Z Z Z
1  q 2 2 0 (m )
d d2 r
2
0 ~ A = d d r 4 2
j j = d d2 r 2 2 j j. (7.111)
2m c 2m 0 q 2q 0

The contribution from large r is


Z
dm 2 1 m 20 m 20
Z
dr 1
d r 04 4 2 =
2
2 = , (7.112)
2q 2 02 16 r d 32 4 q 2 02 d r 3 8 3 q 2 02 d 2

which is also finite for an infinite film.


We conclude that the free energy of an isolated vortex is finite in a superconducting film. It is then plausible
and indeed true that the interaction energy of a vortex-antivortex pair does not diverge for large separations r but
saturates for r  . Since for the far field of a single vortex the magnetic-field energy, Eq. (7.103), dominates
over the energy due to the gradient term, we expect the large-r interaction to be dominated by the Coulomb-type

66
attraction of the magnetic monopoles in the upper and lower half spaces. This supposition is borne out by a
proper analysis. Consequently, for large r the interaction behaves like
 2
0 1
Vint = const (7.113)
2 r

(0 /2 is the monopole strength, according to Eq. (7.102)).

All this showes that, strictly speaking, there will be a non-zero concentration of free vortices at any temperature
T > 0. Thus there is no quasi-long-range order. However, the relevant length scale is = 2 /d, which can
be very large for thin films, even compared to the lateral size L of the sample. In this case the large-r limit is
experimentally irrelevant. But for vortex separations r  , the magnetic-field expulsion on the scale r is very
weak since is the effective penetration depth. Then the fact that the condensate is charged is irrelevant and
we obtain the same logarithmic interaction as for a neutral superfluid.
Thus for thin films of typical size we can use the previously discussed BKT theory. For superconducting
films we even have the advantage of an additional observable, namely the voltage for given current. We give a
hand-waving derivation of V (I). The idea is that a current exerts a Magnus force on a vortex, in the direction
perpendicular to the current. The force is opposite for vortices and antivortices and is thus able to break vortex-
antivortex pairs. As noted above, free vortices lead to dissipation. A vortex moving through the sample in the
orthogonal direction between source and drain contacts leads to a change of the phase difference by 2. We
will see in the chapter on Josephson effects why this corresponds to a non-zero voltage. Since free vortices act
independently, it is plausible to assume that the resistance is

R nv , (7.114)

where nv now denotes the concentration of free vortices. To find it, note that the total potential energy due to
vortex-antivortex attraction and Magnus force can be written as

V = Vint 2FMagnus r (7.115)

with
r
Vint = 2 kB T K ln . (7.116)
r0

67
V

Vint

rbarrier x

There is a finite barrier for vortex-antivortex unbinding at a separation rbarrier determined from V /r = 0. This
gives
2 kB T Kbarrier
rbarrier = , (7.117)
2FMagnus
where Kbarrier := K(rbarrier ). The barrier height is

E := V (rbarrier ) V (r0 )
= V (rbarrier )
rbarrier
= 2 kB T Kbarrier ln 2FMagnus rbarrier
r0
 
rbarrier
= 2 kB T Kbarrier ln 1 . (7.118)
r0
For small currents we have rbarrier  r0 and thus
rbarrier
E
= 2 kB T K(l ) ln . (7.119)
| {z } r0
K

The rate at which free vortices are generated is


 2K
rbarrier
Rgen eE
= . (7.120)
r0
The recombination rate of two vortices to form a pair is

Rrec n2v , (7.121)

since two vortices must meet. In the stationary state we have


 K
p rbarrier
Rgen = Rrec nv Rgen (7.122)
r0
and thus a resistance of  K
rbarrier
R nv . (7.123)
r0
Since the Magnus force is FMagnus I we have
1 1
rbarrier (7.124)
FMagnus I

so that  K
1
R = I K . (7.125)
I

68
Finally, the voltage measured for a current I is

V = RI I K I = I 1+K , (7.126)

where K K(l ) is the renormalized stiffness. Since


2 2 p 
K= + K
= 1 + 2b Tc T (7.127)

we find for the exponent p
1 + K
= 3 + 2| {z2b} Tc T (7.128)
= const

for T . Tc .

1+ K
3

1
ohmic

Tc T

Above Tc we have K = 0 and thus ohmic resistance, V I, as expected. Below Tc , the voltage is sub-ohmic, i.e.,
the voltage is finite for finite current but rises more slowly than linearly for small currents. This behavior has
been observed for thin superconducting films.

69
8

Origin of attractive interaction

In the following chapters we turn to the microsopic theory of superconductivity. While the BCS theory is
reasonably easy to understand if one assumes an attractive interaction between electrons in a superconductor,
it is far from obvious where such an interaction should come from. The only fundamental interaction that is
relevant for (non-radioactive) solids is the electromagnetic one, which naively gives a repulsive interaction of a
typical strength of several eV between nearby electrons. How can this lead to an attraction at the low energy
scale kB Tc 1 meV? We will see that the lattice of ion cores (nuclei with tightly bound inner electrons) plays an
important role. We will use Feynman diagrams for Green functions to describe the physics. Unfortunately, we do
not have the time to introduce these concepts rigorously; this is done in many good textbooks on many-particle
physics as well as in the lecture notes on Vielteilchentheorie (in German), which are available online. For those
familiar with Feynman diagrams, they rigorously represent mathematical expressions, for the others they should
at least be useful as cartoons of the relevant processes.

8.1 Reminder on Green functions


Nevertheless, we start by briefly summarizing some properties of Green functions. In many-particle physics,
we usefully express the Hamiltonian in terms of an electronic field operator (r, t), where = , is the
electron spin. The field operator can be expanded in any convenient basis of single-particle states characterized
by wavefunctions (r), where represents all relevant quantum numbers,
X
(r, t) = (r) c (t), (8.1)

X
(r, t) = (r) c (t). (8.2)

c and c are annihilation and creation operators of electrons in the single-particle state, respectively. Note
that we are working in the Heisenberg picture, i.e., the wavefunctions (r) are independent of time, whereas
the time dependence is carried by the operators.
For example, the Hamiltonian for free electrons reads

~2 2
Z X
H = d3 r (r) (r). (8.3)

2m

70
In this case it is useful to expand into plane waves,
1 X ikr
(r) = e ck (8.4)
V k
~2 (k 0 )2
Z
1 X 0
eikr+ik r ck
X
H= d3 r ck0
V 0
2m
kk
X ~2 k 2
= ck ck . (8.5)
2m
k

Two types of Green functions are defined as follows:

greater Green function: D E


G> (rt, r0 0 t0 ) := i (r, t)0 (r0 , t0 ) , (8.6)

where h. . .i is the equilibrium average

Tr AeH 1
hAi = Tr Aeq = H
Tr AeH , (8.7)
Tr e Z
where = 1/kB T is the inverse temperature and H is the many-particle Hamiltonian. For non-interacting
electrons at temperature T  EF /kB , eq describes the Fermi sea. G> describes the conditional probability
amplitude for an electron created at point r0 with spin 0 at time t0 to be found at point r with spin at
time t.
lesser Green function: D E
G< (rt, r0 0 t0 ) := +i 0 (r0 , t0 ) (r, t) ; (8.8)

it describes the propagation of a hole from time t to time t0 .

These are not the most useful definitions since in quantum theory propagation of an electron forward in time
cannot be separated from propagation of a hole backward in time. It is also useful to distiguish between the cases
t > t0 and t < t0 . This is accomplished by these definitions:
retarded Green function:
Dn oE
GR (rt, r0 0 t0 ) := i (t t0 ) (r, t), 0 (r0 , t0 ) , (8.9)

where {A, B} := AB + BA is the anti-commutator (appropriate for fermions) and the step function
(
0 1 for t > t0
(t t ) = (8.10)
0 for t < t0

selects only contributions with t > t0 (forward in time).


advanced Green function:
Dn oE
GA (rt, r0 0 t0 ) := +i (t0 t) (r, t), 0 (r0 , t0 ) , (8.11)

this Green function analogously contains only contributions backward in time.


The thermal averages in the Green functions introduce operators eH , whereas the time evolution of operators
introduces time-evolution operators according to

A(t) = eiHt/~ A eiHt/~ . (8.12)

71
Specifically, we have
1  0 0
GR (rt, r0 0 t0 ) =i (t t0 ) Tr eH eiHt/~ (r) eiH(tt )/~ 0 (r0 ) eiHt /~
Z 
0 0
+ eiHt /~ 0 (r0 ) eiH(t t)/~ (r) eiHt/~ . (8.13)

In practice, we will want to write the Hamiltonian in the form H = H0 + V and treat V in perturbation theory.
It is not surprising that this is complicated due to the presence of H in several exponential factors. However,
these factors are of a similar form, only in some the prefactor of H is imaginary and in one it is the real inverse
temperature. Can one simplify calculations by making all prefactors real? This is indeed possible by formally
replacing t i , t0 i 0 , which is the main idea behind the imaginary-time formalism. We cannot discuss
it here but only state a few relevant results.
It turns out to be useful to consider the Matsubara (or thermal) Green function
D E
G(rt, r0 0 t0 ) := T (r, t)0 (r0 , t0 ) , (8.14)

where for any operator


A( ) := eH /~ A eH /~ . (8.15)
and T is a time-ordering directive:
(
0 A( )B( 0 ) for > 0
T A( )B( ) = (8.16)
B( )A( ) for < 0
0

(upper/lower sign for bosonic/fermionic operators). For time-independent Hamiltonians, the Green function only
depends on the difference 0 . One can then show that the resulting Green function G(r, r0 0 , ) is defined
only for [~, ~] and satisfies
G(r, r0 0 , + ) = G(r, r0 0 , ) (8.17)
for fermions. This implies that the Fourier transform is a discrete sum over the fermionic Matsubara frequencies
(2n + 1)
n := , n Z. (8.18)
~
The imaginary-time formalism is useful mainly because G is easier to obtain or approximate than the other Green
functions and these can be calculated from G based on the following theorem: The retarded Green function GR ()
in Fourier space is obtained from G(in ) by means of replacing in by + i0+ , where i0+ is an infinitesimal
positive imaginary part,
GR () = G(in + i0+ ). (8.19)
This is called analytic continuation. Analogously,
GA () = G(in i0+ ). (8.20)

G(i n)
in + i 0+

R
G ( )
A
G ( )

in i 0+

72
It will be useful to write the electronic Green function in k space. In the cases we are interested in, momentum
and also spin are conserved so that the Green function can be written as
D E
Gk ( ) = T ck ( ) ck (0) . (8.21)

Analogously, the bosonic Matsubara Green function of phonons can be written as


D E
Dq ( ) = T bq ( ) bq (0) , (8.22)

where enumerates the three polarizations of acoustic phonons. We also note that for bosons we have

D( + ) = +D( ), (8.23)

i.e., the opposite sign compared to fermions. Therefore, in the Fourier transform only the bosonic Matsubara
frequencies
2n
n := , nZ (8.24)
~
occur.

8.2 Coulomb interaction


We now discuss the effect of the electron-electron Coulomb interaction. We will mainly do that at the level of
Feynman diagrams but it should be kept in mind that these represent mathematical expressions that can be
evaluated if needed. The Coulomb interaction can be written in second-quantized form as
Z
1 X
Vint = d3 r1 d3 r2 1 (r1 )2 (r2 ) VC (|r2 r1 |) 2 (r2 )1 (r1 ) (8.25)
2
1 2

with
e2
VC (r) = . (8.26)
r
In momentum space we have
1 X ikr
(r) = e ck (8.27)
V k
so that
1 XZ 0 00
r2 +ik000 r1
ck1 ck0 2
X
Vint = d3 r1 d3 r2 eikr1 ik r2 +ik VC (|r2 r1 |) ck00 2 ck000 1
2V 2
kk0 k00 k000 1 2
1 XZ 0 00
+k000 )R i(kk0 +k00 k000 )/2
ck1 ck0 2
X
= d3 R d3 % ei(kk +k e VC () ck00 2 ck000 1 , (8.28)
2V 2
kk0 k00 k000 1 2

where
r1 + r2
R= , (8.29)
2
= r2 r1 . (8.30)

We can now perform the integral over R:


1 X XZ 0 00 000
Vint = d3 k+k0 k00 k000 ,0 ei(kk +k k )/2 ck1 ck0 2 VC () ck00 2 ck000 1
2V 0 00 000
kk k k 1 2
Z
1 X X 0 00
= d3 ei(k k ) VC () ck1 ck0 2 ck00 2 ck+k0 k00 ,1 . (8.31)
2V 0 00
kk k 1 2

73
Substituting new momentum variables

k1 = k + k0 k00 , (8.32)
00
k2 = k , (8.33)
q = k0 k00 , (8.34)

we obtain
1 X X
Vint = VC (q) ck1 q,1 ck2 +q,2 ck2 2 ck1 1 (8.35)
2V
k1 k2 q 1 2

with Z
VC (q) = d3 eiq VC (). (8.36)

The interaction Vint is a sum over all processes in which two electrons come in with momenta k1 and k2 , a
momentum of q is transfered from one to the other through the Coulomb interaction, and the electrons fly out
with momenta k1 q and k2 + q.

k1 q,1 k2 + q,2
Vc (q)

k11 k 2 2

VC (q) is obviously the Fourier transform of the Coulomb interaction. It is most easily obtained by Fourier
transforming the Poisson equation for a point charge,

2 (r) = 4 (r) = 4Q (r) (8.37)


Z
d3 r eiqr 2 (r) = 4Q (8.38)
Z
by parts
d3 r (iq)2 eiqr (r) = 4Q (8.39)

q 2 (q) = 4Q (8.40)
Q
(q) = 4 2 (8.41)
q
so that
e2
VC (q) = 4 . (8.42)
q2

Screening and RPA


The bare Coulomb interaction VC is strongly repulsive, as noted above. But if one (indirectly) measures the
interaction between charges in a metal, one does not find VC but a reduced interaction. First of all, there will be
a dielectric function from the polarizability of the ion cores,
4 e2
VC (q) , (8.43)
 q2
but this only leads to a quantitative change, not a qualitative one. We absorb the factor 1/ into e2 from now
on. More importantly, a test charge in a metal is screened by a cloud of opposite charge so that from far away
the effective charge is strongly reduced. Diagramatically, the effective Coulomb interaction VCfull (q, in ) is given
by the sum of all connected diagrams with two external legs that represent the Coulomb interaction. With the
representations

VC (8.44)

74
(the minus sign is conventional) and

VCfull (8.45)

as well as

G0 (8.46)
for the bare electronic Green function one would find for the non-interacting Hamiltonian H0 , we obtain
=

+ + +

+ + + +

+ (8.47)
This is an expansion in powers of e2 since VC contributes a factor of e2 . We have exhibited all diagrams up to
order e6 . If we try to evaluate this sum term by term, we accounter a problem: At each vertex , momentum and
energy (frequency) must be conserved. Thus in partial diagrams of the form

k k

q=0

the Coulomb interaction carries momentum q = 0. But


e2
VC (0) = 4 (8.48)
02
is infinite. However, one can show that the closed G loop corresponds to the average electron density so that the
diagram signifies the Coulomb interaction with the average electronic charge density (i.e., the Hartree energy).
But this is compensated by the average charge density of the nuclei. Thus we can omit all diagrams containing
the tadpole diagram shown above.
Since we still cannot evaluate the sum in closed form we need an approximation. We first consider the limiting
cases of small and large q in VCfull (q, in ).
For large q, corresponding to small distances, the first diagram is proportional to 1/q 2 , whereas all the
others are at least of order 1/q 4 and are thus suppressed. We should recover the bare Coulomb interaction
for large q or small distances, which is plausible since the polarization of the electron gas cannot efficiently
screen the interaction between two test charges that are close together.
For small q we find that higher-order terms contain higher and higher powers of 1/q 2 and thus become
very large. This is alarming. The central idea of our approximation is to keep only the dominant term
(diagram) at each order in e2 . The dominant term is the one with the highest power in 1/q 2 . Only the
VC lines forming the backbone of the diagrams (drawn horizontally) carry the external momentum q due
to momentum conservation at the vertices. Thus the dominant terms are the ones with all VC lines in the
backbone:
q k q k q

k+q k+ q

75
Summing up these dominant terms we obtain the approximate effective interaction

VCRPA := +

+ +

+ (8.49)

This is called the random phase approximation (RPA) for historical reasons that do not concern us here, or the
Thomas-Fermi approximation. The most important part of RPA diagrams is clearly the bubble diagram

0 , (8.50)

which stands for


1X 1 X 0 0
0 (q, in ) = Gk+q, (in + in ) Gk (in )
i V
n k
2 XX 1 1
= (8.51)
in
in + in k+q in k
k

(the factor of 2 is due to the spin). We can write the diagramatic series also as

VCRPA (q, in ) = VC (q) + VC (q) 0 (q, in ) VC (q) VC (q) 0 (q, in ) VC (q) 0 (q, in ) VC (q) +
 
= VC (q) 1 0 (q, in ) VC (q) + 0 (q, in ) VC (q) 0 (q, in ) VC (q) . (8.52)

This is a geometric series, which we can sum up,

VC (q)
VCRPA (q, in ) = . (8.53)
1 + VC (q) 0 (q, in )

0 can be evaluated, it is essentially 1 times the susceptibility of the free electron gas. We cannot explain this
here, but it is plausible that the susceptibility, which controls the electric polarization of the electron gas, should
enter into a calculation of the screened Coulomb interaction. Note that VCRPA has a frequency dependence since
0 (or the susceptibility) has one.
In the static limit i 0 + i0+ and at low temperatures T  EF /kB , one can show that

0 (q, 0)
= const = N (EF ), (8.54)

where N (EF ) is the electronic density of states at the Fermi energy, including a factor of two for the spin. Thus
we obtain
2
4 qe2 e2
VCRPA (q) = 2 = 4
1 + 4 qe2 0 q 2 + 4e2 0

e2
= 4
q 2 + 4e2 N (EF )
e2
= 4 2 (8.55)
q + 2s

with p
s := 4e2 N (EF ). (8.56)

76
Note that summing up the more and more strongly diverging terms has led to a regular result in the limit q 0.
The result can be Fourier-transformed to give
es r
VCRPA (r) = e2 (8.57)
r
(not too much confusion should result from using the same symbol e for the elementary charge and the base of
the exponential function). This is the Yukawa potential, which is exponentially suppressed beyond the screening
length 1/s . This length is on the order of 108 m = 10 nm in typical metals. While we thus find a strong
suppression of the repulsive interaction at large distances, there is no sign of it becoming attractive.
On the other hand, for very high frequencies & EF /~ the electron gas cannot follow the perturbation, the
susceptibility and 0 go to zero, and we obtain the bare interaction,
e2
VCRPA (q, )
= VC (q) = 4 2 . (8.58)
q

8.3 Electron-phonon interaction


The nuclei (or ion cores) in a crystal oscillate about their equilibrium positions. The quanta of these lattice
vibrations are the phonons. The many-particle Hamiltonian including the phonons has the form
H = Hel + Hph + Hel-ph , (8.59)
where we have discussed the electronic part Hel before,
 
1
q bq bq +
X
Hph = (8.60)
2
q

is the bare Hamiltonian of phonons with dispersion q , and


1 XX  
Hel-ph = gq ck+q, ck bq + bq, (8.61)
V
k q

describes the electron-phonon coupling. gq is the coupling strength. Physically, an electron can absorb (bq )
or emit (bq, ) a phonon under conservation of momentum. Hence, electrons can interact with one another by
exchanging phonons. Diagrammatically, we draw the simplest possible process as

k1 q,1 0 k2+ q,2


Dq

k11 k 22

In detail, we define, in analogy to the Coulomb interaction


VC (q) , (8.62)

the interaction due to phonon exchange,


1 2 0
Vph (q, , in )
|gq | Dq (in ) . (8.63)
V
We quote the expression for the bare phonon Green function, i.e., the one obtained from Hph alone:
0 1 1 2 q
Dq (in ) = + = . (8.64)
in q in q (in )2 2q
The phonon-mediated interaction is thus frequency-dependent, whereas the Coulomb interaction is static. How-
ever, we could write the Coulomb interaction in a very similar form as the exchange of photons. Since the speed
of light is so much larger than the speed of sound, we can neglect the dynamics for photon exchange but not for
phonon exchange.

77
Jellium phonons
For our discussions we need a specific model for phonons. We use the simplest one, based on the jellium approx-
imation for the nuclei (or ion cores). In this approximation we describe the nuclei by a smooth positive charge
density + (r, t). In equilibrium, this charge density is uniform, + (r, t) = 0+ . We consider small deviations

+ (r, t) = 0+ + + (r, t). (8.65)

Gauss law reads


E = 4 + (r, t), (8.66)
since 0+ is compensated by the average electronic charge density. The density of force acting on + is f = + E
=
0+ E to leading order in + . Thus
f = 4 0+ + . (8.67)
The conservation of charge is expressed by the continuity equation

+ + + v = 0. (8.68)
t |{z}
= j+

To leading order this reads



+ + 0+ v
=0 (8.69)
t
2 Newton f
+
= 0+ v = 0+ , (8.70)
t2 t m
where m is the mass density of the nuclei. With the nuclear charge Ze and mass M we obtain

2 Ze Ze
2
+
= f = 4 0+ + , (8.71)
t M M
which is solved by
+ (r, t) = + (r) eit (8.72)
with r r
Ze 0 Z 2 e2 0
= 4 = 4 n+ , (8.73)
M + M
where n0+ is the concentration of nuclei (or ions). We thus obtain optical phonons with completely flat dispersion,
i.e., we find the same frequency for all vibrations.
One can also obtain the coupling strength gq . It is clear that it will be controlled by the Coulomb interaction
between electrons and fluctuations + in the jellium charge density. We refer to the lecture notes on many-particle
theory and only give the result:
1 2
|gq | = VC (q). (8.74)
V 2
Consequently, the electron-electron interaction due to phonon exchange becomes

1 2 2 2
Vph (q, in ) = |gq | Dq0 (in ) = VC (q) = V C (q) . (8.75)
V 2 (in )2 2 (in )2 2

It is thus proportional to the bare Coulomb interaction, with an additional frequency-dependent factor. The
retarded form reads

R 2 2
Vph (q, ) = Vph (q, in + i0+ ) = VC (q) = V C (q) , (8.76)
( + i0+ )2 2 2 2 + i0+ sgn

where we have used 2 i0+ = i0+ sgn and have neglected the square of infinitesimal quantities.

78
8.4 Effective interaction between electrons
Combining the bare Coulomb interaction and the bare interaction due to phonon exchange, calculated for the
jellium model, we obtain the bare effective interaction between electrons,
Veff (q, in ) := VC (q) + Vph (q, in )
2
= VC (q) + VC (q)
(in )2 2
(in )2
= VC (q) . (8.77)
(in )2 2
The retarded form is
R
Veff (q, ) = Veff (q, in + i0+ )
2
= VC (q) . (8.78)
2 2 + i0+ sgn
R
This expression is real except at = and has a pole there. Moreover, Veff is proportional to VC with a negative
prefactor as long as < .
VeffR

Vc(q )

The effective interaction is thus attractive for < . The exchange of phonons overcompensates the repulsive
Coulomb interaction. On the other hand, for 0, the effective interaction vanishes. This means that in a
quasi-static situation the electrons do not see each other at all.
What happens physically is that the electrons polarize the (jellium) charge density of the nuclei. The nuclei
have a high inertial mass, their reaction to a perturbation has a typical time scale of 1/ or a frequency scale of
. For processes slow compared to , the nuclei can completely screen the electron charge, forming a polaron,
which is charge-neutral. For frequencies > 0, we have to think in terms of the response of the system to
a test electron oscillating with frequency . The jellium acts as an oscillator with eigenfrequency . At the
present level of approximation it is an undamped oscillator. The jellium oscillator is excited at the frequency
. For 0 < < , it is driven below its eigenfrequency and thus oscillates in phase with the test electron. The
amplitude, i.e., the jellium polarization, is enhanced compared to the = 0 limit simply because the system is
closer to the resonance at . Therefore, the oscillating electron charge is overscreened. On the other hand, for
> the jellium oscillator is driven above its eigenfrequency and thus follows the test electron with a phase
difference of . Thus the electron charge is not screened at all but rather enhanced and the interaction is more
strongly repulsive than the pure Coulomb interaction.

Screening of the effective interaction


From our discussion of the Coulomb interaction we know that the real interaction between two electrons in a
metal is strongly screened at all except very short distances. This screening is well described within the RPA.
We now apply the RPA to the effective interaction derived above. We define
Veff = + (8.79)

79
and
RPA
Veff := + + + (8.80)

or
RPA
Veff (q, in ) = Veff (q, in ) + Veff (q, in ) 0 (q, in ) Veff (q, in )
Veff (q, in ) 0 (q, in ) Veff (q, in ) 0 (q, in ) Veff (q, in ) + (8.81)

As above, we can sum this up,


(in )2
RPA Veff (q, in ) (in )2 2
Veff (q, in ) = = VC (q) (in )2
1 + Veff (q, in )0 (q, in ) 1 + VC (q) 0 (q, in )
(in )2 2
(in )2
= VC (q)
(in )2 2
+ (in )2 VC (q)0 (q, in )
VC (q) (in )2 + (in )2 VC (q)0 (q, in )
=
1 + VC (q)0 (q, in ) (in )2 2 + (in )2 VC (q)0 (q, in )
| {z }
= VCRPA (q,in )

(in )2
= VCRPA (q, in ) 2
(in )2 1+VC (q)0 (q,in )
(in )2
= VCRPA (q, in ) (8.82)
(in )2 q2 (in )

with the renormalized phonon frequency



q (in ) := p . (8.83)
1 + VC (q)0 (q, in )
RPA
To see that this is a reasonable terminology, compare Veff to the bare effective interaction

(in )2
Veff (q, in ) = VC (q) . (8.84)
(in )2 2

Evidently, screening leads to the replacements VC VCRPA and q .


For small momenta and frequencies, we have 0 N (EF ), the density of states at EF . In this limit we thus
obtain

q
=q =q
=q 2 = q. (8.85)
2
1 + 4 qe2 N (EF ) 2
1 + q2s
s s
q 2

Due to screening we thus find an acoustic dispersion of jellium phonons. This is of course much more realistic
than an optical Einstein mode.
Beyond the low-frequency limit it is important that 0 and thus q obtains a sizable imaginary part. It smears
RPA
out the pole in the retarded interaction Veff (q, ) or rather moves it away from the real-frequency axisthe
lattice vibrations are now damped. The real part of the retarded interaction is sketched here for fixed q:

80
Re VeffRPA, R

Vc RPA (q )

0
Re q

Note that
the interaction still vanishes in the static limit 0,
the interaction is attractive for 0 < < Re q , where Re q q.
It is important that the static interaction is not attractive but zero. Hence, we do not expect static bound states
of two electrons.
To obtain analytical results, it is necessary to simplify the interaction. The main property required for
superconductivity is that the interaction is attractive for frequencies below some typical phonon frequency. The
typical phonon frequency is the material specific Debye frequency D . We write the effective RPA interaction in
terms of the incoming and transferred momenta and frequencies,
RPA
Veff RPA
= Veff (k, in ; k0 , in0 ; q, in ). (8.86)

k q, i n i n q, i n k+ q , i n + i n

k, i n k, i n

We then approximate the interaction (very crudely) by a constant V0 < 0 if both incoming frequencies are
smaller than D and by zero otherwise,
(
RPA V0 for |in | , |in0 | < D ,
Veff (8.87)
0 otherwise.

81
9

Cooper instability and BCS ground state

In this chapter we will first show that the attractive effective interaction leads to an instability of the normal
state, i.e., of the Fermi sea. Then we will dicuss the new state that takes its place.

9.1 Cooper instability


Let us consider the scattering of two electrons due to the effective interaction. A single scattering event is
represented by the diagram

k, i n k+ q , i n + i n

k, i n k q, i n i n

Electrons can also scatter multiple times:

+ + + + (9.1)

An instability occurs if this series diverges since then the scattering becomes infinitely strong. In this case the
perturbative expansion in the interaction strength V0 represented by the diagrams breaks down. This means that
the true equilibrium state cannot be obtained from the equilibrium state for V0 = 0, namely the noninteracting
Fermi gas, by perturbation theory. A state that is perturbatively connected to the free Fermi gas is called a
Landau Fermi liquid. It is an appropriate description for normal metals. Conversely, a scattering instability
signals that the equilibrium state is no longer a Fermi liquid.
Like in the RPA, it turns out to be sufficient to consider the dominant diagrams at each order. These are
the ladder diagrams, which do not contain crossing interaction lines. Moreover, the instability occurs first for
the scattering of two electrons with opposite momentum, frequency, and spin. We thus restrict ourselves to the

82
diagrams describing this situation. We define the scattering vertex by

k in p i n

k in p i n
k in p i n k in k1 i 1n p i n

k k1 , k1 p,
:= k p, i n i n + + (9.2)
i n i 1n i 1n i n

k in p i n k in k1 i 1n p i n

This is a geometric series, which we can sum up,

= 1 + + + = (9.3)

With our approximation (


RPA V0 for |in | < D ,
Veff (9.4)
0 otherwise,
we obtain
+V0
(in ) 1 1 (9.5)
Gk01 (in1 ) Gk
0
P P
1 V0 1 , |i 1 | <
in n D V k1 1
(in1 )
for |in | < D and zero otherwise. Thus

V0
for |in | < D ,


1 1 0 0
P P 1 1
(in ) 1 V0 in1 , |in1 | < D V k1 Gk1 (in ) Gk1 (in ) (9.6)

0 otherwise.

We see that the scattering vertex diverges if


1 X 1 X 0
V0 Gk1 (in1 ) Gk
0
1
(in1 ) = 1. (9.7)
1
V
in k1
|in1 | < D

This expression depends on temperature. We now evaluate it explicitly:


Z
X 1 X 1 1 X 1
V0 kB T = V0 kB T d D( + ) (9.8)
1
V in1 k1 in1 k1 1
(n1 )2 + 2
in k1 in
| 1
in |
< D | 1
in| < D

with the density of states per spin direction and per unit cell, D(). Assuming the density of states to be

83
approximately constant close to the Fermi energy we get (with ~ = 1)
Z
X d
V0 kB T D(EF )
1
(n1 )2 + 2
in
|in1 | < D | {z
1|
}
= /|n
D /2 D /2
X 1 X 1
= V0 kB T D(EF ) 2 (2n+1)
= V0 
kB
T D(EF )

1
n=0

n=0
n+ 2

  
D
= V0 D(EF ) + ln 4 . (9.9)
2

In the last step we have used an approximation for the sum over n that is valid for D  1, i.e., if the sum has
many terms. Since Tc for superconductors is typically small compared to the Debye temperature D /kB (a few
hundred Kelvin), this is justified. 0.577216 is the Euler constant. Altogether, we find

V0
  (9.10)
1 V0 D(EF ) + ln 2

D

for |in | < D . Coming from high temperatures, but still satisfying kB T  D , multiple scattering enhances .
diverges at T = Tc , where
 
2D
V0 D(EF ) + ln =1 (9.11)
kB Tc
2e D 1
ln = (9.12)
kB Tc V0 D(EF )
2e
 
1
kB Tc = D exp . (9.13)

|{z} V0 D(EF )
1.13387 1

This is the Cooper instability. Its characteristic temperature scale appears to be the Debye temperature of a few
hundred Kelvin. This is disturbing since we do not observe an instability at such high temperatures. However,
the exponential factor tends to be on the order of 1/100 so that we obtain Tc of a few Kelvin. It is important
that kB Tc is not analytic in V0 at V0 = 0 (the function has an essential singularity there). Thus kB Tc cannot
be expanded into a Taylor series around the non-interacting limit. This means that we cannot obtain kB Tc in
perturbation theory in V0 to any finite order. BCS theory is indeed non-perturbative.

9.2 The BCS ground state


We have seen that the Fermi sea becomes unstable due to the scattering of electrons in states |k, i and |k, i.
Bardeen, Cooper, and Schrieffer (BCS) have proposed an ansatz for the new ground state. It is based on the
idea that electrons from the states |k, i and |k, i form (so-called Cooper) pairs and that the ground state is
a superposition of states built up of such pairs. The ansatz reads
Y 
|BCS i = uk + vk ck ck, |0i , (9.14)
k

84
where |0i is the vacuum state without any electrons and uk , vk are as yet unknown complex coefficients. Normal-
ization requires

1 = hBCS |BCS i
Y 
uk0 + vk0 ck0 ck0 , |0i
Y
= h0| (uk + vk ck, ck )
k k0
Y 
uk vk ck ck,
2 2
= h0| |uk | + + uk vk ck, ck + |vk | |0i
k
Y 2 2

= |uk | + |vk | . (9.15)
k

2 2
This is certainly satisfied if we demand |uk | + |vk | = 1 for all k, which we will do from now on. Note that the
occupations of |k, i and |k, i are maximally correlated; either both are occupied or both are empty. Also,
|BCS i is peculiar in that it is a superposition of states with different total electron numbers. (We could imagine
the superconductor to be entangled with a much larger electron reservoir so that the total electron number in
superconductor and reservoir is fixed.) As a consequence, expressions containing unequal numbers of electronic
creation and annihilation operators can have non-vanishing expectation values. For example,
D E D E
ck ck, BCS ck ck, BCS

BCS
Y 
(uk0 + vk 0 ck0 , ck0 ) ck ck, uk00 + vk00 ck00 ck00 , |0i
Y
= h0|
k0 k00
Y  2 2

= h0| vk uk |uk0 | + |vk0 | |0i = vk uk . (9.16)
k0 6=k | {z }
=1

The coefficients uk , vk are chosen so as to minimize the expectation value of the energy, hBCS |H| BCS i, under
2 2
the constraint |uk | + |vk | = 1 for all k. |BCS i is thus a variational ansatz.
We write the Hamiltonian as
k ck ck + Vint
X
H= (9.17)
k

and, in the spirit of the previous section, choose the simplest non-trivial approximation for Vint that takes into
account that
1. only electrons with energies |k | . D relative to the Fermi energy are important and
2. the instability is due to the scattering between electrons in single-particle states |k i and |k, i.
This leads to
1 X
Vint = Vkk0 ck ck, ck0 , ck0 (9.18)
N 0
kk

with (
V0 for |k | < D and |k0 | < D ,
Vkk0 = (9.19)
0 otherwise.
We assume that scattering without momentum transfer, k0 = k, contributes negligibly compared to k0 6= k since
there are many more scattering channels for k0 6= k. We also assume that uk = uk , vk = vk , which is, at worst,

85
a restriction of our variational ansatz. Then we obtain
Y 
uq + vq cq, cq ck ck uq0 + vq0 cq0 cq0 , |0i
X Y 
hBCS |H| BCS i = k h0|
k q q0
1 X Y 
uq + vq cq, cq ck ck, ck0 ck0 uq0 + vq0 cq0 cq0 , |0i
Y 
+ Vkk0 h0|
N 0
kk q q0

k h0| |vk | ck ck, ck ck ck, ck |0i
X 2
X 2
= k h0| |vk | ck, ck ck ck ck ck, |0i +
k k
1 X
+ Vkk0 h0| vk uk0 uk vk0 ck, ck ck ck, ck0 , ck0 ck0 ck0 , |0i
N 0
kk
X 2 1 X
= 2k |vk | + Vkk0 vk uk uk0 vk0 =: EBCS . (9.20)
N 0
k kk

This energy should be minimized with respect to the uk , vk . For EBCS to be real, the phases of uk and vk must
be the same. But since EBCS is invariant under

uk uk eik , vk vk eik , (9.21)

we can choose all uk , vk real. The constraint from normalization then reads u2k + vk2 = 1 and we can parametrize
the coefficients by
uk = cos k , vk = sin k . (9.22)
Then
X 1 X
EBCS = 2k sin2 k + Vkk0 sin k cos k sin k0 cos k0
N 0
k kk
X 1 X Vkk0
= k (1 cos 2k ) + sin 2k sin 2k0 . (9.23)
N 0 4
k kk

We obtain the minimum from


EBCS 1 X Vqk0 1 X Vkq
= 2q sin 2q + cos 2q sin 2k0 + sin 2k cos 2q
q N 0 2 N 2
k k
1 X !
= 2q sin 2q + Vqk0 cos 2q sin 2k0 = 0. (9.24)
N 0
k

We replace q by k and parametrize k by


k
sin 2k =: p (9.25)
k2 + 2k

and write
k
cos 2k = p . (9.26)
k2+ 2k
The last equality is only determined by the previous one up to the sign. We could convince ourselves that the
other possible choice does not lead to a lower EBCS . Equation (9.24) now becomes

k k 1 X k k0
2p 2 2
+ Vkk0 p 2 p =0 (9.27)
k + k N 0 k + 2k k2 0 + 2k0
k
1 X k 0
k = Vkk0 p 2 . (9.28)
N 0 2 k0 + 2k0
k

86
This is called the BCS gap equation for reasons to be discussed below. When we have solved it, it is easy to
obtain the original variational parameters in terms of k ,
!
2 1 k
uk = 1+ p 2 , (9.29)
2 k + 2k
!
2 1 k
vk = 1 p 2 , (9.30)
2 k + 2k
k
u k vk = p 2 . (9.31)
2 k + 2k

The relative sign of uk and vk is thus the sign of k . The absolute sign of, say, uk is irrelevant because of the
invariance of EBCS under simultaneous phase rotations of uk , vk (consider a phase factor of ei = 1).
For our special interaction (
V0 for |k | , |k0 | < D ,
Vkk0 = (9.32)
0 otherwise,
the BCS gap equation becomes
1 k0
X

(V0 ) p 2 for |k | < D ,
k = N 2 k0 + 2k0 (9.33)
k0 , |k0 | < D

0 otherwise,

which can be solved with the ansatz


(
0 > 0 for |k | < D ,
k = (9.34)
0 otherwise.

We obtain
V0
p 0
X
0 = (9.35)
N 2 k2 0 + 20
k0
|k0 | < D
ZD
V0 X 1 1
1= p = V0 d D( + ) p , (9.36)
N 2 k2 0 + 20 2 + 20
2
k0 D
|k0 | < D

where D() is the density of states per spin direction and per unit cell. If the density of states is approximately
constant within D of the Fermi energy, we obtain
ZD
1 d D
1 = V0 D(EF ) p = V0 D(EF ) Arsinh (9.37)
2 2
2 + 0 0
D
1 D
sinh = (9.38)
V0 D(EF ) 0
1
0 = D 1 . (9.39)
sinh V0 D(E F)

In the so-called weak-coupling limit of small V0 D(EF ), this result simplifies to


 
1
0 = 2D exp . (9.40)
V0 D(EF )

87
Interestingly, apart from a numerical factor, the value of 0 agrees with kB Tc for the Cooper instability. We will
return to this observation below. The non-analyticity of the function 0 (V0 ) means that we cannot obtain 0
and thus |BCS i within perturbation theory in V0 .
We can now find the energy gain due to the superconducting state, i.e., the condensation energy. For this, we
insert uk , vk into EBCS ,
!
X 1 k 1 X k k0
EBCS = 2
k 1 p 2 2
+ Vkk0 p 2 p . (9.41)
2 + N 4 + 2k k2 0 + 2k0
k  k k kk0 k

We use the simple form of Vkk0 and assume that D() is constant within D of the Fermi energy but not outside
of this interval. This gives

Z D ZD !

EBCS = N d D( + ) 2 + N d D( + ) 1 p
2 + 20
D

Z 
 ZD ZD
20
d 0 D( + )D( + 0 )(V0 ) p

+N dD(
 + ) 0 + N d p
 4 2 + 20 ( 0 )2 + 20
D D
D


Z D  
D
q

= 2N 2 2 2
d D( + ) + N D(EF ) D D + 0 + 0 Arsinh
0

D
N V0 D2 (EF ) 20 Arsinh2 . (9.42)
0
With the gap equation
D
1 = V0 D(EF ) Arsinh (9.43)
0
this simplifies to

Z D q
EBCS
= 2N d D( + ) N D(EF ) D 2 + 2 .
D 0 (9.44)

The normal-state energy should be recovered by taking 0 0. The energy difference is


q
EBCS := EBCS EBCS |0 0 = N D(EF ) D D 2 + 2 + N D(E ) 2 . (9.45)
0 F D

Since for weak coupling we have 0  D , we can expand this in 0 /D ,


s  2
2 0 2 1
EBCS = N D(EF ) D 1 + + N D(EF ) D = N D(EF ) 20 . (9.46)
D 2

The condensation-energy density is thus (counted positively)

1N
eBCS
= D(EF ) 20 . (9.47)
2V
Type-I superconductivity is destroyed if eBCS equals the energy density required for magnetic-field expulsion. At
H = Hc , this energy is Hc2 /8, as we have seen above. We thus conclude that
r
N
Hc = 4 D(EF ) 0 . (9.48)
V
This prediction of BCS theory is in reasonably good agreement with experiments for simple superconductors.

88
10

BCS theory

The variational ansatz of Sec. 9.2 has given us an approximation for the many-particle ground state |BCS i. While
this is interesting, it does not yet allow predictions of thermodynamic properties, such as the critical temperature.
We will now consider superconductors at non-zero temperatures within mean-field theory, which will also provide
a new perspective on the BCS gap equation and on the meaning of k .

10.1 BCS mean-field theory


We start again from the Hamiltonian
1 X
k ck ck + Vkk0 ck ck, ck0 , ck0 .
X
H= (10.1)
N 0
k kk

A mean-field approximation consists of replacing products of operators A, B according to


AB
= hAi B + A hBi hAi hBi . (10.2)
Note that the error introduced by this replacement is
AB hAi B A hBi + hAi hBi = (A hAi)(B hBi), (10.3)
i.e., it is of second order in the deviations of A and B from their averages. A well-known mean-field approximation
is the Hartree or Stoner approximation, which for our Hamiltonian amounts to the choice A = ck ck0 , B =
ck, ck0 , . However, Bardeen, Cooper, and Schrieffer realized that superconductivity can be understood with
the help of a different choice, namely A = ck ck, , B = ck0 , ck0 . This leads to the mean-field BCS Hamiltonian
1 X  
k ck ck + Vkk0 hck ck, i ck0 , ck0 + ck ck, hck0 , ck0 i hck ck, ihck0 , ck0 i .
X
HBCS =
N 0
k kk
(10.4)
We define
1 X
k := Vkk0 hck0 , ck0 i (10.5)
N 0
k
so that
1 X
k = Vkk0 hck0 ck0 , i. (10.6)
N 0
k
At this point it is not obvious that the quantity k is the same as the one introduced in Sec. 9.2 for the special
case of the ground state. Since this will turn out to be the case, we nevertheless use the same symbol from the
start. We can now write
k ck ck k ck ck, + const.
X X X
HBCS = k ck, ck (10.7)
k k k

89
The constant is irrelevant for the following derivation and is omitted from now on. Since HBCS is bilinear in c, c
it describes a non-interacting effective system. But what is unusual is that HBCS contains terms of the form cc
and c c , which do not conserve the electron number. We thus expect that the eigenstates of HBCS do not have a
sharp electron number. We had already seen that the BCS ground state has this property. This is a bit strange
since superpositions of states with different electron numbers are never observed. One can formulate the theory
of superconductivity in terms of states with fixed electron number, but this formulation is cumbersome and we
will not pursue it here.
To diagonalize HBCS , we introduce new fermionic operators, which are linear combinations of electron creation
and annihilation operators,     
k uk vk ck
= . (10.8)
k, vk uk ck,
This mapping is called Bogoliubov (or Bogoliubov-Valatin) transformation. Again, it is not clear yet that uk , vk
are related to the previously introduced quantities denoted by the same symbols. For the to satisfy fermionic
anticommutation relations, we require
n o

k , k = k k + k k

= uk uk ck ck uk vk ck ck, vk uk ck, ck + vk vk ck, ck,


+ uk uk ck ck uk vk ck ck, vk uk ck, ck + vk vk ck, ck,
n o n o n o
= |uk | ck , ck uk vk {ck , ck, } vk uk ck, , ck + |vk | ck, , ck,
2 2

| {z } | {z } | {z } | {z }
=0
=1 =0 =1
2 2 !
= |uk | + |vk | = 1. (10.9)

Using this constraint, we find the inverse transformation,


    
ck u k vk k
= . (10.10)
ck, vk uk
k,

Insertion into HBCS yields


Xn 


 




HBCS = k uk k + vk k, uk k + vk k, + k vk k + uk k, vk k + uk k,
k
     o

k vk k + uk k, uk k + vk k, k uk k + vk k, vk k + uk k,
X n 2 2


= k |uk | k |vk | + k vk uk + k uk vk k k
k
 
2 2
+ k |vk | + k |uk | + k uk vk + k vk uk k, k,
 
2
+ k uk vk + k uk vk + k vk2 k (uk ) k k,
  o
2
+ k vk uk + k vk uk k u2k + k (vk ) k, k + const. (10.11)

The coefficients uk , vk should now be chosen such that the and terms vanish. This requires
2
2k uk vk + k vk2 k (uk ) = 0. (10.12)

Writing

k = |k | eik , (10.13)
ik
uk = |uk | e , (10.14)
ik
vk = |vk | e , (10.15)

90
we obtain  
2 2
2k |uk | |vk | ei(k k ) + |k | |vk | ei(2k k ) |uk | ei(k 2k ) = 0. (10.16)

A special solution of this equation (we do not require the general solution) is given by

k = 0, (10.17)
k = k , (10.18)
 
2 2
2k |uk | |vk | + |k | |vk | |uk | = 0. (10.19)

From the last equation we obtain


 
2 2 2 4 2 2 4
4k2 |uk | |vk | = |k | |vk | 2 |vk | |uk | + |uk | (10.20)
     2
2 2 2 2 4 2 2 4 2 2 2 2
4 k2 + |k | |uk | |vk | = |k | |vk | + 2 |vk | |uk | + |uk | = |k | |vk | + |uk | = |k | (10.21)
|k |
|uk | |vk | = q (10.22)
2
2 k2 + |k |

so that
2 2 2k |uk | |vk | k
|uk | |vk | = =q . (10.23)
|k | 2
k2 + |k |
2 2
Together with |uk | + |vk | = 1 we thus find

2 1 k
|uk | = 1+ q , (10.24)
2 2 2
k + |k |

2 1 k
|vk | = 1 q . (10.25)
2 2 2
k + |k |

Restoring the phases in Eq. (10.22), we also conclude that

k
uk vk = q . (10.26)
2 2
2 k + |k |

The BCS Hamiltonian now reads, ignoring a constant,



2 2
X
q k |k | 
k + k,

HBCS = +q k k,
2 2
k k2 + |k | k2 + |k |
Xq 2



= k2 + |k | k k + k, k, . (10.27)
k

Using k = k and the plausible assumption |k | = |k |, we obtain the simple form



X
HBCS = Ek k k (10.28)
k

with the dispersion q


2
Ek := k2 + |k | . (10.29)

91
It is instructive to first consider the normal state, for which k 0. Then
  (
2 1 k 0 for k < 0,
|uk | = 1+ = (10.30)
2 |k | 1 for k > 0,
  (
2 1 k 1 for k < 0,
|vk | = 1 = (10.31)
2 |k | 0 for k > 0.

We see that the Bogoliubov quasiparticles described by , , are holes for energies below the Fermi energy
(k < 0) and electrons for energies above (k > 0). Their dispersion is Ek = |k |. For a parabolic normal
dispersion k :
Ek

hole electron
excitations excitations
0 kF k

The excitation energies Ek are always positive except at the Fermi surfaceit costs energy to create a hole in
the Fermi sea and also to insert an electron into an empty
q state outside of the Fermi sea.
2
Superconductivity changes the dispersion to Ek = k2 + |k | :

Ek

kF
0 kF k

Superconductivity evidently opens an energy gap of magnitude |kF | in the excitation spectrum.
We should recall that in deriving HBCS we have ignored a constant, which we now reinsert,

X
HBCS = EBCS + Ek k k . (10.32)
k

The energy of the system is EBCS if no quasiparticles are present and is increased (by at least |kF |) if quasiparti-
cles are excited. The state without any quasiparticles is the pure condensate. Since EBCS depends on temperature
through hck, ck i, the condensate is not generally the BCS ground state discussed previously. However, one can
show that it agrees with the ground state in the limit T 0.

92
1 2
uk

2
vk
0 kF k
2 2
We also find 0 < |uk | < 1 and 0 < |vk | < 1, i.e., the Bogoliubov quasiparticles are superpositions of particles
and holes. Deep inside the Fermi sea, the quasiparticles are mostly hole-like, while far above EF they are mostly
electron-like. But right at the Fermi surface we find, for example for spin = ,
1 1
k = ck eik ck, . (10.33)
2 2
The quasiparticles here consist of electrons and holes with the same amplitude. This means that they are
electrically neutral on average.
So far, we have not determined the gap function k . This can be done by inserting the Bogoliubov transfor-
mation into the definition
1 X
k = Vkk0 hck0 , ck0 i, (10.34)
N 0
k

which yields
1 X D  E
k = Vkk0 vk0 k 0 + uk0 k0 , uk0 k0 + vk0 k
0 ,
N 0
k
1 X n D E D E D Eo
= Vkk0 vk0 uk0 k 0 k0 vk2 0 k 0 k

0 ,

+ u2k0 hk0 , k0 i + uk0 vk0 k0 , k0 , . (10.35)
N 0
k

For selfconsistency, the averages have to be evaluated with the BCS Hamiltonian HBCS . This gives
D E
k 0 k0 = nF (Ek0 ), (10.36)
D E
k 0 k

0 , = 0, (10.37)
hk0 , k0 i = 0, (10.38)
D E

k0 , k0 , = 1 nF (Ek0 ), (10.39)

and we obtain the BCS gap equation, now at arbitrary temperature,


1 X 1 X k0
k = Vkk0 uk0 vk0 [1 2 nF (Ek0 )] = Vkk0 q [1 2 nF (Ek0 )]
N 0 N 0 2
k k 2 k2 0 + |k0 |
1 X k 0
Vkk0 [1 2 nF (Ek0 )] . (10.40)
N 0 2Ek0
k

We see that nF (Ek0 ) 0 for Ek0 > 0 and T 0 so that the zero-temperature BCS gap equation (9.28) is
recovered as a limiting case.
For our model interaction and assuming a k-independent real gap, we obtain, in analogy to the ground-state
derivation,
p 
ZD 1 2 nF 2 + 20 ZD
tanh 2 2 + 20
p
1 = V0 d D( + ) p = V0 d D( + ) p . (10.41)
2 2 + 20 2 2 + 20
D D

93
Note that this only works for 0 6= 0 since we have divided by 0 . If the density of states is approximately
constant close to EF , the equation simplifies to
ZD
p
V0 D(EF ) tanh 2 2 + 20
1= d p . (10.42)
2 2 + 20
D

The integral is easily evaluated numerically, leading to the temperature dependence of 0 :


0

0 (0)

0 Tc T

For weak coupling we have already seen that


 
1
0 (0) = 2D exp . (10.43)
V0 D(EF )
We can also obtain an analytical expression for Tc : If T approaches Tc from below, we can take the limit 0 0
in the gap equation,
ZD ZD
ZD /2
tanh 2 || tanh 2 tanh x
1
= V0 D(EF ) d = V0 D(EF ) d = V0 D(EF ) dx
2 || x
D 0 0
(
ZD /2 )
by parts D D ln x
= V0 D(EF ) ln tanh dx . (10.44)
2 2 cosh2 x
0

In the weak-coupling limit we have D = D /kB T  1 (this assertion should be checked a-posteriori ). Since
the integrand of the last integral decays exponentially for large x, we can send the upper limit to infinity,
( Z )  
V0 D(EF ) ln D D ln x D
1= tanh dx = V 0 D(EF ) ln + ln , (10.45)
2 | {z 2 } cosh2 x 2 4
0

=1

where is again the Euler gamma constant. This implies


 
1 2D
exp = e (10.46)
V0 D(EF )
2e
 
1
kB Tc = D exp . (10.47)
V0 D(EF )
This is exactly the same expression we have found above for the critical temperature of the Cooper instability.
Since the approximations used are quite different, this agreement is not trivial. The gap at zero temperature and
the critical temperature thus have a universal ratio in BCS theory,
20 (0) 2
= 3.528. (10.48)
kB Tc e
This ratio is close to the result measured for simple elementary superconductors. For example, for tin one
finds 20 (0)/kB Tc 3.46. For superconductors with stonger coupling, such as mercury, and for unconventional
superconductors the agreement is not good, though.

94
10.2 Isotope effect
How can one check that superconductivity is indeed governed by a phonon-mediated interaction? BCS theory
predicts  
1
kB Tc , 0 D exp . (10.49)
V0 D(EF )
It would be ideal to compare kB Tc or 0 for superconductors that only differ in the Debye frequency D , not in
V0 of D(EF ). This is at least approximately possible by using samples containing different isotopes (or different
fractions of isotopes) of the same elements.
The eigenfrequency of a harmonic oscillator scales with the mass like
1
. (10.50)
m
The entire phonon dispersion, and thus in particular the Debye frequency, also scales like
1 1
q , D (10.51)
M M
with the atomic mass M for an elementary superconductor. The same scaling has been found above for the
jellium model, see Eq. (8.73). Consequently, for elementary BCS weak-coupling superconductors,
1
kB Tc , 0 M with =
. (10.52)
2
This is indeed found for simple superconductors. The exponent is found to be smaller or even negative for materials
that are not in the weak-coupling regime V D(EF )  1 or that are not phonon-mediated superconductors. In
particular, if the relevant interaction has nothing to do with phonons, we expect = 0. This is observed for
optimally doped (highest Tc ) cuprate high-temperature superconductors.

10.3 Specific heat


We now discuss further predictions following from BCS theory. We start by revisiting the heat capacity or specific
heat. The BCS Hamiltonian

X
HBCS = EBCS + Ek k k (10.53)
k
with q
2
Ek = k2 + |k | (10.54)
leads to the internal energy X
U = hHBCS i = EBCS + 2 Ek nF (Ek ). (10.55)
k
However, this is inconvenient for the calculation of the heat capacity C = dU/dT since the condensate energy
EBCS depends on temperature through hck, ck i. We better consider the entropy, which has no contribution
from the condensate. It reads
X
S = kB [(1 nF ) ln(1 nF ) + nF ln nF ], (10.56)
k

where nF nF (Ek ). From the entropy, we obtain the heat capacity


dS dS
C=T =
dT d
X d X dnF
= 2kB [(1 nF ) ln(1 nF ) + nF ln nF ] = 2kB [ ln(1 nF ) 1 + ln nF + 1]
d | {z } d
k k nF
= ln 1nF = ln eEk = Ek
X d nF (Ek )
= 2kB 2 Ek . (10.57)
d
k

95
q nF (Ek ) depends on or temperature both explicitly and through the temperature dependence of
Note that
2
Ek = k2 + |k (T )| :
! !
2 2
2
X nF nF 1 d |k | X nF 1 d |k |
C = 2kB Ek + = 2kB Ek2 + . (10.58)
Ek 2Ek d Ek 2 d
k | {z } k
Ek nF
= Ek

The first term is due to the explicit dependence, i.e., to the change of occupation of quasiparticle states with
temperature. The second term results from the temperature dependence of the quasiparticle spectrum and is
2
absent for T Tc , where |k | = 0 = const. The sum over k contains the factor

nF 1 eEk
= = 2 = nF (Ek )[1 nF (Ek )] (10.59)
Ek Ek eEk + 1 (eEk + 1)

so that !
2
2
X 1 d |k |
C = 2kB nF (1 nF ) Ek2 + . (10.60)
2 d
k

Here, nF (1 nF ) is exponentially small for Ek  kB T . This means that for kB T  min , where min is the
minimum superconducting gap, all terms in the sum are exponentially suppressed since kB T  min Ek . Thus
the heat capacity is exponentially small at low temperatures. This result is not specific to superconductorsall
systems with an energy gap for excitations show this behavior.
For the simple interaction used above, the heat capacity can be obtained in terms of an integral over energy.
The numerical evaluation gives the following result:
C

l state
norma
0 Tc T

We find a downward jump at Tc , reproducing a result obtained from Landau theory in section 6.1. The jump
occurs for any mean-field theory describing a second-order phase transition for a complex order parameter. Since
BCS theory is such a theory, it recovers the result.
The height of the jump can be found as follows: The Ek2 term in Eq. (10.60) is continuous for T Tc
(0 0). Thus the jump is given by

d20

1 X
C = 2 3 nF (k ) [1 nF (k )] . (10.61)
kB Tc d T Tc
k

To obtain 0 close to Tc , we have to solve the gap equation


ZD ZD
tanh 2 2 + 20 tanh 2 2 + 20
p p
1 = V0 D(EF ) d p = V0 D(EF ) d p (10.62)
2 2 + 20 2 + 20
D 0

for small 0 . Writing


1 1
= = (10.63)
kB T kB (Tc T )

96
and expanding for small T and small 0 , we obtain
ZD ZD
tanh 2kB Tc V0 D(EF ) d
1
= V0 D(EF ) d + T 2
2kB Tc2 cosh 2kB Tc
0 0
| {z }
= 1 (gap equation)
ZD
2kB Tc tanh 2kB Tc
!
V0 D(EF ) 2 d 1
+ 0 . (10.64)
4kB Tc 2 cosh2
2kB Tc

0

In the weak-coupling limit, D  kB Tc , we can extend the integrals to infinity, which yields
7(3)
V0 D(EF ) V0 D(EF ) 2 2 T 7(3) 20
0
= T 2k B Tc 0 = V 0 D(EF ) V 0 D(EF ) (10.65)
2kB Tc2 4kB Tc 2kB Tc Tc 8 2 (kB Tc )2
8 2 2
20
= k Tc T. (10.66)
7(3) B
Thus
d20 dT d20 2
8 2 3 3

2 d0
= = kB Tc = k T (10.67)
d T Tc
d dT T 0
dT T 0
7(3) B c
Z
1 X 8 2 3 3 8 2
C = 2 3 nF (k ) [1 nF (k )] k T = kB N D(EF ) d nF () [1 nF ()]
kB Tc 7(3) B c 7(3)
k
| {z }
= kB Tc

8 2 2
= N D(EF ) kB Tc , (10.68)
7(3)

where N is the number of unit cells. The specific-heat jump is

C 8 2 2
c = = d(EF ) kB Tc , (10.69)
V 7(3)

where d(EF ) = D(EF ) N/V is the density of states per volume.

10.4 Density of states and single-particle tunneling


Detailed experimental information on the excitation spectrum in a superconductor can be obtained from single-
particle tunneling between a superconductor and either a normal metal or another superconductor. We discuss
this in the following assuming, for simplicity, that the density of states D(E) in the normal state is approximately
constant close to the Fermi energy. We restrict ourselves to single-electron tunneling; pair tunneling, which leads
to the Josephson effects, will be discussed later.

Quasiparticle density of states


The density of states per spin of the Bogoliubov quasiparticles created by is easily obtained from their disper-
sion:  
1 X 1 X
q
2
Ds (E) = (E Ek ) = E k2 + |k | . (10.70)
N N
k k

97
The subscript s of Ds (E) stands for superconducting. For the case of approximately constant normal-state
density of states Dn and constant gap 0 , we find
Z  q  Z  q 
Ds (E) = 2 2
d Dn ( ) E + 0 = Dn (EF ) 2 2
d E + 0

   
Z
p p
+ E 2 2 E 2 2
0 0
= Dn (EF ) d +

|| 2 2

2 2
+0 +0

Dn (EF ) p 2E

for E > 0 ,
= E 2 20 (10.71)
0 for E < 0 .

Ds (E )
Dn (EF)

0 0 E

There are of course no states in the energy gap. Importantly, we find a divergence at the gap edge at E = 0 .
In the normal-state limit, 0 0, we obtain Ds (E)/Dn (EF ) 2, deviating from the result of unity given in
Tinkhams book. The origin is that E is an excitation energy relative to the Fermi sea, i.e., both electron and
hole excitations contribute to the density of states at positive E.

Single-particle tunneling
It is plausible that the density of states Ds (E) can be mapped out by tunneling experiments, for example using a
normal-metal/insulator/superconductor structure. However, we have to keep in mind that the particles tunneling
out of or into a normal metal are real electrons, whereas the quasiparticles in a superconductor are superpositions
of electrons and holes. To study tunneling effects theoretically, we employ a tunneling Hamiltonian of the form

H = HL + HR + HT , (10.72)

where HL and HR describe the materials to the left and right of the tunneling region. Either can be a normal
metal or a superconductor and is assumed to be unaffected by the presence of the tunneling region. For example,
while translational invariance is necessarily broken in a tunneling device, we nevertheless write HL,R in terms of
lattice-momentum states. HT describes the tunneling between the two materials,

tkq ck dq + h.c.
X
HT = (10.73)
kq

Here, c and d are electronic operators referring to the two sides of the tunneling barrier and tkq is a tunneling
matrix element, which might depend on the momenta of the incoming and outgoing electron. We assume a
non-magnetic tunneling barrier so that the electron spin is conserved and tkq does not depend on it. Note that in
the presence of interactions HT is an approximation valid for weak tunneling. We treat the two bulk materials in
the mean-field approximation so that HL,R are effectively non-interacting mean-field Hamiltonians. We further
assume constant normal-state densities of states and momentum-independent tunneling matrix elements t = tkq .

98
For tunneling between two normal metals, we can calculate the current for an applied voltage V using the
Landauer formula:
Inn = ILR IRL (10.74)
| {z } | {z }
left to right right to left

with
Z
2e n n 2
ILR d DL ( + ) DR ( + + eV ) |t| nF () [1 nF ( + eV )], (10.75)
h

Z
2e n n 2
IRL d DL ( + ) DR ( + + eV ) |t| nF ( + eV ) [1 nF ()], (10.76)
h

n n
where DL,R is the density of states per spin direction in the left/right normal metal. The factors of DL (+) nF ()
etc. can be understood as the probabilities that the relevant initial states exist and are occupied and the final
states exist and are empty. We get
Z
2e n n 2
Inn d DL ( + ) DR ( + + eV ) |t| [nF () nF ( + eV )]
h

Z
2e n n 2
= D (EF ) DR (EF ) |t| d [nF () nF ( + eV )]
h L

2e2 n n 2
= DL (EF ) DR (EF ) |t| V. (10.77)
h
We thus find ohmic behaviour,
Inn = Gnn V. (10.78)
We next consider the case that one material is normal while the other (without loss of generality the left) is
superconducting. Now there are additional factors because the quasiparticles in the superconductor are not pure
electrons or holes. Let us say an electron is tunneling out of the superconductor with energy > 0. First of all,
this is only possible if 0 because of the energy gap. Now the electron can either come from an electron-like
quasiparticle (k > kF ), which contains an electron portion of
! p !
2 1 k 1 2 20
|uk | = 1+ p 2 = 1+ , (10.79)

2 k + 20 2 2 2

k = +0

or from a hole-like quasiparticle (k 0 < kF ) with an electron portion of


! p !
2 1 k0 1 2 20
|uk0 | = 1+ p 2 = 1 . (10.80)

2 2 2 2 2

k0 + 0 k0 = +0

On the other hand, an electron tunneling out with energy < 0 is best described as a hole tunneling in with
energy > 0. The relevant factors are
p !
2 1 2 20
|vk | = 1 (10.81)
2 ||

and
p !
2 1 2 20
|vk0 | = 1+ . (10.82)
2 ||

99
Thus the current flowing from left to right is
Z
2e s n 2
ILR d DL () DR ( + + eV ) |t| nF () [1 nF ( + eV )]
h
0
( p ! p !)
1 2 20 1 2 20
1+ + 1
2 2
| {z }
=1
Z0
2e s n 2
+ d DL (||) DR ( + + eV ) |t| nF () [1 nF ( + eV )]
h

( p ! p !)
1 2 20 1 2 20
1 + 1+
2 || 2 ||
| {z }
=1
Z
2e s n 2
= d DL (||) DR ( + + eV ) |t| nF () [1 nF ( + eV )], (10.83)
h

s
where DL () now is the superconducting density of states neither containing a spin factor of 2 nor the factor
of 2 due describing both electrons and holes as excitations with positive energypositive and negative energies
are here treated explicitly and separately. We see that the electron-hole mixing does not lead to any additional
factors beyond the changed density of states. With the analogous expression
Z
2e s n 2
IRL d DL (||) DR ( + + eV ) |t| nF ( + eV ) [1 nF ()] (10.84)
h

we obtain
Z
2e s n 2
Isn d DL (||) DR ( + + eV ) |t| [nF () nF ( + eV )]
h

Z s
2e n n 2 DL (||)
= D (EF ) DR (EF ) |t| d [nF () nF ( + eV )], (10.85)
h L n
DL (EF )

where
s
DL (||) p ||

for || > 0 ,
n = 2 20 (10.86)
DL (EF )
0 for || < 0 .
Thus
Z s
Gnn DL (||)
Isn = d n [nF () nF ( + eV )]. (10.87)
e DL (EF )

It is useful to consider the differential conductance
Z
Ds (||)
 
dIsn nF ( + eV )
Gsn := = Gnn d nL
dV DL (EF )

Z s
DL (||)
= Gnn d n nF ( + eV )[1 nF ( + eV )]. (10.88)
DL (EF )

100
In the limit kB T 0 this becomes
Z s
DL (||) Ds (|eV |)
Gsn = Gnn d n ( + eV ) = Gnn Ln . (10.89)
DL (EF ) DL (EF )

Thus low-temperature tunneling directly measures the superconducting density of states. At non-zero tempera-
tures, the features are smeared out over an energy scale of kB T . In a band picture we can see that the Fermi
energy in the normal material is used to scan the density of states in the superconductor.

2 0

+ eV

S I N
For a superconductor-insulator-superconductor contact, we only present the result for the current without deriva-
tion. The result is very plausible in view of the previous cases:
Z s s
Gnn DL (||) DR (| + eV |)
Iss = d n n (E ) [nF () nF ( + eV )]. (10.90)
e DL (EF ) DR F

2 L
2 R
+ eV

S I S
Now we find a large change in the current when the voltage V is chosen such that the lower gap edge at one
side is aligned with the upper gap edge at the other side. This is the case for |eV | = L + R . This feature
will remain sharp at non-zero temperatures since the densities of states retain their divergences at the gap edges
as long as superconductivity is not destroyed. Effectively, we are using the density-of-states singularity of one
superconductor to scan the density of states of the other. A numerical evaluation of Iss gives the following typical
behavior (here for L = 2kB T , R = 3kB T ):

3.5

3.0

2.5

2.0
Iss

1.5

1.0

|L R|
0.5 L + R
0.0
0 1 2 3 4

eV

101
10.5 Ultrasonic attenuation and nuclear relaxation
To conclude the brief survey of experimental consequences of BCS theory, we discuss the effects of time-dependent
perturbations. They will be exemplified by ultrasonic attenuation and nuclear relaxation, which represent two
distinct ways in which quasiparticle interference comes into play. Quite generally, we write the perturbation part
of the Hamiltonian as
Bk0 0 k ck0 0 ck ,
XX
H1 = (10.91)
kk0 0

where Bk0 0 k are matrix elements of the perturbation between single-electron states of the non-interacting system.
In the superconducting state, we have to express c, c in terms of , ,

ck = uk k + vk k, , (10.92)

ck = uk k vk k, , (10.93)

where we have assumed uk = uk , vk = vk . Thus, with , 0 = 1,


  

Bk0 0 k uk0 k 0 0 + 0 vk 0 k0 ,0 uk k + vk k,
XX
H1 =
kk0 0

Bk0 0 k uk0 uk k 0 0 k + uk0 vk k 0 0 k,

XX
=
kk0 0


+ 0 vk 0 uk k0 ,0 k + 0 vk 0 vk k0 ,0 k, . (10.94)

It is useful to combine the terms containing Bk0 0 k and Bk,,k0 ,0 since both refer to processes that change
momentum by k0 k and spin by 0 . If the perturbation couples to the electron concentration, which is the
case for ultrasound, one finds simply
Bk,,k0 ,0 = Bk0 0 k . (10.95)
This is often called case I. Furthermore, spin is conserved by the coupling to ultrasound, thus

Bk0 0 k = 0 Bk0 k . (10.96)

Adding the two terms, we obtain


1 XX 
Hultra = Bk0 k uk0 uk k 0 k + uk0 vk k 0 k,

2 0
kk

+ vk 0 uk k0 , k + vk 0 vk k0 , k, + uk uk0 k, k0 ,


uk vk0 k, k 0 vk uk0 k k0 , + vk vk0 k k 0 . (10.97)

Assuming uk , vk , k R for simplicity, we get, up to a constant,


1 XX h  
Hultra = Bk0 k (uk0 uk vk0 vk ) k 0 k + k,

k0 ,
2 0
kk
 i
+ (uk0 vk + vk0 uk ) k 0 k,

+ k0 , k . (10.98)

We thus find effective matrix elements Bk0 k (uk0 uk vk0 vk ) for quasiparticle scattering and Bk0 k (uk0 vk +
vk0 uk ) for creation and annihilation of two quasiparticles. Transition rates calculated from Fermis golden rule
contain the absolute values squared of matrix elements. Thus the following two coherence factors will be impor-

102
tant. The first one is
( ! !
21 k0 k
(uk0 uk vk0 vk ) = 1+ p 2 1+ p 2
4 k0 + 2k0 k + 2k
! !)
k0 k0 k
2p 2 p k + 1 p 2 1 p 2
k0 + 2k0 k2 + 2k k0 + 2k0 k + 2k
 
1 k k0 k k 0
= 1+ (10.99)
2 Ek Ek0 Ek Ek0
p
with Ek = k2 + 2k . If the matrix elements B and the normal-state density of states are even functions of the
P relative to the Fermi energy, k , the second term, which is odd in k and k , will drop out
normal-state energy 0

under the sum kk0 . Hence, this term is usually omitted, giving the coherence factor
 
0 1 k k0
F (k, k ) := 1 (10.100)
2 Ek Ek0

relevant for quasiparticle scattering in ultrasound experiments. Analogously, we obtain the second coherence
factor  
k k0
2
(u
|{z} k k0 v + v 0
k ku )2
= 1 + =: F+ (k, k0 ) (10.101)
Ek Ek0
=1

for quasiparticle creation and annihilation.


We can gain insight into the temperature dependence of ultrasound attenuation by making the rather crude
approximation that the matrix element B is independent of k, k0 and thus of energy. Furthermore, typical
ultrasound frequencies satisfy ~  0 and ~  kB T . Then only scattering of quasiparticles by phonons
but not their creation is important since the phonon energy is not sufficient for quasiparticle creation.
The rate of ultrasound absorption (attenuation) can be written in a plausible form analogous to the current
in the previous section:
Z
2
s d Ds (||) Ds (| + |) |B| F (, + ) [nF () nF ( + )] , (10.102)

where now
20
 
0 1
F (, ) = 1 . (10.103)
2 || | 0 |
With the approximations introduced above we get
Z
20
 
2 Ds (||) Ds (| + |) 1
s Dn2 (EF ) |B| d 1 [nF () nF ( + )] . (10.104)
Dn (EF ) Dn (EF ) 2 || | + |

The normal-state attentuation rate is found by letting 0 0:


Z
2 nF () nF ( + ) 1 2
n Dn2 (EF ) |B| d = Dn2 (EF ) |B| (10.105)
2 2

Z
s 1 Ds (||) Ds (| + |) || | + | 20
= d [nF () nF ( + )] . (10.106)
n Dn (EF ) Dn (EF ) || | + |

103
Since is small, we can expand the integral up to linear order in ,
Z 2 2
20

s 1 Ds (||) nF
= d ()
n Dn (EF ) 2


Z 0 Z Z 0 Z
" #2
2 2
|| 0 nF nF
= + d p 2
= + d
2
2 0
0 0

= nF () + nF (0 ) nF (0 ) + nF () = nF (0 ) + 1 nF (0 )
| {z } | {z }
=0 =1
2
= 2 nF (0 ) = . (10.107)
e0 + 1
Inserting the BCS prediction for 0 (T ), we can plot s /n vs. temperature:
s
n
1

0 Tc T

We now turn to the relaxation of nuclear spins due to their coupling to the electrons. We note without derivation
that the hyperfine interaction relevant for this preocess is odd in momentum if the electron spin is not changed
but is even if the electron spin is flipped, i.e.,

Bk,,k0 ,0 = 0 Bk0 0 k (10.108)

(recall 0 = 1). This is called case II. The perturbation Hamiltonian now reads
1 XX 
HNMR = Bk0 0 k uk0 uk k 0 0 k + uk0 vk k 0 0 k,

2 0 0
kk

+ 0 vk 0 uk k0 ,0 k + 0 vk 0 vk k0 ,0 k, 0 uk uk0 k, k0 ,0


+ uk vk0 k, k 0 0 + 0 vk uk0 k k0 ,0 vk vk0 k k 0 0 . (10.109)

Assuming uk , vk , k R, we get, up to a constant,


1 XX h  
HNMR = Bk0 0 k (uk0 uk + vk0 vk ) k 0 0 k 0 k,

k0 ,0
2 0 0
kk
 i
+ (uk0 vk vk0 uk ) k 0 0 k,

0 k0 ,0 k . (10.110)

Compared to ultrasonic attenuation (case I) there is thus a change of sign in both coherence factors. An analogous
derivation now gives the interchanged coherence factors
 
0 1 k k0
F+ (k, k ) = 1+ (10.111)
2 Ek Ek0
for quasiparticle scattering and  
0 1 k k0
F (k, k ) = 1 (10.112)
2 Ek Ek0

104
for quasiparticle creation and annihilation. The relevant energy ~ is the Zeeman energy of a nuclear spin in the
applied uniform magnetic field and is small compared to the gap 0 . Thus we can again restrict ourselves to
the small- limit. The derivation is initially analogous to case I, but with F replaced by F+ . The nuclear-spin
relaxation rate is
Z
2
s d Ds (||) Ds (| + |) |B| F+ (, + ) [nF () nF ( + )] (10.113)

with
20
 
10
F+ (, ) = 1+ . (10.114)
2 || | 0 |
Thus
Z
20
 
2 Ds (||) Ds (| + |) 1
s Dn2 (EF ) |B| d 1+ [nF () nF ( + )] (10.115)
Dn (EF ) Dn (EF ) 2 || | + |

Z
s 1 Ds (||) Ds (| + |) || | + | + 20
= d [nF () nF ( + )] . (10.116)
n Dn (EF ) Dn (EF ) || | + |

If we now expand the integral for small as above, we encounter a problem:


Z " #2 Z
s 1 || 2 + 20 nF 2 + 20 nF
= 2 d () = 2 d . (10.117)
2 2 20
p
n 2 20
0 0

This integral diverges logarithmically at the lower limit. Keeping a non-zero but realistically small removes the
divergence. However, the calculated s /n is still too large compared to experiments. The origin of this problem
is the strong singularity in the superconducting density of states. A k-dependent gap k removes the problem;
in a realistic theory k always has a k dependence since it cannot have higher symmetry than the underlying
normal dispersion k . Introducing some broadening of the density of states by hand, we numerically find the
following temperature dependence:

s
n
case II (NMR)

case I

0 Tc T

There is a large maximum below the transition temperature, called the Hebel-Slichter peak. It results from the
factor Ds2 (||) (for 0) in the integrand,

2
Ds2 (||)
= Dn2 (EF ) , (10.118)
2 20

which for nuclear relaxation is not canceled by the coherence factor F+ , whereas for ultrasonic attenuation it is
canceled by F . Physically, the strong enhancement below Tc of the density of states of both initial and final
states at & 0 leads to increased nuclear relaxation.

105
10.6 Ginzburg-Landau-Gorkov theory
We conclude this chapter by remarking that Lev Gorkov managed, two years after the publication of BCS theory,
to derive Ginzburg-Landau theory from BCS theory. The correspondence is perfect if the gap is sufficiantly small,
i.e., T is close to Tc , and the electromagnetic field varies slowly on the scale of the Pippard coherence length 0
(see Sec. 5.4). These are indeed the conditions under which Ginzburg and Landau expected their theory to be
valid.
Gorkov used equations of motion for electronic Green functions, which he decoupled with a mean-field-like
approximation, which allowed for spatial variations of the decoupling term (r). The derivation is given in
Schrieffers book and we omit it here. Gorkov found that in order to obtain the Ginzburg-Landau equations, he
had to take

q = 2e, (10.119)
m = 2m, (10.120)

as anticipated, and (using our conventions)


p
7(3) p 0 (r)
(r) = ns , (10.121)
4 kB Tc
where
ns
n0s := . (10.122)
1 TTc

T Tc

Recall that ns 1 T /Tc close to Tc in Ginzburg-Landau theory. The spatially dependent gap is thus locally
proportional to the Ginzburg-Landau condensate wavefunction or order parameter (r).
Since we have already found that London theory is a limiting case of Ginzburg-Landau theory, it is also a
limiting case of BCS theory. But London theory predicts the two central properties of superconductors: Ideal
conduction and flux expulsion. Thus Gorkovs derivation also shows that BCS theory indeed describes a super-
conducting state. (Historically, this has been shown by BCS before Gorkov established the formal relationship
between the various theories.)

106
11

Josephson effects

Brian Josephson made two important predictions for the current flowing through a tunneling barrier between two
superconductors. The results have later been extended to various other systems involving two superconducting
electrodes, such as superconductor/normal-metal/superconductor heterostructures and superconducting weak
links. Rather generally, for vanishing applied voltage a supercurrent Is is flowing which is related to the phase
difference of the two condensates by
 Z 
2
Is = Ic sin ds A . (11.1)
0

We will discuss the critical current Ic presently. We consider the case without magnetic field so that we can
choose the gauge A 0. Then the Josephson relation simplifies to

Is = Ic sin . (11.2)

It should be noted that this DC Josephson effect is an equilibrium phenomenon since no bias voltage is applied.
The current thus continues to flow as long as the phase difference is maintained.
Secondly, Josephson predicted that in the presence of a constant bias voltage V , the phase difference would
evolve according to
d 2e
= V (11.3)
dt ~
(recall that we use the convention e > 0) so that an alternating current would flow,
 
2e
Is (t) = Ic sin 0 Vt . (11.4)
~

This is called the AC Josephson effect. The frequency


2eV
J := (11.5)
~
of the current is called the Josephson frequency. The AC Josephson effect relates frequencies (or times) to voltages,
which makes it important for metrology.

11.1 The Josephson effects in Ginzburg-Landau theory


We consider a weak link between two identical bulk superconductors. The weak link is realized by a short wire
of length L  and cross section A made from the same material as the bulk superconductors. We choose this
setup since it is the easiest to treat in Ginzburg-Landau theory since the parameters and are uniform, but the
only property that really matters is that the phase of the order parameter (r) only changes within the weak

107
link. We employ the first Ginzburg-Landau equation for A 0 assuming (r) to depend only on the coordinate
x along the wire,
2 f 00 (x) + f (x) f 3 (x) = 0 (11.6)
with r
(x)
f (x) = = (x). (11.7)
|()|

0 L x

We assume the two bulk superconductors to be uniform and to have a relative phase of . This allows us to
write (
1 for x 0,
f (x) = i
(11.8)
e for x L.
For the wire we have to solve Eq. (11.6) with the boundary conditions

f (0) = 1, f (L) = ei . (11.9)

Since L  , the first term in Eq. (11.6) is larger than the other two by a factor of order 2 /L2 , unless = 0,
in which case the solution is trivially f 1. It is thus sufficient to solve f 00 (x) = 0, which has the solution

Lx x
f (x) = + ei . (11.10)
L L
Inserting f (x) into the second Ginzburg-Landau equation (with A 0), we obtain
 
q~ e~
js = i [( 0 ) 0 ] = i [(f 0 ) f f f 0 ]
2m 2m
     
e~ 1 1 Lx x Lx x 1 1
= i 2ns + ei + ei + ei + ei
2m L L L L L L L L
   
e~ns x L x i e~ns x Lx
2 ei ei i
 
= i e e = 2 sin sin
m L L2 m L2 L2
e~ns
= 2 sin . (11.11)
mL
The current is obviously obtained by integrating over the cross-sectional area,
2e~ns A
Is = sin , (11.12)
m L
so that we get
2e~ns A
Ic = . (11.13)
m L
The negative sign is due to the negative charge 2e of the Cooper pairs. The amplitude of the current-phase
relation is clearly |Ic |.

108
Ginzburg-Landau theory also gives us the free energy of the wire. Since we have neglected the and terms
when solving the Ginzburg-Landau equation, we must for consistency do the same here,

ZL ZL 2 ZL
~2 0 ~2 ~2 2ns
 
2 1 1 i
F =A dx | (x)| = A dx + e =A dx 2 (1 cos )
4m 4m L L 4m L2
0 0 0
A ~2 ns
= (1 cos ) . (11.14)
L m
F

The free energy is minimal when the phases of the two superconductors coincide. Thus if there existed any
mechanism by which the phases could relax, they would approach a state with uniform phase across the junction,
a highly plausible result.
We can now also derive the AC Josephson effect. Assuming that the free energy of the junction is only changed
by the supercurrent, we have
d
F = Is V, (11.15)
dt
i.e., the electrical power. This relation implies that
F d
= Is V (11.16)
dt
A ~2 n s d 2e~ns A
sin = sin V (11.17)
L m dt m L
d 2e
= V, (11.18)
dt ~
as stated above. Physically, if a supercurrent is flowing in the presence of a bias voltage, it generates power. Since
energy is conserved, this power must equal the change of (free) energy per unit time of the junction.

11.2 Dynamics of Josephson junctions


For a discussion of the dynamical current-voltage characteristics of a Josephson junction, it is crucial to realize
that a real junction also
1. permits single-particle tunneling (see Sec. 10.4), which we model by an ohmic resistivity R in parallel to
the junction,
2. has a non-zero capacitance C.
This leads to the resistively and capacitively shunted junction (RCSJ) model represented by the following circuit
diagram:

C junction
R

109
The current through the device is the sum of currents through the three branches,
V dV
I= +C Ic sin , (11.19)
R dt
where we take Ic > 0 and have made the sign explicit. With
d 2e
= V (11.20)
dt ~
we obtain
~ d ~C d2
I= Ic sin . (11.21)
2eR dt 2e dt2
We introduce the plasma frequency r
2eIc
p := (11.22)
~C
and the quality factor
Q := p RC (11.23)
of the junction. This leads to
I 1 d2 1 d
= 2 2 sin (11.24)
Ic p dt Qp dt
and with := p t finally to
d2 1 d I
+ + sin = . (11.25)
d 2 Q d Ic
Compare this equation to the Newton equation for a particle moving in one dimension in a potential Vpot (x) with
Stokes friction,
dVpot
mx + x = (11.26)
dx
1 dVpot
x + x = . (11.27)
m m dx
This Newton equation has the same form as the equation of motion of if we identify

t , (11.28)
x , (11.29)
1
, (11.30)
m Q
1 I
Vpot (x) cos . (11.31)
m Ic
Thus the time dependence of ( ) corresponds to the damped motion of a particle in a tilted-washboard
potential

I
Ic
Vpot m

110
Equation (11.25) can be used to study a Josephson junction in various regimes. First, note that a stationary
solution exists as long as |I| Ic . Then

I
sin = const = and V 0. (11.32)
Ic
This solution does not exist for |I| > Ic . What happens if we impose a time-independent current that is larger
than the critical current? We first consider a strongly damped junction, Q  1. Then we can neglect the
acceleration term and write
1 d I
+ sin = (11.33)
Q d Ic
1 d I
= sin (11.34)
Q d Ic
d
I = Q d (11.35)
Ic + sin

Z 0
0 1+ I
tan

d I > Ic 2 Ic 2

Q ( 0 ) = I
= q arctan q . (11.36)
Ic + sin 0 I 2

1 I 2

1


0 Ic Ic 0

We are interested in periodic solutions for ei or mod 2. One period T is the time it takes for to change
from 0 to 2 (note that d /d < 0). Thus
2
d 0
Z
I > Ic 2
Qp T = I
= q (11.37)
Ic + sin 0 I 2

1
0 Ic
2 1 ~ 1 ~ 1
T = q = 2 q = p . (11.38)
Qp I 2

1 2eIc R I 2

1 eR I Ic2
2
Ic Ic

The voltage V d /dt is of course time-dependent but the time-averaged voltage is simply

ZT ZT
1 ~ 1 d ~ 1 ~ 1 p
V = dt V (t) = dt = [(T ) (0)] = = R I 2 Ic2 (11.39)
T 2e T dt 2e T | {z } e T
0 0 = 2
p
for I > Ic . By symmetry, V = R I 2 Ic2 for I < Ic . The current-voltage characteristics for given direct
current thus look like this:
I

Ic
V/R

0 V
Ic

For |I| Ic , the current flows without resistance. At Ic , non-zero DC and AC voltages set in gradually. For
|I|  Ic , the DC voltage approaches the ohmic result for a normal contact.

111
The solution for general Q requires numerical calculation but we can analyze the opposite case of weak
damping, Q  1. The stationary solution = const, V = 0 still exists for I Ic . The mechanical analogy
suggests that the time-dependent solution with periodic ei will be a very rapid slide down the washboard,
overlaid by a small-amplitude oscillation,

= t + , (11.40)
where  p and is periodic in time and small. Inserting this ansatz into the equation of motion we find
I
=Q p  p , (11.41)
Ic
p2

= 2 sin t. (11.42)

|{z}
1

Thus
p2

= 2 sin . (11.43)
p p
We convince ourselves that this is a good solution for Q  1: Inserting it into Eq. (11.25), we obtain for the
left-hand side
!
1 1 p p2
sin cos sin + 2 sin
p Q p Q p p p
   2  
 I 1 Ic  1 Ic 1
= sin 2 cos sin cos 2
cos +O . (11.44)
Ic Q I p p Q I p Q4
 p  p

To leading order in 1/Q this is just I/Ic , which agrees with the right-hand side. We thus find an averaged
voltage of
ZT  
1 ~ d ~ ~ I ~ 2eIc I
V = dt = = Q p = RC = RI, (11.45)
T 2e dt 2e 2e Ic 2e ~C Ic
0
i.e., the ohmic behavior of the normal junction. Note that the time-dependent solution exists for all currents,
not just for |I| > Ic . Thus for |I| Ic there are now two solutions, with V 0 and with V = RI. If we would
change the imposed current we could expect hysteretic behavior. This is indeed observed.
I

V/R
Ic

0 V
Ic

If we instead impose a constant voltage we obtain, for any Q,


d 2e d2
= V = const = 0 (11.46)
dt ~ dt2
and thus
 
1 2e 2e I(t)
V + sin V t + 0 = (11.47)
Qp ~ ~ Ic
 
Ic 2e 2e
I(t) = V + Ic sin V t 0 . (11.48)
Qp ~ ~

112
The averaged current is just
Ic 2e ~C Ic 2e V
I = V = V = . (11.49)
Qp ~ 2eIc RC ~ R
Note that this result holds for any damping. It is evidently important to carefully specify whether a constant
current or a constant voltage is imposed.

11.3 Bogoliubov-de Gennes Hamiltonian


It is often necessary to describe inhomogeneous systems, Josephson junctions are typical examples. So far, the only
theory we know that is able to treat inhomogeneity is the Ginzburg-Landau theory, which has the disadvantage
that the quasiparticles are not explicitly included. It is in this sense not a microscopic theory. We will now discuss
a microscopic description that allows us to treat inhomogeneous systems. The essential idea is to make the BCS
mean-field Hamiltonian spatially dependent. This leads to the Bogoliubov-de Gennes Hamiltonian. It is useful to
revert to a first-quantized description. To this end, we introduce the condensate state |BCS i as the ground state
of the BCS Hamiltonian

k ck ck k ck ck, + const.
X X X
HBCS = k ck, ck (11.50)
k k k

|BCS i agrees with the BCS ground state defined in Sec. 9.2 in the limit T 0 (recall that k and thus HBCS
is temperature-dependent). We have
HBCS |BCS i = EBCS |BCS i , (11.51)
where EBCS is the temperature-dependent energy of the condensate.
Since HBCS is bilinear, it is sufficient to consider single-particle excitations. Many-particle excitations are
simply product states, or more precisely Slater determinants, of single-particle excitations. We first define a
two-component spinor    
|k1 i ck
|k i := |BCS i . (11.52)
|k2 i ck,
It is easy to show that

[HBCS , ck ] = k ck k ck, , (11.53)


[HBCS , ck, ] = k ck, k ck . (11.54)

With these relations we obtain


 
HBCS |k1 i = HBCS ck |BCS i = k ck k ck, + ck HBCS |BCS i
= (EBCS + k ) |k1 i k |k2 i (11.55)

and
 
HBCS |k2 i = HBCS ck, |BCS i = k ck, k ck + ck, HBCS |BCS i
= (EBCS k ) |k2 i k |k1 i . (11.56)

Thus for the basis {|k1 i , |k2 i} the Hamiltonian has the matrix form
 
EBCS + k k
. (11.57)
k EBCS k
This is the desired Hamiltonian in first-quantized form, except that we want to measure excitation energies relative
to the condensate energy. Thus we write as the first-quantized Hamiltonian in k space
 
k k
HBdG (k) = . (11.58)
k k

113
This is the Bogoliubov-de Gennes Hamiltonian for non-magnetic superconductors. Its eigenvalues are
q
2
k2 + |k | = Ek (11.59)

with corresponding eigenstates


 
uk |k1 i vk |k2 i = uk ck vk ck, |BCS i = k

|BCS i (11.60)

and
 
vk |k1 i + uk |k2 i = vk ck + uk ck, |BCS i = k, |BCS i (11.61)

with uk , vk defined as above. (A lengthy but straightforward calculation has been omitted.) We can now un-
derstand why the second eigenvalue comes out negative: The corresponding eigenstate contains a quasiparticle
annihilation operator, not a creation operator. Hence, HBdG (k) reproduces the excitation energies we already
know.
The next step is to Fourier-transform the Hamiltonian to obtain its real-space representation, which we write
as  
1 X ikr H0 (r) (r)
HBdG (r) := e HBdG (k) = , (11.62)
N (r) H0 (r)
k

where we expect
~2 2
H0 (r) = + V (r) (11.63)
2m
as the free-electron Hamiltonian. But in this form it becomes easy to include spatially inhomogeneous situations:
Both V (r) and (r) can be chosen spatially dependent (and not simply lattice-periodic). The corresponding
Schrdinger equation
HBdG (r) (r) = E (r) (11.64)
with  
1 (r)
(r) = (11.65)
2 (r)
is called the Bogoliubov-de Gennes equation. Note that in this context the gap (r) is usually defined with the
opposite sign, which is just a phase change, so that the explicit minus signs in the off-diagonal components of
HBdG are removed. Furthermore, in Bogoliubov-de Gennes theory, the gap function is typically not evaluated
selfconsistantly from the averages hck, ck i. Rather, (r) is treated as a given function characterizing the
tendency of superconducting pairing.

11.4 Andreev reflection


As an application of the Bogoliubov-de Gennes approach, we study what happens to an electron that impinges on
a normal-superconducting interface from the normal side. We model this situation by the Bogoliubov-de Gennes
Hamiltonian !
~2
2m 2 0 (x)
HBdG = ~2 2
(11.66)
0 (x) 2m +

(note the changed sign of (r)) so that


HBdG (r) = E (r). (11.67)
In the normal region, x < 0, the two components 1 (r), 2 (r) are just superpositions of plane waves with wave
vectors k1 , k2 that must satisfy

k12 = 2m( + E) = kF2 + 2mE, (11.68)


k22 = 2m( E) = kF2 2mE, (11.69)

114
where ~ = 1. In the superconductor, x > 0, we have
 
1 2
1 (r) + 0 2 (r) = E 1 (r), (11.70)
2m
 
1 2
+ 2 (r) + 0 1 (r) = E 2 (r) (11.71)
2m
1
E + 2m 2 +
2 (r) = 1 (r) (11.72)
0
 2
1 2
+ 1 (r) = (E 2 20 ) 1 (r) (11.73)
2m

and analogously
 2
1 2
+ 2 (r) = (E 2 20 ) 2 (r). (11.74)
2m
If the energy is above the gap, |E| > 0 , the solutions are again plain wave vectors q1 , q2 , where now
!2
2
q1,2
= E 2 20 > 0 (11.75)
2m
 q 
2 2 2
q1,2 = 2m + E 0 , (11.76)

and amplitudes coupled by Eq. (11.72).


We are here interested in the more surprising case |E| < 0 . Since the solution must be continuous across
the interface and is plane-wave-like in the normal region, we make the ansatz

1 (r) = ei(k1y y+k1z z) 1 (x), (11.77)

from which !2
2
k1k1 d2
+ + 1 (x) = (E 2 20 ) 1 (x) (11.78)
2m 2m dx2
with k1k := (k1y , k1z ). Since this equation is linear with constant coefficients, we make an exponential ansatz

1 (x) = ex+iqx (11.79)

with , q R. This leads to


2
!2
k1k ( + iq)2
+ + = E 2 20 < 0 (11.80)
2m 2m
2
!2 !2
k1k + q2 iq 2
2
k1k + q2 2 iq
++ + =
2m m 2m 2m 2m m
2
! 2 !
k1k + q2 2 2 q 2
2
iq k1k + q
2
2
= 2 = E 2 20 . (11.81)
2m 2m m2 m 2m 2m

Since the right-hand side is real, we require


2
k1k + q2 2
=0 (11.82)
2m 2m
2 = k1k
2
+ q 2 2m. (11.83)

115
For the real part it follows that
2 q 2
= E 2 20 (11.84)
m2  
2
2 q 2 k1k + q 2 2m q 2
= = 20 E 2 > 0 (11.85)
m2  m

2

q 4 + k1k 2
2m q 2 m2 (20 E 2 ) = 0 (11.86)
s
2 2m k1k2 2
2m k1k

2
q = + m2 (20 E 2 )
2 2
" r #
1 2 2
2 2 2 2
= k k1k kF k1k + 4m2 (0 E 2 ) . (11.87)
2 F

Both solutions are clearly real but the one with the minus sign is negative so that q would be imaginary, contrary
to our assumption. Thus the relevant solutions are
s r
1 2
q = q1 := 2 2
kF k1k + kF2 k1k
2 + 4m2 (20 E 2 ). (11.88)
2
From this we get
" r #
1 2 2
2 2
= k1k kF2 + 2
k k1k + kF2 k1k
2 + 4m2 (20 E 2 )
2 F
" r #
1 2  2
= k kF2 + kF2 k1k
2 + 4m2 (20 E 2 ) (11.89)
2 1k

and s r 2
1 2 2
= 1 := k1k kF + kF2 k1k
2 + 4m2 (20 E 2 ). (11.90)
2
The positive root exists but would lead to a solution that grows exponentially for x . For 2 (r) the derivation
is completely analogous. However, 1 (r) and 2 (r) are related be Eq. (11.72), which for exponential functions
becomes a simple proportionality. Therefore, we must have k1k = k2k , which already implies q1 = q2 and 1 = 2 .
Then we have
" 2
#
1
E + 2m 2 + 1 (1 iq1 )2 k1k
2 (r) = 1 (r) = E+ + 1 (r)
0 0 2m 2m
" 2
#
1 k1k + q2 21 1 q1
= E ++ i 1 (r). (11.91)
0 | 2m {z 2m} m
=0

Since we already know that


21 q12
= 20 E 2 (11.92)
m2
and 1 , q1 have been defined as positive, we get
p
E i 20 E 2
2 (r) = 1 (r). (11.93)
0
We now write down an ansatz and show that it satisfies the Bogoliubov-de Gennes equation and the continuity
conditions at the interface. The ansatz reads
 
eik1 r + r eik1 r
(r) = for x 0, (11.94)
a eik2 r

116
2 2
with k1 := (k1x , k1y , k1z ) and k2 = (k2x , k1y , k1z ) with k2x > 0, where k1x + k1k = kF2 + 2mE and k2x
2 2
+ k1k =
2
kF 2mE, and

+ ei(q1 x+k1y y+k1z z) + ei(q1 x+k1y y+k1z z)


 
(r) = e1 x for x 0. (11.95)
+ ei(q1 x+k1y y+k1z z) + ei(q1 x+k1y y+k1z z)

Note that k1 is the wave vector of a specularly reflected electron. From Eq. (11.93) we get
p
E i 20 E 2
= . (11.96)
0
From the continuity of 1 , 2 , and their x-derivatives we obtain

1 + r = + + , (11.97)
a = + + , (11.98)
ik1x r ik1x = + (1 + iq1 ) + (1 iq1 ), (11.99)
a ik2x = + (1 + iq1 ) + (1 iq1 ). (11.100)

We thus have six coupled linear equations for the six unknown coefficients r, a, + , , + , . The equations
are linearly independent so that they have a unique solution, which we can obtain by standard methods. The
six coefficients are generally non-zero and complex. We do not give the lengthy expressions here but discuss the
results physically.
The solution in the superconductors decays exponentially, which is reasonable since the energy lies in the
superconducting gap.
In the normal region there is a secularly reflected electron wave (coefficient r), which is also expected. So
far, the same results would be obtained for a simple potential step. However, explicit evaluation shows that
2
in general |r| < 1, i.e., not all electrons are reflected.
There is also a term
2 (r) = a eik2 r for x 0. (11.101)
Recall that the second spinor component was defined by

|k2 i = ck, |BCS i . (11.102)


2
Hence, the above term represents a spin-down hole with wave vector k2 . Now k1k = k2k and k1x =
2 2 2 2 2
kF k1k + 2mE and k2x = kF k1k 2mE. But the last terms 2mE are small since

kF2
|E| < 0  = (11.103)
2m
in conventional superconductors. Thus |k2x k1x | is small and the hole is traveling nearly in the opposite
direction compared to the incoming electron wave. This phenomenon is called Andreev reflection.
N y,z S
electron electronlike
~
k1 quasiparticle
(evanescent)

electron k1 holelike
k2 quasiparticle
hole (evanescent)

117
Since not all electrons are reflected and in addition some holes are generated, where does the missing charge go?
The quasiparticle states in the superconductor are evanescent and thus cannot accommodate the missing charge.
The only possible explanation is that the charge is added to the superconducting condensate, i.e., that additional
Cooper pairs are formed. (The whole process can also run backwards, in which case Cooper pairs are removed.)
Recall that the condensate does not have a sharp electron number and can therefore absorb or emit electrons
without changing the state. But it can only absorb or emit electrons in pairs. The emerging picture is that if
an incoming electron is not specularly reflected, a Cooper pair is created, which requires a second electron. This
second electron is taken from the normal region, creating a hole, which, as we have seen, travels in the direction
the original electron was coming from.
N S

e
Cooper
h pair

Andreev bound states


An interesting situation arises if a normal region is delimited by superconductors on two sides. We here only
qualitatively consider a superconductor-normal-superconductor (SNS) hetero structure. Similar effects can also
occur for example in the normal core of a vortex.
If no voltage is applied between the two superconductors, an electron in the normal region, with energy
within the gap, is Andreev reflected as a hole at one interface. It is then Andreev reflected as an electron at
the other interface. It is plausible that multiple reflections can lead to the formation of bound states. The real
physics is somewhat more complicated since the electron is also partially specularly reflected as an electron. It is
conceptually clear, though, how to describe Andreev bound states within the Bogoliubov-de Gennes formalism:
We just have to satisfy continuity conditions for both interfaces.
S N S

electron
2 0 2 0
hole

If Andreev reflection dominates, as assumed for the sketch above, a Cooper pair is emitted into the right su-
perconductor for every reflection at the right interface. Conversely, a Cooper pair is absorbed from the left
superconductor for every reflection at the left interface. This corresponds to a supercurrent through the device.
Andreev bound states thus offer a microscopic description of the Josephson effect in superconductor-normal-
superconductor junctions.
If we apply a voltage V , the situation changes dramatically: If an electron moving, say, to the right, increases
its kinetic energy by eV due to the bias voltage, an Andreev reflected hole traveling to the left also increases its
kinetic energy by eV since it carries the opposite charge. An electron/hole Andreev-reflected multiple times can
thus gain arbitrarily high energies for any non-vanishing bias voltage.

118
S N S

2 0
2eV eV

In particular, an electron-like quasiparticle from an occupied state below the gap in, say, the left superconductor
can after multiple reflections emerge in a previously unoccupied state above the gap in the right superconductor.
A new transport channel becomes available whenever the full gap 20 is an odd integer multiple of eV :

20 = (2n + 1) eV, n = 0, 1, . . . (11.104)


0
eV = , n = 0, 1, . . . (11.105)
n + 12

The case n = 0 corresponds to direct quasiparticle transfer from one superconductor to the other, similar to
quasiparticle tunneling in a superconductor-insulator-superconductor junction. The opening of new transport
channels for n = 0, 1, . . . , i.e., at
2 2 2
eV = 0 , 0 , 0 , . . . (11.106)
3 5 7
leads to structures in the current-voltage characteristics below the gap, specifically to peaks in the differential
conductance dI/dV .
dI
dV

2
2 0 0 3 0 2 0 eV
2
5 0

119
12

Unconventional pairing

In this chapter we first discuss why interactions different from the phonon-mediated one might lead to uncon-
ventional pairing, that is to a gap function k with non-trivial k dependence. Then we will briefly consider the
origin of such interactions.

12.1 The gap equation for unconventional pairing


We will still use the BCS gap equation even when discussing unconventional superconductors. While the BCS
mean-field theory is not quantitatively correct in such cases, it will give a clear understanding of why the gap k
can have non-trivial symmetry. To get started, we briefly review results from BCS theory. The screened effective
interaction was derived in Sec. 8.4,
RPA (in )2
Veff (q, in ) = VCRPA (q, in ) , (12.1)
(in )2 q2 (in )
where VCRPA is the screened Coulomb interaction and q is the renormalized phonon dispersion. The retarded
interaction at small frequencies is
e2 2
R
Veff (q, in )
= 4 , (12.2)
q 2 + 2s 2 qR ()2 + i0+ sgn
where s is the inverse screening length. The bevavior at small distances r and small but non-zero frequency
R
is determined by Veff at large q, where q can by approximated by the Debye frequency. The interaction is thus
attractive and decays like 1/r for small r. The interaction is strongest at the same site in a tight-binding model.
In order to understand the physics, it makes sense to replace the interaction by a simplified one that is completely
local (attractive Hubbard model) or, equivalently, constant in k space, as we have done above. However, the BCS
gap equation
1 X k0
k = Vkk0 [1 nF (Ek0 )] (12.3)
N 0 2Ek0
k
is in fact much more general. In the gap equation, Vkk0 describes the amplitude for scattering of two electrons
with momenta k0 and k0 and opposite spins into states with momenta k and k.
Let us first consider the case that the interaction is local in real space (flat in k space) but repulsive. This
would apply if the phonons were for some reason ineffective in overscreening the Coulomb interaction. Then we
obtain
1 X k 0
k = V0 [1 nF (Ek0 )] (12.4)
N 0 2Ek0
k
with V0 > 0. The right-hand side is clearly independent of k so that we have k = 0 and can cancel a factor
of 0 if it is non-zero:
V0 X 1 nF (Ek0 )
1= . (12.5)
N 0 2Ek0
k

120
But now the right-hand side is always negative. Consequently, there is no non-trivial solution and thus no
superconductivity for a k-independent repulsion.
Now let us look at a strong interaction between nearest-neighbor sites. We consider a two-dimensional square
lattice for simplicity and since it is thought to be a good model for the cuprates. In momentum space, a nearest-
neighbor interaction is written as
Vkk0 = 2V1 [cos (kx kx0 )a + cos (ky ky0 )a], (12.6)
where V1 > 0 (V1 < 0) for a repulsive (attractive) interaction.
The electronic properties of a superconductor are most strongly affected by the states close to the normal-state
Fermi surface since elsewhere |k | is small compared to the normal-state energy |k |. We therefore concentrate on
these states. Also note that the coupling to states away from the Fermi surface is suppressed in the gap equation
by the factor k0 /Ek0 . This is a relatively weak, power-law suppression.
Let us first consider the repulsive case V1 > 0. Then Vkk0 is most strongly repulsive (positive) for momentum
transfer k k0 0. But k should be a smooth function of k, thus for k and k0 close together, k and k0 are
also similar. In particular, k will rarely change its sign between k and k0 . Consequently, the right-hand side
of the gap equation always contains a large contribution with sign opposite to that of k , coming from the sum
over k0 close to k. Hence, a repulsive nearest-neighbor interaction is unlikely to lead to superconductivity.
For the attractive case, V1 < 0, Vkk0 is most strongly attractive for k k0 0 and most strongly repulsive
for k k0 (/a, /a) and equivalent points in the Brillouin zone. The attraction at small q = k k0 is always
favorable for superconductivity. However, we also have an equally strong repulsion around q (/a, /a). A
critical situation thus arises if both k and k0 lie close to the Fermi surface and their difference is close to (/a, /a).
The central insight is that this can still help superconductivity if the gaps k and k0 at k and k0 , respectively,
have opposite sign. In this case the contribution to the right-hand side of the gap equation from such k0 has the
same sign as k since Vkk0 > 0 and there is an explicit minus sign.
This effect is crucial in the cuprates, which do have an effective attractive nearest-neighbor interaction and
have a large normal-state Fermi surface shown here for a two-dimensional model:
ky

0 kx

The vector Q in the sketch is Q = (/a, /a). Following the previous discussion, k close to the Fermi surface
should have different sign between points separated by Q. On the other hand, the small-q attraction favors gaps
k that change sign as little as possible. By inspection, these conditions are met by a gap changing sign on the
diagonals:
ky

0 kx

121
This type of gap is called a dx2 y2 -wave (or just d-wave) gap since it has the symmetry of a dx2 y2 -orbital
(though in k-space, not in real space). The simplest gap function with this symmetry and consistent with the
lattice structure is
k = 0 (cos kx a cos ky a). (12.7)
Recall that the gap function away from the Fermi surface is of limited importance.
The d-wave gap k is distinct from the conventional, approximately constant s-wave gap in that it has zeroes
on the Fermi surface. These zeroes are called gap nodes. In the present case they appear in the (11) and equivalent
directions. The quasiparticle dispersion in the vicinity of such a node is
q q
Ek = k2 + |k |
2
= (k )2 + 20 (cos kx a cos ky a)2 . (12.8)

One node is at  
kF 1
k0 = . (12.9)
2 1
Writing k = k0 + q and expanding for small q, we obtain
q
Ek0 +q
= (vF q)2 + 20 [(sin kx0 a)qx a + (sin ky0 a)qy a]2 , (12.10)

where  
k vF 1
vF := = (12.11)
k k=k0 2 1
is the normal-state Fermi velocity at the node. Thus
s   2 q
2 kF a kF a
Ek0 +q = (vF q) + 0 a sin , a sin
2 q = (vF q)2 + (vqp q)2 , (12.12)
2 2

where  
kF a 1
vqp := 0 a sin vF . (12.13)
2 1
Thus the quasiparticle dispersion close to the node is a cone like for massless relativistic particles, but with
different velocities in the directions normal and tangential to the Fermi surface. Usually one finds

vF > vqp . (12.14)

The sketch shows equipotential lines of Ek .


ky

kx

The fact that the gap closes at some k points implies that the quasiparticle density of states does not have a gap.

122
At low energies we can estimate it from our expansion of the quasiparticle energy,
1 X
Ds (E) = (E Ek )
N
k
 
4 X
q

= E (vF q)2 + (vqp q)2
N q
d2 q
Z  q 

= 4auc E (v F q)2 + (v
qp q)2
(2)2
d2 u 
Z
4auc q 
= 2
E u2x + u2y
vF vqp (2)
Z
2auc 2auc
= du u (E u) = E, (12.15)
vF vqp 0 vF vqp

where auc is the area of the two-dimensional unit cell. We see that the density of states starts linearly at small
energies. The full dependence is sketched here:

Ds (E )
Dn (EF)

0 max E

An additional nice feature of the d-wave gap is the following: The interaction considered above is presumably
not of BCS (Coulomb + phonons) type. However, there should also be a strong short-range Coulomb repulsion,
which is not overscreened by phonon exchange. This repulsion can again be modeled by a constant V0 > 0 in
k-space. This additional interaction adds the term

1 X 1 nF (Ek0 )
V0 k 0 (12.16)
N 0 2Ek0
k

to the gap equation. But since Ek0 does not change sign under rotation of k0 by /2 (i.e., 90 ), while k0 does
change sign under this rotation, the sum over k0 vanishes. d-wave pairing is thus robust against on-site Coulomb
repulsion.

12.2 Cuprates
Estimates of Tc based on phonon-exchange and using experimentally known values of the Debye frequency,
the electron-phonon coupling, and the normal-state density of states are much lower than the observed critical
temperatures. Also, as we have seen, such an interaction is flat in k-space, which favors an s-wave gap. An s-wave
gap is inconsistant with nearly all experiments on the cuprates that are sensitive to the gap. The last section has
shown that dx2 y2 -wave pairing in the cuprates is plausible if there is an attractive interaction for momentum
transfers q (/a, /a). We will now discuss where this attraction could be coming from.

123
T

AFM SC
0 hole doping x

A glance at typical phase diagrams shows that the undoped cuprates tend to be antiferromagnetic. Weak hole
doping or slightly stronger electron doping destroy the antiferromagnetic order, and at larger doping, supercon-
ductivity emerges. At even larger doping (the overdoped regime), superconductivity is again suppressed. Also
in many other unconventional superconductors superconductivity is found in the vicinity of but rarely coexist-
ing with magnetic order. This is true for most pnictide and heavy-fermion superconductors. The vicinity of a
magnetically ordered phase makes itself felt by strong magnetic fluctuations and strong, but short-range, spin
correlations. These are seen as an enhanced spin susceptibility.
At a magnetic second-order phase transition, the static spin susceptibility q diverges at q = Q, where Q is
the ordering vector. It is Q = 0 for ferromagnetic order and Q = (/a, /a) for checkerboard (Nel) order on
a square lattice. Even some distance from the transition or at non-zero frequencies , the susceptibility q ()
tends to have a maximum close to Q. Far away from the magnetic phase or at high frequencies this remnant of
magnetic order becomes small. This discussion suggests that the exchange of spin fluctuations, which are strong
close to Q, could provide the attractive interaction needed for Cooper pairing.

The Hubbard model


The two-dimensional, single-band, repulsive Hubbard model is thought (by many experts, not by everyone) to be
the simplest model that captures the main physics of the cuprates. The Hamiltonian reads, in real space,

tij ci cj + U
X
ci ci ci ci
X
H= (12.17)
ij i

with U > 0 and, in momentum space,


U X
k ck ck + c 0
X
H= c ck0 ck . (12.18)
N 0 k+q, k q,
k kk q

Also, the undoped cuprate parent compounds have, from simple counting, an odd number of electrons per unit
cell. Thus for them the single band must be half-filled, while for doped cuprates it is still close to half filling. The
underlying lattice in real space is a two-dimensional square lattice with each site i corresponding to a Cu+ ion.

Cu+ O 2
The transverse spin susceptibility is defined by

+ (q, ) = T S + (q, ) S (q, 0) ,




(12.19)

124
where is the imaginary time, T is the time-ordering directive, and

S (q, ) := S x (q, ) iS y (q, ) (12.20)

with
1 X 0
S (q, ) := ck+q, ( ) ck0 ( ) (12.21)
N k0 2

are electron-spin operators. = ( x , y , z ) is the vector of Pauli matrices. The susceptibility can be rewritten
as
1 XD     E
+ (q, ) = T ck+q, ck ( ) ck0 q, ck0 (0) . (12.22)
N 0
kk

In the non-interacting limit of U 0, the average of four fermionic operators can be written in terms of products
of averages of two operators (Wicks theorem). The resulting bare susceptibility reads
1 XD E D E
+
0 (q, ) = T ck+q, ( ) ck+q, (0) T ck ( ) ck (0)
N 0 0
k
1 X D E D E
= T ck+q, (0) ck+q, ( ) T ck ( ) ck (0)
N 0 0
k
1 X
0 0
= Gk+q, ( ) Gk ( ). (12.23)
N
k

The Fourier transform as a function of the bosonic Matsubara frequency in is

Z
+
0 (q, in ) = d ein +
0 (q, )
0
Z
1 X1X 0 1 X in 0
= d ein ein ( ) Gk+q,
0
(in0 ) e Gk (in )
N 0 i
0 k in n

1 X 1 X 1 X1X 0
= 0
n +n0 ,n Gk+q, (in0 ) Gk
0
(in ) = G 0
(in in ) Gk (in )
N 2 0
N i k+q,
k in ,in k n

1 X1X 1 1
= . (12.24)
N i in in k+q in k
k n

This expression can be written in a more symmetric form by making use of the identity k = k and replacing
the summation variables k by k q and in by in + in . The result is
1 X1X 1 1
+
0 (q, in ) = . (12.25)
N i in k in + in k+q
k n

The Matsubara frequency sum can be evaluated using methods from complex analysis. We here give the result
without proof,
1 X nF (k ) nF (k+q )
+
0 (q, in ) = . (12.26)
N in + k k+q
k

Note that the same result is found for the bare charge susceptibility except for a spin factor of 2. Diagramatically,
the result can be represented by the bubble diagram

k , , in
1
+
0 (q, in ) = 0 (q, in ) = . (12.27)
2
k + q , , in + i n

125
What changes when we switch on the Hubbard interaction U ? In analogy with the RPA theory for the screened
Coulomb interaction we might guess that the RPA spin susceptibility is given by

? U
+
RPA = + + (12.28)

but the second and higher terms vanish since they contain vertices at which the Hubbard interaction supposedly
flips the spin,
U , (12.29)

which it cannot do. On the other hand, the following ladder diagrams do not vanish and represent the RPA
susceptibility:

+
RPA = + U + + (12.30)

Since the Hubbard interaction does not depend on momentum, this series has a rather simple mathematical form,

+ + + + + + +
RPA (q, in ) = 0 (q, in ) + 0 (q, in ) U 0 (q, in ) + 0 (q, in ) U 0 (q, in ) U 0 (q, in ) +
= + + + +
0 (q, in )[1 + U 0 (q, in ) + U 0 (q, in ) U 0 (q, in ) + ]
+
0 (q, in )
= (12.31)
1 U +
0 (q, in )

[the signs in the first line follow from the Feynman rules, in particular each term contains a single fermionic loop,
which gives a minus sign, which cancels the explicit one in Eq. (12.30)]. The RPA spin susceptibility can be
evaluated numerically for given dispersion k . It is clear that it predicts an instability of the Fermi liquid if the
static RPA spin susceptibility

0+,R (q, 0)
+,R + +
RPA (q, 0) = RPA (q, in + i0 )|0 (12.32)
1 U +,R
0 (q, 0)

diverges at some q = Q, i.e., if


U +,R
0 (Q, 0) = 1. (12.33)
Since the spin susceptibility diverges, this would be a magnetic ordering transition with ordering vector Q. Note
that the RPA is not a good theory for the antiferromagnetic transition of the cuprates since it is a resummation
of a perturbative series in U/t, which is not small in cuprates. t is the typical hopping amplitude. Nevertheless
it gives qualitatively reasonable results in the paramagnetic phase, which is of interest for superconductivity.
The numerical evaluation at intermediate doping and high temperatures gives a broad and high peak in
+,R
RPA (q, 0) centered at Q = (/a, /a). This is consistant with the ordering at Q observed at weak doping.

+RPA
, R
((q, q), 0)

q
0 2
a a

126
At lower temperatures, details become resolved that are obscured by thermal broadening at high T . The RPA
and also more advanced approaches are very sensitive to the electronic bands close to the Fermi energy; states
with |k |  kB T have exponentially small effect on the susceptibility. Therefore, the detailed susceptibility at low
T strongly depends on details of the model Hamiltonian. Choosing nearest-neighbor and next-nearest-neighbor
hopping in such a way that a realistic Fermi surface emerges, one obtains a spin susceptiblity with incommensurate
peaks at a (1, 1 ) and a (1 , 1).

qy

0 qx
a

These peaks are due to nesting: Scattering is enhanced between parallel portions of the Fermi surface, which in
turn enhances the susceptibility [see M. Norman, Phys. Rev. B 75, 184514 (2007)].
ky


a
(1 , 1)
a
(1, 1 )
a

kx
a

The results for the spin susceptibility are in qualitative agreement with neutron-scattering experiments. However,
the RPA overestimates the tendency toward magnetic order, which is reduced by more advanced approaches.

Spin-fluctuation exchange
The next step is to construct an effective electron-electron interaction mediated by the exchange of spin fluctua-
tions. The following diagrammatic series represents the simplest way of doing this, though certainly not the only
one:
U
Veff := + + + (12.34)

127
Note that the external legs do not represent electronic Green functions but only indicate the states of incoming
and outgoing electrons. The series is very similar to the one for the RPA susceptibility. Indeed, the effective
interaction is
Veff (q, in ) = U + U + + +
0 (q, in ) U + U 0 (q, in ) U 0 (q, in ) +
= U + U 2 [+ + +
0 (q, in ) + 0 (q, in ) U 0 (q, in ) + ]
= U + U 2 +
RPA (q, in ). (12.35)
Typically one goes beyond the RPA at this point by including additional diagrams. In particular, also charge
fluctuations are included through the charge susceptibility and the bare Green function G 0 is replaced by a
selfconsistent one incorporating the effect of spin and charge fluctuations on the electronic self-energy. This leads
to the fluctuation-exchange approximation (FLEX ). One could now obtain the Cooper instability due to the FLEX
effective interaction in analogy to Sec. 9.1 and use a BCS mean-field theory to describe the superconducting state.
However, since the system is not in the weak-coupling limitthe typical interaction times the electronic density
of states is not smallone usually employs a strong-coupling generalization of BCS theory known as Eliashberg
theory. Since the effective interaction is, like the spin susceptibility, strongly peaked close to (/a, /a), it favors
dx2 y2 -wave pairing, as we have seen.
The numerical result of the FLEX for Tc and for the superfluid density ns are sketched here:

Tc ns

0 doping x
The curve for Tc vs. doping does not yet look like the experimentally observed dome. There are several aspects
that make the region of weak doping (underdoping) difficult to treat theoretically. One is indicated in the sketch:
ns is strongly reduced, which indicates that superconductivity may be in some sense fragile in this regime.
Furthermore, the cuprates are nearly two-dimensional solids. In fact we have used a two-dimensional model so
far. If we take this seriously, we know from chapter 7 that any mean-field theory, which Eliashberg theory with
FLEX effective interaction still is, fails miserably. Instead, we expect a BKT transition at a strongly reduced
critical temperature. We reinterpret the FLEX critical termperature as the mean-field termperature TMF and the
FLEX superfluid density as the unrenormalized superfluid density n0s . We have seen in chapter 7 that the bare
stiffness K0 is proportional to n0s . The BKT transition temperature Tc is defined by K(l ) = 2/. It is thus
reduced by small n0s , corresponding to a small initial value K(0) K0 . This is physically clear: Small stiffness
makes it easy to create vortex-antivortex pairs. Tc is of course also reduced by TMF and can never be larger than
TMF . A BKT theory on top of the FLEX gives the following phase diagram, which is in qualitative agreement
with experiments:
T
TMF ns

vortex
fluctuations

Tc superconductor
0 x
This scenario is consistent with the Nernst effect (an electric field measured normal to both an applied magnetic
field and a temperature gradient) in underdoped cuprates, which is interpreted in terms of free vortices in a broad
temperature range, which we would understand as the range from Tc to TMF .

128
Also note that spin fluctuations strongly affect the electronic properties in underdoped cuprates up to a
temperature T significantly higher than TMF . For example, below T the electronic density of states close to
the Fermi energy is suppressed compared to the result of band-structure calculations. The FLEX describes this
effect qualitatively correctly. This suppression is the well-known pseudogap.

Quantum critical point


The previously discussed approach relies on a resummation of a perturbative series in U/t. This is questionable
for cuprates, where U/t is on the order of 3. Many different approaches have been put forward that supposedly
work in this strong-coupling regime. They emphasize different aspects of the cuprates, showing that it is not even
clear which ingredients are the most important for understanding the phase diagram. Here we will review a lign
of thought represented by Chandra Varma and Subir Sachdev, among others. Its starting point is an analysis of
the normal region of the phase diagram.
T
T*
strange metal
~T
pseudo Fermi
gap liquid
AFM

superconductor
0 x
Roughly speaking, there are three regimes in the normal-conducting state:
a pseudo-gap regime below T at underdoping, in which the electronic density of states at low energies is
suppressed,
a strange-metal regime above the superconducting dome, without a clear suppression of the density of
states but with unusual temperature and energy dependencies of various observables, for example a resistivity
linear in temperature, T ,
an apparently ordinary normal-metal (Fermi-liquid) regime at overdoping, with standard const + T 2
dependence.
The two crossover lines look very much like what one expects to find for a quantum critical point (QCP), i.e., a
phase transition at zero temperature.
T
quantum
critical
affected affected
by phase I by phase II

phase I phase II
QCP tuning parameter
c

The regions to the left and right have a characteristic energy scale () inherited from the ground state (T = 0),
which dominates the thermal fluctuations. The energy scales go to zero at the QCP, i.e., for c from both
sides. Right at the QCP there then is no energy scale and energy or temperature-dependent quantities have to be
power laws. The emerging idea is that the superconducting dome hides a QCP between antiferromagnetic (spin-
density-wave) and paramagnetic order at T = 0. The spin fluctuations associated with this QCP become stronger

129
as it is approached. Consequently, the superconductivity caused by them is strongest and has the highest Tc right
above the QCP. Compare the previous argument: There, the spin fluctuations are assumed to be strongest close
to the (finite-temperature) antiferromagnetic phase and one needs to envoke small ns and vortex fluctuations to
argue why the maximum Tc is not close to the antiferromagnetic phase.

12.3 Pnictides
The iron pnictide superconductors are a more heterogeneous group than the cuprates. However, for most of
them the phase diagram is roughly similar to the one of a typical cuprate in that superconductivity emerges at
finite doping in the vicinity of an antiferromagnetic phase. This antiferromagnetic phase is metallic, not a Mott
insulator, though, suggesting that interactions are generally weaker in the pnictides.
T
CeFeAsO1x Fx
tet horo
140 K ort
rag mb
on
al
ic

AFM superconductor

0 0.05 doping x

The crystal structure is quasi-two-dimensional, though probably less so than in the cuprates. The common
structural motif is an iron-pnictogen, in particular Fe2+ As3 , layer with Fe2+ forming a square lattice and As3
sitting alternatingly above and below the Fe2+ plaquettes.

As 3
Fe 2+

While the correct unit cell contains two As and two Fe ions, the glide-mirror symmetry with respect to the Fe
plane allows to formulate two-dimensional models using a single-iron unit cell. The fact that different unit cells
are used leads to some confusion in the field. Note that since the single-iron unit cell is half as large as the
two-iron unit cell, the corresponding single-iron Brillouin zone is twice as large as the two-iron Brillouin zone.
A look at band-structure calculations or angular resolved photoemission data shows that the Fermi surface of
pnictides is much more complicated than the one of cuprates. The kz = 0 cut typically shows five Fermi pockets.

130
ky
M
X holelike

Q2
electronlike
X kx
Q1

singleiron Brillouin zone

The (probably outer) hole pocket at and the electron pockets at X and X0 are well nested with nesting vectors
Q1 = (/a, 0) and Q2 = (0, /a), respectively. Not surprisingly, the spin susceptibility is peaked at Q1 and Q2 in
the paramagnetic phase and in the antiferromagnetic phase the system orders antiferromagnetically at either of
three vectors. Incidentally, the antiferromagnet emerges through the formation and condensation of electron-hole
pairs (excitons), described by a BCS-type theory. The same excitonic instability is for example responsible for
the magnetism of chromium.
Assuming that the exchange of spin fulctuations in the paramagnetic phase is the main pairing interaction,
the gap equation
1 X 1 nF (Ek0 )
k = Vkk0 k0 (12.36)
N 0 2Ek0
k

with Vkk0 large and positive for k k0 = Q1 or Q2 , favors a gap function k changing sign between k and
k + Q1 and between k and k + Q2 , see Sec. 12.1. This is most easily accomodated by a nodeless gap changing
sign between the electron and hole pockets:
ky

kx

131
This gap is said to have s-wave symmetry in that it does not have lower symmetry than the lattice, unlike d-wave.
To emphasize the sign change, it is often called an s -wave gap. The simplest realization would be

k = 0 cos kx a cos ky a. (12.37)

12.4 Triplet superconductors and He-3


So far, we have assumed that Cooper pairs are formed by two electrons with opposite spin so that the total
spin of pair vanishes (spin-singlet pairing). This assumption becomes questionable in the presence of strong
ferromagnetic interactions, which favor parallel spin alignment. If superconductivity is possible at all in such a
situation, we could expect to find spin-1 Cooper pairs. Since they would be spin triplets, one is talking of triplet
superconductors. This scenario is very likely realized in Sr2 RuO4 (which is, interestingly, isostructural to the
prototypical cuprate La2 CuO4 ), a few organic salts, and some heavy-fermion compounds. It is even more certain
to be responsible for the superfluidity of He-3, where neutral He-3 atoms instead of charged electrons form Cooper
pairs, see Sec. 2.2.
Formally, we restrict ourselves to a BCS-type mean-field theory. We generalize the effective interaction to
allow for an arbitrary spin dependence,
1 X X
k ck ck + V 0 0 (k, k0 ) ck ck, ck0 , 0 ck0 0 ,
X
H= (12.38)
N 0 0 0
k kk

where , , 0 , 0 =, are spin indices. In decomposing the interaction, we now allow the averages hck, ck i to
be non-zero for all , . Thus the mean-field Hamiltonian reads
1 X X D E 
k ck ck + ck ck, ck0 , 0 ck0 0 + ck ck, hck0 , 0 ck0 0 i + const.
X
HMF = V 0 0 (k, k0 )
N 0
k kk 0 0
(12.39)
We define
1 XX
(k) := V 0 0 (k, k0 ) hck0 , 0 ck0 0 i (12.40)
N 0 0 0
k

so that
1 XX D E
(k) = V0 0 (k0 , k) ck0 0 ck0 , 0 . (12.41)
N 0 0 0
k

Here, we have used that 


V 0 0 (k, k0 ) = V0 0 (k0 , k),

(12.42)
which follows from hermiticity of the Hamiltonian H. Then

k ck ck (k) ck ck, + const.


X X X
HMF = (k) ck, ck (12.43)
k k k

The gap function (k) is now a matrix in spin space,


 
(k) (k)
(k) = . (12.44)
(k) (k)


The function (k) has an important symmetry property that follows from the symmetry of averages hck, ck i:
It is clear that
hck ck, i = hck, ck i. (12.45)
Furthermore, by relabeling , 0 0 , k k, k0 k0 in the interaction term of H, we see that the
interaction strength must satisfy the relation

V 0 0 (k, k0 ) = V 0 0 (k, k0 ). (12.46)

132
Thus
1 XX 1 XX
(k) = V 0 0 (k, k0 ) hck0 0 ck0 , 0 i = + V 0 0 (k, k0 ) hck0 , 0 ck0 0 i = (k)
N 0 0 0 N 0 0 0
k k
(12.47)
or, equivalently,

(k) T (k).
= (12.48)

We now want to write (k) in terms of singlet and triplet components. From elementary quantum theory, a
spin-singlet pair is created by
ck ck, ck ck,
sk := , (12.49)
2
while the m = 1, 0, 1 components of a spin-triplet pair are created by
tk1 := ck ck, , (12.50)
ck ck, + ck ck,
tk0 := , (12.51)
2
tk,1 := ck ck, , (12.52)
respectively. Alternatively, we can transform onto states with maximum spin along the x, y, and z axis. This is
analogous to the mapping from (l = 1, m) eigenstates onto px , py , and pz orbitals for the hydrogen atom. The
new components are created by

ck ck, + ck ck,
tkx := , (12.53)
2
ck ck, + ck ck,
tky := i , (12.54)
2
ck ck, + ck ck,
tkz := = tk0 , (12.55)
2
which form a vector tk . The term in HMF involving (k) can now be expressed in terms of the new operators,

tkx itky tkx itky


" #
X
X sk + tkz sk + tkz
(k)ck ck, = (k) + (k) + (k) + (k)
k k
2 2 2 2
!
X h
i
= 2 k + d(k) tk (12.56)
k

(the factor 2 is conventional), which requires
(k) (k)
k = , (12.57)
2
(k) + (k)
dx (k) = , (12.58)
2
(k) + (k)
dy (k) = i , (12.59)
2
(k) + (k)
dz (k) = . (12.60)
2
The term in HMF involving (k) is just the hermitian conjugate of the one considered. We have now identified
the singlet component of the gap, k , and the triplet components, k(k). Since
 
dx (k) + idy (k) k + dz (k)
(k) = , (12.61)
k + dz (k) dx (k) + idy (k)

133
we can write the gap matrix in a compact form as

(k) = (k 1 + d(k) ) i y , (12.62)

where is the vector of Pauli matrices. The mean-field Hamiltonian HMF is diagonalized by a Bogoliubov
transformation and the gap function (k) is obtained selfconsistently from a gap equation in complete analogy

to the singlet case discussed in Sec. 10.1, except that (k) is now a matrix and that we require four coefficients
uk , uk , vk , vk . We do not show this here explicitly.
A few remarks on the physics are in order, though. The symmetry (k) = T (k) implies

(k 1 + d(k) ) i y = i ( y )T k 1 + d(k) T

(12.63)
| {z }
i y
k 1 + d(k) = y
k 1 + d(k) T y


y x T y
( )
= k 1 + d(k) y ( y )T y
y ( z )T y
y x y

= k 1 + d(k) y y y
y z y
= k 1 d(k) . (12.64)

Thus we conclude that k is even,


k = k , (12.65)
whereas d(k) is odd,
d(k) = d(k). (12.66)
Hence, the d -vector can never be constant, unlike the single gap in Sec. 10.1. Furthermore, we can expand k
into even basis functions of rotations in real (and k) space,

k = s s (k) + dx2 y2 dx2 y2 (k) + d3z2 r2 d3z2 r2 (k) + dxy dxy (k) + dyz dyz (k) + dzx dzx (k) + . . . ,
(12.67)
and expand (the components of) d(k) into odd basis functions,

d(k) = dpx px (k) + dpy py (k) + dpz pz (k) + . . . (12.68)

For the singlet case we have already considered the basis functions s (k) = 1 for conventional superconductors,
s (k) = cos kx a cos ky a for the pnictides, and dx2 y2 (k) = cos kx a cos ky a for the cuprates. A typical basis
function for a triplet superconductor would be px = sin kx a. He-3 in the so-called B phase realized at now too
high pressures (see Sec. 2.2) has the d -vector

d(k) = x px (k) + y py (k) + z pz (k), (12.69)

where the basis functions are that standard expressions for px , py , and pz orbitals (note that there is no Brillouin
zone since He-3 is a liquid),
r
3
px (k) = sin k cos k , (12.70)
4
r
3
py (k) = sin k sin k , (12.71)
4
r
3
pz (k) = cos k . (12.72)
4
This is the so-called Balian-Werthamer state.

134
It is plausible that in a crystal the simultaneous presence of a non-vanishing k-even order parameter k and
a non-vanishing k-odd order parameter d(k) would break spatial inversion symmetry. Thus in an inversion-
symmetric crystal singlet and triplet superconductivity do not coexist. However, in crystals lacking inversion
symmetry (noncentrosymmetric crystals), singlet and triplet pairing can be realized at the same time. More-
over, one can show that spin-orbit coupling in a noncentrosymmetric superconductor mixes singlet and triplet
pairing. In this case, k and d(k) must both become non-zero simultaneously below Tc . Examples for such
superconductors are CePt3 Si, CeRhSi3 , and Y2 C3 .

135

You might also like